QUANTITATIVE HUMAN PHYSIOLOGY QUANTITATIVE HUMAN PHYSIOLOGY AN INTRODUCTION Second Edition Joseph Feher Department of Physiology and Biophysics, Virginia Commonwealth University School of Medicine AMSTERDAM • BOSTON • HEIDELBERG • LONDON • NEW YORK • OXFORD • PARIS SAN DIEGO • SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO Academic Press is an imprint of Elsevier Academic Press is an imprint of Elsevier 125 London Wall, London EC2Y 5AS, United Kingdom 525 B Street, Suite 1800, San Diego, CA 92101-4495, United States 50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom Copyright r 2012, 2017 Elsevier Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions. This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein). Notices Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary. Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility. To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress ISBN: 978-0-12-800883-6 For Information on all Academic Press publications visit our website at https://www.elsevier.com Publisher: Katey Birtcher Acquisition Editor: Steve Merken Editorial Project Manager: Nate McFadden Production Project Manager: Stalin Viswanathan Designer: Maria Ines Cruz Typeset by MPS Limited, Chennai, India Preface Welcome to the second edition of Quantitative Human Physiology! This new edition has been updated with numerous enhancements, many of which were suggested or inspired by instructors and students who used the first edition. These important changes include (but are not limited to): G G G G G G Substantial updating of the text throughout to reflect the latest research results, with many sections expanded to include relevant and important information Enhanced, updated, and improved figures for better understanding and clarification of challenging topics Addition of several new appendices covering statistics, nomenclature of transport carriers, and structural biology of important items such as the neuromuscular junction and calcium release unit Addition of new problems within the problem sets and example calculations in the text Addition of some Clinical Applications such as dual energy X-ray bone densitometry Addition of commentary to power point presentations. The goal of this new edition was to make important improvements while retaining the features that make this text uniquely suited to the needs of undergraduate bioengineering students. While it is sometimes very tempting to make drastic and sweeping changes in an attempt satisfy everyone, this new edition focuses on refinements and updates that, we hope, most instructors and students will find helpful, informative, and instructionally sound. THIS TEXT IS A PHYSIOLOGY TEXT FIRST, AND QUANTITATIVE SECOND The second edition of this text remains faithful to its primary goals: it remains, first and foremost, a physiology text. It is explicitly designed for a certain class of students, those majoring in Biomedical Engineering at Virginia Commonwealth University, and their suggestions, limitations, and desires have shaped the text from the outset. Specifically, the text is designed for students who have never been exposed to Physiology, students who know neither the language nor the concepts of the subject. The text contains all elements of physiology in nine units: physical and chemical foundations; cell physiology; excitable tissue physiology; neurophysiology; cardiovascular physiology; respiratory physiology; renal physiology; gastrointestinal physiology; and endocrinology. The course is best taught in the order of the text but it is possible to present the material in other sequences. Secondly, the text affirms its aim to be quantitative. Being quantitative has two aspects. The first is about knowing the numerical value for the ranges of crucial aspects of physiology, such as the flows or forces within the body. The second is about discovering the relations between physiological parameters. For many aspects of physiology, there currently is not enough information to make a detailed quantitative analysis, or the analysis at that depth is beyond the scope of this text. In these cases, this text is not very different from more traditional texts. Where possible, the text takes an analytical and quantitative approach. THE TEXT USES MATHEMATICS EXTENSIVELY Carl Frederick Gauss famously said, “Mathematics is the queen of sciences.” Mathematics is not just about the numerical value of something, such as the magnitude of the arterial blood pressure or the rate of salivary secretion, although that is what many people think of when they think of quantitation. Rather, mathematics is about the relationship between things that can vary with time or position. These relationships cannot be fully understood with words alone. They require the language of mathematics. Students should be able to articulate these relationships with words, but this text demands more. Mathematical statements—equations—are simply logical sentences. You can read an equation in words. But the mathematical statement uses an economy of words. Mathematics also has specific rules for the manipulation of the logical parts of the sentences, so that rearrangement or combination of equations leads to new insight. NOT ALL THINGS WORTH KNOWING ARE WORTH KNOWING WELL This text uses lots of mathematics at the level of the calculus and elementary differential equations. These mathematical tools are used to encode the physics or chemistry of processes into mathematical symbolism, and mathematical manipulation leads to useful equations. ix x PREFACE The point of the derivations is the useful equation, not the derivations themselves. The derivations are included, sometimes as an appendix, to satisfy the students that the final equations are not magical, but come from the application of mathematics to physical and chemical principles that apply to the body. The point is to be able to apply the final equations. Physics, chemistry, and math at the level of calculus are used to get the equations, but generally algebra is all that is required to apply them. In my view, the derivations are important to teach students the process of encoding the physics and chemistry and deriving the relationship between variables that constitute the useful equations, but rote memorization of derivations is boring and useless. PERFECT IS THE ENEMY OF GOOD: EQUATIONS AREN’T PERFECT, BUT THEY’RE OFTEN GOOD ENOUGH The text does not say what I tell my students about the applicability of equations in general. I tell my students in lecture that all of these equations are wrong. They are wrong in something of the same sense that Newtonian mechanics is wrong. Relativity and quantum mechanics supplant Newtonian mechanics (even in the macroscopic world) but Newtonian mechanics will still get the rocket ship to the moon. So I tell them that these equations are theoretically wrong but they give satisfactorily correct answers, in much the same way that Newtonian mechanics still does, even though theoretically incomplete. Fick’s Law of Diffusion depends on continuum mathematics for an inherently discrete process. But the discreteness is so finely divided that the distinction is unimportant. EXAMPLES AND PROBLEM SETS ALLOW APPLICATION OF THE USEFUL EQUATIONS There are several aids in the text to foster a quantitative and analytical understanding. First, there are some worked calculations in the text that are set apart in text boxes as Examples. These aim to show the students how to apply some of the ideas presented in the text. Second, there are a total of 17 problem sets scattered throughout the text. All units have at least one, most have two, and the cardiovascular section has three. These are meant to cover about three chapters each, so that a problem set can be assigned approximately once a week, for three lectures per week. Students have repeatedly told me that they want a solution set to the problems to see how they can do them, but this makes them useless as a graded assignment. There is no better teacher than wrestling with a problem. The second edition has expanded on these problem sets with new problems. The problems themselves are generally meant to cover some new idea or concept that could not be, or was not, effectively covered in the text itself, and to expand on the student’s understanding of the material. They are not merely busy work or “plug and chug” exercises. The alert student should not merely do the problem, but think about what the result means. In many cases, the problems are written as a sequence of questions, each of which sets up the student for further questions or insights. These illustrate the process necessary for answering a larger question, breaking it up into parts of the answer that must come first. This method aids the students in thinking about larger problems: break it down into its simpler components. Some subject matter unfortunately does not lend itself easily to such problems, but I have attempted to find problems that students can do. The students in my classes find the problems challenging. I allow them to work on them collaboratively, because they are meant to be part of the instruction and less of the evaluation, but the problem sets are graded and contribute significantly to the final grades. Of course, such policies are set at the discretion of the instructor. LEARNING OBJECTIVES, SUMMARIES, AND REVIEW QUESTIONS GUIDE STUDENT LEARNING The Learning Objectives are meant as a guide to the construction of examination questions, either directly or indirectly. These learning objectives are fairly broad and together they cover the breadth of physiology. This is my advice to students: read the Learning Objectives first, the chapter Summary second, and then read the text. Next, attempt to answer the Review Questions and return to the Learning Objectives. If you are mystified by any of the Review Questions or Learning Objectives, read the pertinent part of the text again. CLINICAL APPLICATIONS PIQUE INTEREST Pathological situations often illuminate normal physiology. Clinical Applications are scattered throughout the text. Because it takes a lot of background material to understand these clinical applications, Clinical Applications are less prevalent in the early parts of the text, which are foundational, than in the latter parts of the text. There are more than 50 such Clinical Applications, with several new additions in the second edition. The students like them because these Clinical Applications tell them that there is a reason for learning what might otherwise appear arcane. HOW INSTRUCTORS CAN USE THIS TEXT This text is meant for anyone with a fairly modest level of mathematical skill, at the level of the calculus with elementary differential equations. Students without this level of training will find this book too difficult. I have developed this book specifically with undergraduate Biomedical PR E FACE Engineering students in mind, and have taught this material for 16 years. Each Chapter is intended to be a single lecture, and the length of the chapter is meant to be readable in a single sitting. There are 77 chapters, so it is most useful in a two-semester sequence. I cover all chapters in that time, and all problem sets. The text would also be useful for instructors with less time by adjusting the breadth and depth of coverage. Units 1, 8, and 9, (Physical and Chemical Foundations, Gastrointestinal Physiology and Endocrinology, respectively) for example, could be eliminated or covered more superficially. Unit 1 could be eliminated because it is a review, Units 8 and 9 because they are relatively peripheral to Biomedical Engineering. Unit 2, Cell Physiology, could be covered more superficially or eliminated if students previously have had Cell Biology. Alternatively, it is possible to cover the breadth of physiology if the depth is reduced. This can be done by focusing on the chapter summaries, which give a broach picture of what is happening with less detail, and combining lectures or omitting some topics. In this regard, the text is useful similarly to a cafeteria: the instructor is free to choose those sections of most interest and to downplay those of lesser interest. Some instructors may feel that knowing how to think about problems is the most important thing, and that the details of the physiology are secondary. Such an instructor may want to focus more on the appendices than on the chapter matter themselves, and more on the problems and how to solve them. ANCILLARY MATERIALS FOR INSTRUCTORS For instructors adopting this text for use in a course, the following ancillaries are available: Power Point lecture slides, electronic images from the text, solutions manual for the problem sets, laboratory manuals with example data, and examination questions with answers. The Power Point slides have been updated to include the new figures and more commentary as appropriate for the lectures. Please visit https://textbooks.elsevier.com/ 9780128008836. HOW STUDENTS CAN USE THIS TEXT A student’s goal ought to be to learn as much physiology as possible within the constraints of available time. This text is written with considerable detail. The Learning Objectives and Review Questions set the stage for the kinds of things you should be able to do, and the kinds of questions you should be able to answer. The Chapter Summaries encapsulate each chapter in an economy of words and detail. Start with the Learning Objectives, read the Summary, and then look at the Review Questions. Then read the chapter and repeat Learning Objectives, Summary, and Review Questions. What you cannot answer at that point requires you to re-read the pertinent sections a second time. The approach to the problem sets is different. The first job is: find a bright fellow student to work with. Second, read the question for understanding of what it is asking. Then ask yourself, what is needed to answer this question? How can you find out what is needed? If it is a physical constant, where can you find it? Do you know a relationship or equation that relates what is being asked to what is given? Write it down and rearrange it to give the desired answer. Plug and chug what is given and what else you have looked up to get a numerical answer. It is very simple in principle but sometimes very difficult in execution. ANCILLARY MATERIALS FOR STUDENTS Student resources available with this text include a set of online flashcards, a selection of animations based on the figures in the text, and online quizzes for self-study. Please visit https://booksite.elsevier.com/9780128008836. STUDENT FEEDBACK Students who have completed the course regularly tell me that it is both one of the most challenging and one of the most rewarding courses of their undergraduate career. This is generally the case: what you get out of an educational enterprise is proportionate to what you put in. Physiology is an integrative science. There is great satisfaction in understanding how a system works from the cellular and subcellular level all the way up to the organism level. Human beings do not come with an owner’s manual. The idea that this text is a first approximation to an owner’s manual resonates with the students. Joseph Feher xi Acknowledgments I would like to thank all the reviewers of the proposal and drafts of this project. Their feedback was very helpful in improving the final version: Brett BuSha Jingjiao Guan Nuran M. Kumbaraci Marie Luby Ken Yoshida Ryan Zurakowski The College of New Jersey Florida State University Stevens Institute of Technology University of Connecticut Indiana University—Purdue University Indianapolis University of Delaware I would also like to acknowledge the help and support of my colleagues at the VCU School of Medicine, especially Clive Baumgarten and Ray Witorsch, for their criticism of early drafts, and Margaret Biber, who expressed the confidence in publishing this text that kept me going. I would like to thank George Ford, VCU School of Medicine, who helped develop the early outlines of the cardiovascular section; Scott Walsh, who provided the basis for some of the endocrine chapters; and Andy Anderson, who helped critique multiple teaching efforts of mine. I especially acknowledge the feedback from many years of sophomore and junior BME students who pointed out errors and difficulties in the material and suggested improvements to help their learning experience, which is really what this text is all about. Special thanks go to students Woon Chow, Yasha Mohajer, Matthew Caldwell, Matthew Painter, Mary Beth Bird, Richard Boe, Linda Scheider, William Eggleston, Alex Sherwood, Ross Pollansch, Kate Proffitt, and Roshan George. Teaching these students and many more like them, and getting to know them, has been tremendously rewarding. I would like to thank the publishing team at Elsevier including Katey Birtcher, Publisher; Steve Merken, Senior Acquisitions Editor; Nate McFadden, Senior Developmental Editor; Maria Inês Cruz, Cover Designer; and Stalin Viswanathan, Production Project Manager. Much more indirectly, I wish to thank a long line of scientific mentors who instilled in me the academic integrity and the desire to be thorough and correct. Included in this list are Lemuel Wright, and Donald B. McCormick, of Cornell University, who advised me on my master’s thesis; Robert Hall of Upstate Medical Center, with whom I worked for a short but meaningful time and who provided the epiphany for my pursuit of an academic career; Robert Wasserman, of Cornell University, who first inspired me to apply mathematics to transport processes; Norm Briggs, of VCU School of Medicine, who first thought I would be a good bet for a tenure track position; Don Mikulecky, who introduced me to many ideas of theoretical biology; and Margaret Biber, who first presented me with the daunting task of developing a year-long physiology sequence for BME students, with laboratories, that formed the text. Although I desire to be thorough and correct, often I am mistaken. I acknowledge the help of all those listed above, but reserve to myself the blame for any errors in the text. If the reader finds any errors of fact or analysis, I would appreciate them letting me know so that I can evaluate and correct it. Lastly, I express my deep appreciation for my wife, Lee, who put up with innumerable late nights and weekends with a husband glued to the chair in front of the computer screen. Her support and encouragement made the work possible. The first edition of this text was dedicated to the memory of Karen Esterline Feher, who died tragically during the writing of the first edition. The second edition was written shortly after the death of Mildred Hastings Feher, who was a most gracious human being. Most of it was also written while my daughter, Teresa, gave birth to three boys. These four ladies in my life have made me understand more fully the circle of life and death, and have taught me to value all parts of it. They are the bringers of life and I can only hope to preserve and protect it, and to finally let it go. I dedicate this effort to Mildred, Lee, Karen and Teresa. xiii The Core Principles of Physiology Learning Objectives G G G G G G G G G G G G Define the discipline of physiology Describe in general terms how each organ system contributes to homeostasis Define reductionism and compare it to holism Describe what is meant by emergent properties Define homeostasis List the four Aristotelian causes and define teleology Define mechanism Describe how evolution is a cause of human form and function Write equations for the conservation of mass and energy for the body Give an example of signaling at the organ or cellular level of organization List the core principles of physiology Contrast feed-forward or anticipatory control to negative feedback control HUMAN PHYSIOLOGY IS THE INTEGRATED STUDY OF THE NORMAL FUNCTION OF THE HUMAN BODY ORGAN SYSTEMS WORK TOGETHER TO PRODUCE OVERALL BODY BEHAVIOR Almost any explanation of something begins with a description of that something’s parts. The human body consists of many parts. We consider an assortment of parts that usually relate to each other in defined ways to be a system. In physiology, a system is usually considered to be a group of organs that serve some welldefined function of the body. The parts of these systems can be described separately, but they work together to produce the overall system behavior. That is, the individual behavior of the parts is integrated to produce overall behavior. The various organ systems and their functions in the body are summarized in Table 1.1.1. These systems are further integrated to produce overall bodily function. Physiology is the study of the integrated normal function of the human body. Each of these organ systems is essential to the survival of the organism, the living human being. It is possible © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00001-X 1.1 in an artificial environment to survive with a single compromised system—such as persons with failed kidneys or failed immunological systems—but these persons could not survive in natural ecosystems. REDUCTIONISM EXPLAINS SOMETHING ON THE BASIS OF ITS PARTS The process of explaining something on the basis of its parts is called reductionism. Thus the behavior of the body can be explained by the coordinated behavior of its component organ systems. In turn, each organ system can be explained in terms of the behavior of the component organs. In this reduction recursion, the behavior of the component organs can be explained by their components, the individual cells that make up the organ. These cells, in turn, can be explained by the behavior of their component subcellular organelles; the subcellular organelles can be explained by the macrochemicals and biochemicals that make up these organelles; the biochemicals can be explained by their component atoms; the atoms can be explained by their component subatomic particles; the subatomic particles can be explained by fundamental particles. According to this reductionist recursion, we might anticipate that the final explanation of our own bodies lies in the physics of the fundamental particles. Beyond being impractical, there is a growing realization that it is theoretically impossible to describe complex and complicated living beings solely on this basis of fundamental physics, because at each step in the process some information is lost. This situation should not trouble us too much. In physics, we speak of the force of friction even though friction is not a fundamental force. It results from tremendous numbers of electromagnetic interactions between particles on two surfaces so that friction really is just a trace of all of those microscopic electromagnetic interactions onto the macroscopic bodies. The details of the surfaces produce those forces and we can experimentally reduce the force simply by polishing the surfaces that interact, making the tiny bumps and valleys on the surfaces smaller. We don’t know the details of the surface and so we ignore the reality and lump all of those interactions together and call it the “force” of friction. We have lost some information and abandon the idea of recursing reductionism down to the level of fundamental physics. 3 4 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION parts, the molecules that make it up, and how the organelle’s function is regulated by and contributes to the function of the cell. TABLE 1.1.1 Organ Systems of the Body Organ System Function Nervous system/ endocrine system Sensory input and integration; command and control Musculoskeletal system Support and movement Cardiovascular system Transportation between tissues and environmental interfaces Gastrointestinal system Digestion of food and absorption of nutrients Respiratory system Regulation of blood gases and exchange of gas with the air Renal system Regulation of volume and composition of body fluids Integumentary system (skin) Protection from microbial invasion, water vapor barrier, and temperature control Reproductive system Pass life on to the next generation Immune system Removal of microbes and other foreign materials Level Discipline Universe Cosmology Society Sociology Body Explanation Physiological systems Function Physiology Organs Cells and cell products Cell physiology Subcellular organelles Biochemistry Molecules Chemistry Atoms Subatomic particles High-energy physics FIGURE 1.1.1 Hierarchical description of physical reality as applied to physiological systems. We attempt to “explain” something in terms of its component parts and describe a function for a part in terms of its role in the “higher” organizational entity. PHYSIOLOGICAL SYSTEMS ARE PART OF A HIERARCHY OF LEVELS OF ORGANIZATION The recursion of explanation described above for reductionism involves various levels of complexity in a hierarchical description of living beings, as shown in Figure 1.1.1. Understanding any particular level entails relating that level to the one immediately above it and the one immediately below it. For example, scientists studying a particular subcellular organelle can be said to have mastered it when they can explain how the function of the organelle derives from the activities of its REDUCTIONISM IS AN EXPERIMENTAL PROCEDURE; RECONSTITUTION IS A THEORETICAL PROCEDURE The processes used in going “down” or “up” in this hierarchy are not the same. We use reductionism to explain the function of the whole in terms of its parts, by going “down” in the levels of organization. We describe the function of the parts at one level by showing how they contribute to the behavior of the larger level of organization, going “up” in Figure 1.1.1. These processes are fundamentally different. Reductionism involves actually breaking the system into its parts and studying the parts’ behavior in isolation under controlled conditions. For example, we can take a sample of tissue and disrupt its cells so that the cell membranes are ruptured. We can then isolate various subcellular organelles and study their behavior. This procedure characterizes the behavior of the subcellular organelle. Knowing the behavior of the individual parts and paying close attention to how these parts are connected, it is possible to predict system behavior from the parts’ behavior using simulation or other techniques. Because it is impossible, except in rare and limited cases, to reassemble broken systems (we cannot unscramble the egg!), we must test our ideas of subcellular function theoretically. HOLISM PROPOSES THAT THE BEHAVIOR OF THE PARTS IS ALTERED BY THEIR CONTEXT IN THE WHOLE Holism conveys the idea that the parts of an organism are interconnected and that each part affects others. The parts cannot be studied in isolation because important aspects of their behavior depend solely on their interaction with other parts. Reductionism seems to imply that the whole is the sum of the parts, whereas holism suggests that the whole is greater than the sum of the parts. The system takes on new properties, emergent properties, that arise from complex interactions among the parts. Examples of emergent properties include selfreplication. The ability of cells to form daughter cells is a system property that does not belong to any one part but belongs to the entire system. Consciousness is also an emergent property that arises from neuronal function but at a much higher level of organization. PHYSIOLOGICAL SYSTEMS OPERATE AT MULTIPLE LEVELS OF ORGANIZATION SIMULTANEOUSLY It should be clear from Figure 1.1.1 that all the levels of organization simultaneously operate in the living human being. Processes occur at the molecular, subcellular, cellular, organ, and system levels simultaneously and dynamically. The C ore P rinciples of Physiology THE BODY CONSISTS OF CAUSAL MECHANISMS THAT OBEY THE LAWS OF PHYSICS AND CHEMISTRY When we say that we explain something, usually we mean that we can trace the output of the system—its behavior—to some cause. That is, we can seamlessly trace cause and effect from some starting point to some ending point. Aristotle (384322 BC) posited four different kinds of causality: 1. Material Cause A house is a house because of the boards, nails, shingles, and so on that make it up. We are what we are because of the cells and the cell products that make us up. 2. Efficient Cause A house is a house because of the laborers who assembled the materials to make the house. We are what we are because of the developmental processes that produced us and because of all of the experiences we have had that alter us. 3. Formal Cause A house is a house because of the blueprint that directed the laborers to assemble the materials in a particular way. We are what we are because of the DNA that directs our cells to make some proteins and not others, and because of epigenetic effects—those effects resulting from the environment interacting with the genome. 4. Final Cause A house is a house because someone needed shelter. We are what we are because. . . The final cause for humans has a variety of possible answers. This is the only cause that addresses the idea of a purpose. We make a house for a purpose: to provide someone with shelter. What is the purpose of human beings? This cause asks the question of why rather than how. TELEOLOGY IS AN EXPLANATION IN REFERENCE TO A FINAL CAUSE A description or explanation of a system based on the reference to the final cause is called a teleological explanation. Teleology has long been ridiculed by scientists because it appears to reverse the scientific notion of cause and effect. In normal usage, cause-and-effect linkages describe only the efficient cause. When a force acts on something, that something reacts in a predictable way. Its predictability is encoded in physical law. Teleology describes the behavior in terms of its final purpose, and not its driving force, which reverses the cause-andeffect link. HOW? QUESTIONS ADDRESS CAUSAL LINKAGES. WHY? QUESTIONS ADDRESS FUNCTION To clarify this process, let us ask some questions about blood pressure. How does your body regulate the average aterial blood pressure? This answer takes some time to develop fully (see Unit 5) but we can simplify the answer by saying: by increasing or decreasing the caliber of the arteries, by increasing or decreasing the output of the heart, and by increasing or decreasing the volume of fluid in the arteries. All of these actions—and more— interact to determine the arterial blood pressure. All of these parts to the answer involve actions, and actions are efficient causes. For example, increasing or decreasing the caliber of the arteries depends on a complicated network of signaling pathways and biochemical reactions, but each starts with a cause and ends with an effect. Now we ask: why does your body regulate average aterial pressure at the level it does? The answer is again not so simple but we can simplify it as: to assure perfusion of the tissues. This addresses a purpose. In this sense, it is teleological, but it also makes sense to us. It is not that the cardiovascular system knows what it is doing, but its operation has been selected to produce the desired output. Within the framework of the body, organs do serve a purpose and that purpose is their function. All of the functions listed in Table 1.1.1 are final causes for the organ systems. HUMAN BEINGS ARE NOT MACHINES BUT STILL OBEY PHYSICAL LAW Aristotle had a different idea about the mind. He posited that the mind was not a material entity, much like an idea is not a material thing. He posited that the mind was what perceived, imagined, thought, emoted, desired, felt pain, reasoned, remembered, and controlled the body. This philosophy in which the mind controls behavior is called mentalism. These ideas went largely unchallenged until the Renaissance. René Descartes (15961650) wrote a book, Treatise on Man, in which he tried to explain how the nonmaterial mind might interact with the material body. In this process, he constructed mechanical analogues to explain sensation and command of movement. He thought that the operation of all things, however complex, could be explained by some mechanism. Each mechanism consists of a sequence of events that link an initial causal input to an effect that becomes the cause of the next step in the mechanism. In this view, human beings are very complex machines that obey natural law. In the late 1800s, W.O. Atwater (18441907) built a calorimeter to study heat production, gas exchange, and fuel consumption in humans. He found that the energy output of humans matched the chemical energy of the food consumed, within narrow experimental error. This result confirmed that the law of conservation of energy held for the transformation of energy by the human body as well as for inanimate transformations. IS THERE A GHOST IN THE MACHINE? The core principle of physiology that states the human body is a mechanism strikes at the heart of the concept of vitalism. Vitalism states that living things cannot be described in mechanistic terms alone, and that some organizing force or vital principle forever distinguishes living things from nonliving things. For human beings, we could call this the soul and be in reasonable agreement with the vitalists. So far, science has found 5 6 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION no reliable, scientific verification that the human body violates any physical law. The existence of emergent properties gives credence to the idea of vitalism. Emergent properties are new properties that arise from complex systems because of their complexity and topology—the way that subsystems are connected. These emergent properties do not appear to be predictable. That is, we cannot see how we would predict their emergence given the fundamental laws of physics. Examples of these emergent properties are life itself and consciousness. These properties appear to arise from interactions of parts that appear to obey physical law alone, but how these properties arise remains mysterious. These emergent properties are system properties, not mechanism properties. Because of overwhelming evidence that specific deficits in brain function produce specific deficits in mental function, we have come to believe that the brain somehow produces the mystical thing that is conscious and self-aware. This thing is not material in the ordinary sense of the word, just like an idea is not a material thing. The new mindbody problem is the inverse of Descarte’s mindbody problem: how can a material thing (the brain) produce the nonmaterial thing that we identify as self. Is this relationship a one-way street, or does the mind have a reciprocal effect on the brain? Although these are extremely important questions, most physiologists take a narrower aim of explaining only those phenomena for which we have satisfactory mechanistic models and attempt to extend the range to include all physiologic phenomena, including consciousness. UNDERSTANDING A SYSTEM IS EQUIVALENT TO MAKING A MODEL OF IT What we have been discussing so far is what it means to say that we understand something. The something we are trying to understand we will label “the real system.” To understand this system, we make a model of it in a process that we will call encoding (B in Figure 1.1.2). The model does not have to remain solely in our minds. It could be written down as a set of equations or as a computer algorithm, for example. In the real system, perturbations cause changes in the real system. This causal link (A in Figure 1.1.2) between perturbation and effect is what we desire to predict or explain using the model. In the model, the cause is also encoded, and its predicted effect we will call an inference (C in Figure 1.1.2). That is, some perturbation of the model is predicted to cause some effect in the model. When we decode this effect (D in Figure 1.1.2), the inference of the model is translated as our prediction of the behavior of the real system. A correct model is one that correctly predicts the behavior of the real system. That is ½1:1:1 A5B1C1D THE CORE PRINCIPLES OF PHYSIOLOGY The rest of this chapter will discuss several Core Principles of Physiology. These are: G G G G G G Cells are the organizational unit of Life. Homeostasis is a central theme of physiology. We have evolved from prior life forms and our pedigree is revealed in our genome. Physiological systems transform matter and energy while obeying the conservation laws. Coordinated command and control requires signaling at all levels of organization. Control systems using negative feedback, positive feedback, anticipatory and threshold mechanisms. CELLS ARE THE ORGANIZATIONAL UNIT OF LIFE THE CELL THEORY IS A UNIFYING PRINCIPLE OF BIOLOGY The cell theory states that all biological organisms are composed of cells; cells are the unit of life and all life come from preexisting life. The cell theory is so established today that it forms one of the unifying principles of biology. The word cell was first used by Robert Hooke (16351703) when he looked at cork with a simple microscope and found what appeared to be blocks of material making up the cork. The term today describes a microscopic unit of life that separates itself from its surroundings by a thin partition, the cell membrane. D: Decoding step FIGURE 1.1.2 The modeling relation. The real system is characterized by causal links, A, that correspond to the behavior of the real system to external or internal perturbation. The real system is converted by an encoding step to a model (B). In the model, the response of the model to a perturbation is determined as an inference (C). The behavior of the model is then decoded (D) in the last step in testing the validity of the model. When B 1 C 1 D 5 A, the model has successfully predicted the behavior of the real system and we say that it is a valid model. Adapted from R. Rosen, Theoretical Biology and Complexity, 1985, Academic Press, NY. Perturbation A: Causal link Effect Real system Effect Model C: Inference Perturbation B: Encoding step The C ore P rinciples of Physiology Erythrocyte Leukocyte Eosinophil Inner hair cell Hepatocyte Enterocyte Parietal cell Motor neuron Cardiomyocyte FIGURE 1.1.3 Examples of the different cells that populate the human body. Motor neurons such as the one illustrated are found in the ventral horn of the spinal cord. The inner hair cells are found in the cochlea and form part of our response to sound. Erythrocytes, leukocytes, and eosinophils are all found in the blood. Cells typically are not colored but may be seen in color by their adsorption of histological stains. Hepatocytes in the liver help package nutrients, form bile, and detoxify foreign chemicals. The enterocytes line the small intestine and absorb nutrients from the food into the blood. The parietal cells secrete HCl in the stomach. The cardiomyocyte shown is a ventricular cell whose contraction contributes to the pumping action of the heart. This is a small sampling of the diversity of cell forms in the human body. Most biologists believe that life arose spontaneously from inanimate matter, but the details of how this could have happened remain unknown, and the time scale was long. Rudolf Virchow, a German pathologist (18211902), famously wrote “omnis cellula e cellula”—all cells come from other cells—meaning that spontaneous generation of living things from inanimate matter does not occur over periods as short as our lifetimes. CELLS WITHIN THE BODY SHOW A MULTITUDE OF FORMS Large multicellular organisms such as ourselves consist of a vast number of different cells that share some features but vary in size, structure, biochemical makeup, and functions. A sampling of the spectrum of cells that make up the body is illustrated in Figure 1.1.3. Almost all cells in the body have a cell membrane, also called the plasma membrane, and most contain a nucleus. The simplest cell in the human is probably the erythrocyte, which is the only cell in the body that lacks a nucleus. THE DIVERSITY OF CELLS IN THE BODY DERIVES FROM DIFFERENTIAL EXPRESSION OF THE GENOME The outward appearance and behavior of an organism define its phenotype, which is related to but not identical to the organism’s genetic material, its genotype. The genotype consists of the set of alternate forms of genes, called alleles, that the organism has, and these alternate forms of genes are further defined by the sequence of nucleotides in their DNA. DNA is the genetic material that is passed on through the generations. It determines the kind of proteins that cells can produce, and these materials make up the phenotype. The genome is the entirety of the hereditary information, including all of the genes and regions of the DNA that are not involved in producing proteins. Nearly all cells in the body contain the entire genome. The exceptions include the erythrocytes and the reproductive cells. Those cells that are not reproductive cells are called somatic cells (from the Greek soma, meaning body). Thus the great majority of body cells are somatic cells, and they all contain the same amount and kind of DNA. The astounding diversity of the types of human cells derives from their expression of different parts of the genome. Here expression means using DNA to produce proteins. THE CONCEPT OF HOMEOSTASIS IS A CENTRAL THEME OF PHYSIOLOGY EXTRACELLULAR FLUID SURROUNDS ALL SOMATIC CELLS As described above, each cell in the body is surrounded by a cell membrane that defines the limits of the cell and separates the interior of the cell from its exterior. The interior consists of a number of subcellular organelles suspended in a fluid, the intracellular fluid. The exterior consists of an extracellular matrix that holds things in place and an extracellular fluid. The extracellular fluid has two components: the plasma and the interstitial fluid. The plasma is that part of the extracellular fluid that is contained in the blood vessels. The interstitial fluid is that part of the extracellular fluid between the cells and the walls of the vasculature. Nearly all cells of the body come in intimate contact 7 8 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Red blood cells Body cells Plasma Intracellular fluid Nutrients EVOLUTION RESULTS FROM CAUSE AND EFFECT SUMMED OVER LONGTIME PERIODS Wastes Capillary walls would shift to those better suited to the environment. New variations in the genotype arise by mutation. Over geological time, such natural selection gradually changes the population. Darwin believed that such slow changes in the genetic makeup of populations could eventually produce new species, and he termed this slow formation of new species evolution. Interstitial fluid FIGURE 1.1.4 Relationship between cells and the extracellular fluid. All cells of the body are surrounded by a thin layer of extracellular fluid from which they immediately derive nutrients such as amino acids, sugars, and oxygen, and to which they discharge wastes such as carbon dioxide and other end products of metabolism. Nutrients are delivered to the cells and waste products are removed through the circulation, which does not make direct contact with the interstitial fluid, but is separated from it by the walls of the vascular system. with the extracellular fluid. The last step in the delivery of nutrients and the first step in removal of wastes is achieved through the extracellular fluid (see Figure 1.1.4). The extracellular fluid was called the milieu interieur, or the internal environment, by the great French physiologist, Claude Bernard (18131878). Survival of the cells depends on the maintenance of a constant internal environment. The maintenance of a constant internal environment is called homeostasis, which is literally translated as same standing. Contributing to the maintenance of a constant internal environment appears to the “goal” (or final cause or purpose) of many physiological systems, and this homeostasis is the central theme of physiology. EVOLUTION IS AN EFFICIENT CAUSE OF THE HUMAN BODY WORKING OVER LONGTIME SCALES EVOLUTION WAS POSTULATED TO EXPLAIN THE DIVERSITY OF LIFE FORMS Charles Darwin (18091892) wrote On The Origin of Species in 1858 as his attempt to explain the origin of the tremendous variety of animals and plants in today’s ecosystems. He noted that any one species consists of a population of individuals that are capable of breeding among themselves but not with members of other species. The similarity among members of a species defines the species; the differences between them define the individual. These outward appearances constitute the phenotype, as described earlier, which arises from the response of the genotype to the environment. Some of the individual members of a species are better suited to their environment than others and produce more offspring as a consequence. With sufficient time, the frequency of genotypes represented in the population Evolution is like a higher level on the hierarchy of cause-and-effect relationships. As an example, consider a mutation that alters the structure of a critical protein located in a selected group of cells in the body that enhances the function of these cells. The mutation causes an altered protein, which in turn causes enhanced behavior of the organism. This altered behavior of the organism causes greater success in reproduction. Over time, greater success in reproduction replaces the less-fit genotype with the mutated, superior genotype. Thus evolution results from thousands of independent cause-and-effect linkages played out over a population of individuals, over longtime periods. EVOLUTION WORKS ON PREEXISTING FORMS: COMPARATIVE GENOMICS REVEALS PEDIGREE At some time in the distant past, there was no life on earth. The origin of life is unknown and, in some sense, how it arose is not a scientific question because we cannot test any hypothesis of events in the past. We can, however, search for the trace of past events in the world today, much like a detective searches for clues to determine what happened earlier. This search has some of the character of an experiment. In this way, the fossil record illuminates the march of evolution to the present day. Similarly, we carry traces of our evolution in our own genome in the form of “fossil genes.” Because evolution works on preexisting forms, and because the multicellular organism plan entails the same challenges to homeostasis, and the same problems of cell maintenance, the genomes for many diverse animals and plants share profound similarities. For this reason, similarities in the genome can be used to trace the evolution of the proteins and shed light on the pedigree of species. EVOLUTION TAILORS THE PHENOTYPE TO THE ECOSYSTEM Humans live and reproduce within the context of an ecosystem. Our evolution has occurred because of our fit, or lack of it, with a specific environment. This explains some of the diversity of human forms within our species. Skin color and overall body shape, for example, are adaptations that arose to better fit the different levels of sunlight and air temperatures at different latitudes. Evolution has prepared us to meet the challenges of our environment but has not prepared us for unusual challenges. For example, we are adapted to survive short periods without water or longer periods without food, but we cannot do without air even for short periods. The C ore P rinciples of Physiology ROBUSTNESS MEASURES THE ABILITY OF THE BODY TO RESPOND TO ENVIRONMENTAL CHALLENGES A robust system is one that continues to function even when faced with difficult challenges. A fragile system fails easily. Engineered systems aim for a degree of robustness and usually achieve that end by adding redundant or back-up control systems and by building in safety factors. These systems are robust for some kinds of failure while remaining vulnerable to others. For example, autopilot systems in aircraft use multiple computers with different programs so that failure of one does not cause failure of the entire system. But these systems remain vulnerable to general power failure. We also have redundant or reserve function for several physiological systems. We have two kidneys, yet generally we can survive with only one. Liver, intestine, brain, and heart have more functional capacity than generally used so that we can survive if part of these organs fails. Strokes and heart attacks damage parts of the brain or the heart, respectively. Persons who suffer these cardiovascular accidents often recover much of their function, depending on the degree of damage and its location. If the damage is severe or involves a critical area, the victim may be permanently impaired or they may die. History is replete with astonishing stories of the tenacity of humans for life under amazingly harsh conditions. On the other hand, history also tells of the crumbling of civilizations when exposed to unfamiliar pathogens. Thus the human body may surprise us either because of its robustness or its fragility. REGULATION OF THE GENOME MAY EXPLAIN THE FAST PACE OF EVOLUTION There is a growing realization among evolutionary biologists that mutations in the genes that encode for somatic proteins—the ones that make us up—are only a small part of the story and cannot account for the rapid pace of evolution. Instead, much of evolution is accounted for in the genes that regulate the expression of other genes. Many modern birds, for example, do not have teeth. Yet it is possible experimentally to induce birds to make teeth, because they retain the genes for making teeth but also have genes that suppress the expression of the genes for teeth. Many diverse groups of animals share most of the genes involved in body building but differ in how and when these genes are used. The result is the differing body forms that are found in the animal kingdom. EVOLUTION HELPS LITTLE IN EXPLAINING THE NORMAL FUNCTION OF THE BODY Although evolution is one of the few unifying principles of biology and is one cause of the structure and function of the human body, it answers the question of the efficient cause of the body on a different time scale than the normal operating time scale of the body. Evolution does not aid us very much when trying to explain how our bodies work on a minute-to-minute basis. Instead, we look to control theory to explain the normal functioning of the human body. LIVING BEINGS TRANSFORM ENERGY AND MATTER Repair, maintenance, growth, activity, and reproduction all require input of energy and matter in the form of food. The gastrointestinal system breaks down the food, which is absorbed into the blood and distributed among the tissues according to need. The available building blocks must be transformed into cellular or extracellular components, and all of this metabolism requires energy. The energy comes from the oxidation of food and subsequent production of wastes. In addition to this conversion of chemical energy of one compound to another, we also transform chemical energy into other forms of energy, including electrical and mechanical energy. These processes obey physical and chemical laws that govern the transformation of energy and matter. In particular, we can write two equations that describe overall mass and energy balance in the body: Min 5 Mout 1 ΔMbody Mfood 1 Mdrink 1 Minspiredair 5 Mfeces 1 Murine 1 Mexpiredair 1 Mexfoliation 1 ΔMbody ½1:1:2 where M indicates mass and the subscripts indicate the origin of the mass (in 5 input; out 5 output; most of these are self-explanatory) and ΔMbody indicates change in body mass. These equations describe mass balance and simply indicate that all of the mass that enters the body must either stay there (ΔMbody) or exit the body through one of several routes. A similar equation can be written for energy balance: Ein 5 Eout 1 ΔEbody Efood 1 Edrink 1 Einspiredair 5 Efeces 1 Eurine 1 Eexpiredair 1 Eexfoliation 1 ΔEbody 1 Eheat 1 Ework ½1:1:3 The overall mass and energy balance is shown schematically in Figure 1.1.5. FUNCTION FOLLOWS FORM Almost all processes carried out by the body, at all levels of organization, depend on the three-dimensional structure of some component. The structure both 9 10 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Air system level at which the structure and arrangement of nerves and tissues is vital for the proper coordination of activity such as the heart beat or gastrointestinal motility. As another example, both the lungs and the gastrointestinal tract involve transfer of gas or nutrients from the environment to the blood. Both lungs and intestine have enormous surface areas and thin barriers—consequences of their structure—to maximize the rate of transport. Food and drink Lungs Skin COORDINATED COMMAND AND CONTROL REQUIRES SIGNALING AT ALL LEVELS OF ORGANIZATION Gastrointestinal system Heart Nutrients and water Internal work Kidney Heat Work Fat Tissues Energy released in metabolism Feces Urine FIGURE 1.1.5 Overall mass and energy balance in the body. The lighter arrows indicate energy transfer. The black arrows denote transfer of mass. The chemical energy of the ingested food is released by oxidation in metabolism, as indicated by the starburst in the tissues. This released chemical energy is used for internal work, which usually eventually degrades to heat, for external work and for storage in the chemical energy of body components such as glycogen and fat. Growth also entails a form of storage of ingested mass and chemical energy. The laws of conservation of mass and energy (in ordinary chemical reactions) require that the matter and energy that enter the body must be equal to the matter and energy that leave the body plus any change in the matter and energy content of the body. enables function and constrains it, by determining what can be done and how fast it can be accomplished. These structural considerations apply at the molecular level, at which the three-dimensional shape of protein surfaces determines what binds to the protein, how it is chemically altered, and how it interacts with other surfaces. These structural considerations apply at the subcellular level, at which the organelles themselves can compartmentalize chemicals and so determine or limit rates of reactions by regulating transfer between the compartments. Structural considerations are also important at the tissue level, at which the topology or spatial distribution of cellular processes allows countercurrent flows, for example, that are crucial in clearing metabolites from the blood or concentrating the urine. Structural considerations are important at the organ Success of an organism requires adaptive responses to change in the environment. This, in turn, requires sensory apparatus that senses both the external environment (exteroreceptors) and the interior environment (interoreceptors). These originate signals that pass either to nearby cells or to the central nervous system either for specific, reflex responses or for global responses. These signals are important at all levels of organization. At the subcellular level, these signals regulate the activities of subcellular components such as the expression of specific genes or the regulation of the rates of energy transformation. At the tissue level, local signals can regulate smooth muscle contraction to regulate blood flow within the organ or secretion into ducts; at the organ system level, signals traveling through the blood (hormones) or over nerves can coordinate activity of the system. At the whole organism level, signals at all levels must be used to adapt to whole-body responses such as running to avoid predators. Coordinating command and control for muscle contraction using sensory information from the environment (exteroreceptors) and from the muscle (interoreceptors) is illustrated in Figure 1.1.6. Signaling at the cellular level is illustrated schematically in Figure 1.1.7. MANY CONTROL SYSTEMS OF THE BODY USE NEGATIVE FEEDBACK LOOPS One of the main themes of physiological control systems comprises a negative feedback loop. This consists of controlled parameters such as plasma calcium concentration, body temperature, plasma glucose concentration, and plasma pH, a sensor for that parameter, a comparator, and an effector. For many physiologically controlled parameters, there is a set point or reference (see Figure 1.1.8). This is the desired value for the controlled parameter. Its value can change under some circumstances. When the controlled parameter varies from its set point, the variation is detected by the sensor and comparator. The comparator then engages some effector or actuator mechanism to correct the departure of the parameter from its normal, set-point value. An example of this is core body temperature, whose normal setpoint value is about 37 C. Whenever heat loss exceeds heat production, body temperature falls below the set point and the person shivers. In this case, the sensor The C ore P rinciples of Physiology POSITIVE FEEDBACK CONTROL SYSTEMS HAVE DIFFERENT SIGNS FOR THE ADJUSTMENT TO PERTURBATIONS Central nervous system Command signal Exteroreceptor sensory signal Skin Positive feedback systems exist for the mechanism of blood clotting, parts of the menstrual cycle, aspects of the action potential, and for parturition (child birth). In these cases, the disturbance is followed by an adjustment in the same direction of the disturbance, so that there is a rapid increase in some component. These positive feedback systems generally are self-limiting and, after the rapid increase in some component, there is a gradual return to baseline levels. ANTICIPATORY OR FEED-FORWARD CONTROL AVOIDS WIDE SWINGS IN CONTROLLED PARAMETERS Muscle Interoreceptor sensory signal Motor signal Spinal cord FIGURE 1.1.6 Neural signaling in the control of muscle. Neurons consist of cell bodies (dark circles) that have long processes (black lines) that bring signals into the central nervous system (dark lines with arrows) or take signals out toward the periphery (light lines with arrows). Branches at the ends of the long processes signify the junction of one neuron with another or with muscle. Neural signal transmission across these junctions is discussed in Chapter 4.2. Muscles are controlled by motor neurons whose cell bodies lie in the spinal cord. These can be activated in reflexes initiated by exteroreceptors that sense perturbations on the skin and send signals to the spinal cord and eventually activate the motor neurons by a simple reflex involving just a few interneurons in the spinal cord. Muscles can also be activated by another reflex involving a stretch receptor internal to the muscle (interoreceptor). In a third pathway, motor neurons can be activated by command signals originating in the brain. Nervous control of muscle is considered in Chapters 4.4 and 4.5. detects the temperature, the comparator determines that it has fallen below the set point, and it engages the skeletal muscles as an effector to produce heat by shivering to help raise the temperature back to the set point. In a fever, the set point is elevated and the individual feels chilled even when the temperature is elevated to, say, 40 C. When the fever “breaks,” the set point is reset back to 37 C and the person perspires because now the body temperature is elevated above the set point. When the controlled parameter varies from its set point, the variation is detected by the comparator. The comparator then engages some actuator mechanism to correct the departure of the parameter from its normal, set-point value. The controlled parameter can vary from its set point by the action of some physiological disturbance. These control systems are called negative feedback loops because the causality forms a loop (y.x.u.a.y) and the adjustment is typically the opposite sign, or the negative, of the disturbance. If d adds to y, the adjustment a subtracts from y; if d subtracts from y, a adds to it. There are many systems that operate through negative feedback loops. Sometimes rapid changes in a controlled parameter can outstrip the physiological mechanisms for reacting to these changes, resulting in potentially catastrophic changes in the internal environment. To avoid this, some physiological systems anticipate changes in controlled parameters and begin to do something about it even before the parameter changes. Most wide swings in controlled parameters have to do with behavior. Eating, for example, is followed by an influx of nutrients into the blood. The nervous system prepares the gastrointestinal tract for a meal by using sensory cues—the sight, aroma, and taste of food—to induce the secretion of gastrointestinal fluids even before food is swallowed. In another example, controlled parameters including blood pH, PCO2 (the partial pressure of CO2 in the blood—a measure of CO2 concentration) and blood PO2 help regulate the depth and frequency of breathing. Negative feedback mechanisms keep these controlled parameters within narrow ranges during normal activity. During strenuous activity, there appears to be little or no error in these controlled parameters, though the depth and frequency of breathing is markedly increased. This occurs through an anticipatory response of the central nervous system in which the depth and frequency of breathing is activated simultaneously with activity. DEVELOPMENTAL AND THRESHOLD CONTROL MECHANISMS REGULATE NONCYCLICAL AND CYCLICAL PHYSIOLOGICAL SYSTEMS Although negative feedback control is a major theme in physiology, it does not account for a variety of important physiological events. Developmental events include the onset of puberty and menopause. Pregnancy, parturition (birth), and cyclical events such as the menstrual cycle and the sleep/wake cycle are episodic events that do not obey negative feedback mechanisms and may involve positive feedback mechanisms. WE ARE NOT ALONE: THE MICROBIOTA In natural ecosystems, human beings are literally covered with tiny contaminating organisms. We typically have 10 times more bacteria than body cells, but each 11 12 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Small molecular weight, lipophilic signal Polypeptide Polypeptide hormone 4 hormone 3 Nuclear receptors Trimeric G-proteins Amplifying enzyme Catalytic receptors AC PLC 5 Ligand-gated channels P JAK 2 IP3 Ca2+ Ca2+ ATP cAMP ATP Ca2+ PKA Ca2+ 1 P Voltage-gated channels + mRNA Secretory protein Ribosome Proteins RNA polymerase II + - Bind to response element Nuclear receptors Nucleus FIGURE 1.1.7 Synopsis of signaling mechanisms on the cellular level. Cells receive electrical signals that can be converted into chemical signals through voltage-gated channels (1). The voltage-gated calcium channel is shown. Chemical signals released from nearby cells can also open ion channels, producing electrical signals in the cell (2). Cells receive chemical signals in the form of polypeptide hormones that cannot penetrate the cell membrane. These can affect the cell by coupling to heterotrimeric G-proteins (3) or to catalytic receptors on the surface of the cell (5). These are coupled to amplifying enzymes or to kinases that phosphorylate intracellular proteins. Small molecular weight, permeant chemical messengers (4) can enter the cell and bind to receptors in the nucleus, which then alter the kind or amount of specific proteins made by the cell. These signaling mechanisms are discussed in detail in Chapter 2.8. 4 5 Reference value or set point The actuator makes adjustment, a, based on input u from the comparator Disturbance r Comparator + – d u x Actuator a + FIGURE 1.1.8 Component parts of a negative feedback loop. A controlled parameter, y, is sensed by a sensor that releases a signal that related to the value of y [x 5 g(y)]. This signal is fed into a comparator, which compares the signal x to some reference or set-point value, causing the comparator to produce a signal u that is a function of the error of x from its reference [u 5 h(r 2 x), where u is the signal from the comparator, r 2 x is the error and h(r 2 x) gives the functional dependence of u on the error]. The signal u turns on an actuator that makes an adjustment, a, to the value of y. In negative feedback, the value of a reduces the error (r 2 x) so that the value of y returns towards its normal, set-point level. Disturbances, d, can alter the value of y and the negative feedback loop is engaged to minimize the departure of y from its set-point level. The reference value, r, is like a set point; its input into the comparator results in a signal u = h(r–x) 1 y is the controlled parameter y Controlled parameter Sensor 3 The comparator compares the sensor's signal, x, to a reference signal, r, to produce an output signal u, to the actuator 2 y is sensed by a sensor which produces an output signal, x, related to y (x = g(y)) The C ore P rinciples of Physiology of these is small, and so the total mass of these bacteria amounts to 0.51.5 kg. These organisms are present as surface contaminants, but they are the usual surface contaminants and so they can be considered to be part of us. They live on the surface of the body, and invasion within the body constitutes infection, and must be fought off by our immune systems. They live on the skin, in the oral cavity, in the airways, and in the gastrointestinal system and on the terminal parts of the reproductive system where it melds with the skin. Most of these are within the gastrointestinal system, and fully half of the feces is estimated to consist of bacteria. In addition to the bacteria, many natural ecosystems provide us with a load of other organisms: tapeworms, flukes, helminths (hookworms, pinworms, and roundworms), leeches, fleas, various fungal organisms, and a host of viruses. Although in natural ecosystems, infections with some of the multicellular parasites may be unusual, the load of bacterial contaminants is unavoidable. The aggregate of these hangers-on is called the microbiota. The microbiota engage signaling systems of the body and thereby alter our physiological states. This is true of the rhinoviruses that make us sneeze, thereby spreading them around to other individual hosts, and intestinal bugs that induce diarrhea, using the host signaling systems, to likewise spread them to other hosts. Evidence is accruing that even those “benevolent” strains that do not make us frankly sick still manipulate our physiology to their advantage. Thus the microbiota become part of our physiology. PHYSIOLOGY IS A QUANTITATIVE SCIENCE As described earlier, homeostasis refers to the maintenance of a constant internal environment, where the internal environment refers to the extracellular fluid that surrounds the cells. This internal environment is characterized by the concentration of a host of materials, and each of these concentrations has a unit and a numerical value. Many of these materials are metabolized by the tissues so that maintaining constant values requires matching supply to consumption. The rates of supply and consumption also have units and numerical values. As Figure 1.1.5 shows, the circulatory system unites all organs of the body by virtue of their perfusion with a common fluid, the blood. Maintenance of this flow requires pressure differences that also have units and magnitudes. Understanding the flows and forces that keep the blood moving and keep its composition relatively constant requires a quantitative approach. SUMMARY Physiology is the integrated study of the normal function of the human body. Like many complicated things, the body can be viewed as a set of subcomponents that interact by linking the output of one component to the input of another. These subcomponents are the organ systems. These include the cardiovascular system, the respiratory system, the renal system, the gastrointestinal system, the neuroendocrine system, the musculoskeletal system, the integument, and the reproductive system. Understanding how the body works as a whole requires us to make a model, either implicit or explicit, that explains the integration of the structures that make up organ systems, and the integration of the organ systems that produces the overall system behavior. Explanation requires that cause and effect in the model faithfully predicts cause and effect in the real system. Understanding can occur on different hierarchical levels of integration: the systems level, the organ level, the cell level, and the subcellular level. Each level seeks to explain behavior at that level on the basis of the components that make up that level. This is reductionism, the explanation of the behavior of a complicated object on the basis of its parts. We say that we understand something when we can explain function in terms of the parts one level below and we can show how behavior at that level contributes to behavior one level above. Aristotle identified four classes of causality: the material cause, the efficient cause, the formal cause, and the final cause. Explanation of something on the basis of the final cause is called teleology. Although science presumes that each component of living things obeys physical law alone, systems produce emergent properties that we seem to be unable to predict. These emergent properties belong to the system as a whole rather than to individual parts within it. Evolution is one cause of the human form and function, but it aids us only a little in understanding how physiological systems work. In the hierarchy of levels of organization, cells are the fundamental unit of life. The various cells of the body show a remarkable diversity of form and function, but they all carry the complete genome, with the exception of erythrocytes and reproductive cells. The diversity of form arises from the use of only parts of the genome for each type of cell. The overriding principle of human physiology is homeostasis, meaning the maintenance of a constant internal environment. Our internal environment is the extracellular fluid that bathes all cells in the body. A combination of internal control systems and external behavior maintains homeostasis. In the final analysis, our life depends on inputs from the environment and so our behavior (feeding, drinking, and temperature control) is crucial to our survival. Internal regulation of the internal environment relies primarily on negative feedback control, in which controlled parameters feed into a sensor that compares the value of the controlled parameter to the desired set point or reference levels. Variation from the set point results in the activation of effector mechanisms that increase or decrease the controlled parameter so that it more closely approximates the set-point level. Anticipatory control also contributes to homeostasis without allowing wide swings in the values of controlled parameters. Some important physiological events, such as puberty, menstruation, pregnancy, parturition, menopause, and the sleep/wake cycle, are not homeostatic. Instead, they incorporate switches between one physiological state and another. 13 14 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION These control systems contribute to the robust control of bodily functions that enable homeostasis under harsh environmental conditions. REVIEW QUESTIONS 1. What argument does holism make against reductionism? 2. Give some examples of emergent properties 3. What is a cell? Why is it considered to be a fundamental unit of organization of life? 4. Contrast genotype with phenotype 5. What is homeostasis? Why is it central to physiology? 6. What is the internal environment in large multicellular animals such as ourselves? 7. What would constitute proof for the theory of evolution? Do you think science has provided it? Why or why not? 8. From Einstein’s equation E 5 mc2, you have learned in physics that mass and energy are interconvertible. Why can we say that mass and energy are conserved in physiological systems? What does it mean to say that living things “transform” matter and energy? 9. Describe the hierarchical organization of the body. 10. Give examples of signaling at the organism level and cellular level. Physical Foundations of Physiology I: Pressure-Driven Flow Learning Objectives G G G G G G G G G G Define intensive and extensive variables Define flow and flux Describe the driving principle for heat flow, electrical current, diffusive flow, and volume flow Explain what is meant by fluxes moving downhill Write a continuity equation and describe its meaning Explain why steady-state flux requires a linear gradient of T, Ψ, C, or P List four capacitances commonly encountered in physiology Define pressure and be able to convert pressure between atm, mmHg, and Pa Write Poiseuille’s law, state its assumptions, and be able to calculate flow using it Write the Law of Laplace for cylindrical tubes and for spheres FORCES PRODUCE FLOWS EXTERNAL AND INTERNAL MOVEMENT IS A HALLMARK OF HUMAN LIFE For humans in their natural environment, movement is essential for survival. This movement refers to translation of the body from one location to another, and movement of the limbs relative to one another. In addition to this movement of body parts with respect to the external world, movement of materials within the body is also essential. Most important among these internal movements are the movement of the blood, movement of the air in and out of the lungs, movement of food and fecal material along the gastrointestinal tract, and movement of the urine from its formation to elimination. In addition, the body transports materials across barriers such as the gastrointestinal tract lining, lungs, and kidney tubule. Transport also occurs within cells. All of these movements require the continued application of force to overcome inertia and friction. 1.2 TRANSPORT OF MATERIAL IS DESCRIBED AS A FLOW OR A FLUX The transport of material is quantitatively expressed as a flow, which we will symbolize by the variable Q. The flow can be expressed as G G G G volume of material or fluid transported per unit time; mass of material transported per unit time; number of particles or moles transported per unit time; number of ions or unit charges transported per unit time (electrical current). FLOW DEPENDS ON THE AREA; FLUX IS FLOW PER UNIT AREA The total flow of volume or solute is an extensive variable: the flow depends on the extent or the amount of the system that gives rise to the flow. In the case of two compartments separated by a membrane, doubling the area or extent of the membrane would produce twice as much flow between the two compartments. Dividing the flows by the area normalizes the flows. The normalized flow is the flux, and the flux is an intensive variable whose value is independent of the extent of the system. Flux is defined as QV A QS JS 5 A JV 5 ½1:2:1 where QV is the volume flow, JV is the volume flux, A is the cross-sectional area through which flow occurs, oriented at right angles to the direction of flow, QS is the amount of material (solute) transported per unit time, JS is the solute flux, and A is the area. The units of flux are amount or volume or mass or charge per unit time per unit area. Strictly speaking, fluxes and flows are vectors, consisting of the magnitude of the flux or flow and its direction. Unless otherwise noted, we will consider flux or flow only in one direction and therefore we will suppress the vector nature of flux and flow. 15 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00002-1 16 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION FLUX DEPENDS LINEARLY ON ITS CONJUGATE FORCE For a variety of forces and fluxes, the flux that results from a net force varies linearly with the force: ½1:2:2 Jx 5 LFx ½1:2:4 where Jx is the flux of something, L is a phenomenological coefficient, and Fx is the net force that drives the flux. This generic equation holds for a variety of kinds of fluxes. The flux of heat energy, electrical flux (the current density), diffusion of solute and pressure-driven flow all obey this general phenomenological law. In each of these cases, the net force is proportional to the gradient of an intensive variable. Strictly speaking, the gradient is a vector quantity, but we use it here to denote the slope of these intensive variables along one dimension: ½1:2:3 diffusion gradient will also modify the flux of solute. In general, flows produced by multiple flow processes are not independent. If there are two forces driving flows, we write Fourier0 s law of heat conduction: JH 5 2λ dT dx Ohm0 s current law: Je 5 2σ dψ dx Fick0 s first law of diffusion: JS 5 2D Pressure-driven flow: JV 5 2LP dC dx dP dx where JH is the flux of heat energy, dT/dx is the temperature gradient, λ is the coefficient of thermal conductivity, Je is the electrical current flux, dψ/dx is the voltage gradient, σ is the electrical conductivity, JS is the solute flux, dC/dx is the concentration gradient, D is the diffusion coefficient, JV is the volume flux, dP/dx is the pressure gradient, and LP is the hydraulic conductivity. All of these phenomenological equations find application in physiological systems. They are all analogues of Ohm’s law. These equations are true only if the only driving force is the one specified. For example, diffusion of electrolytes, charged solutes, is influenced by electric fields. If a voltage gradient is also present along with a concentration gradient, Fick’s first law of diffusion would need to be modified to reflect that influence. A pressure difference that produces a volume flow in the presence of a J1 5 L11 F1 1 L12 F2 J2 5 L21 F1 1 L22 F2 where L11 is the coefficient relating flux 1 to its primary driving force 1 and L22 relates flux 2 to its primary driving force 2, and L12 and L21 are the coupling coefficients that describe how secondary forces affect the flows. An example of this is a bimetallic junction. When two unlike metals are joined together, passing a current through the junction causes it to either heat up or cool, and this is called the Peltier effect. The coupling coefficient implies that if you heat up or cool the junction, a current will flow. This is the basis of the thermocouple. In this case, the two fluxes are heat and current and the two forces are temperature gradient and voltage gradient. Because of a principle called microscopic reversibility, it turns out that the cross-coupling coefficients are equal: L12 5 L21. This is called Onsager reciprocity, in honor of Lars Onsager (19031976) who earned the Nobel Prize in Chemistry in 1968 for this discovery. FLUX MOVES DOWNHILL The relations in Eqn [1.2.3] describe fluxes in one dimension. Both fluxes and gradients are actually vectors, but we consider a single direction here for simplicity. Consider Fick’s law of diffusion for solutes. If the gradient of concentration is constant, we may write ½1:2:5 JS 5 2D ðC1 2 C2 Þ ðx1 2 x2 Þ for two points (C1, x1) and (C2, x2). If C1 . C2 and x1 , x2, the slope is negative and the flux is positive. If C1 , C2 and x1 , x2, then the slope is positive and the flux is negative. Thus, the flux always goes from regions of high concentration to regions of low concentration (see Figure 1.2.1). This is true for all the intensive variables for the fluxes in Eqn [1.2.3]. These fluxes always move downhill, unless acted upon by additional forces. x-axis dC < 0 dx C J>0 dC > 0 dx J<0 C X (distance, m) X (distance, m) FIGURE 1.2.1 Flux moves downhill. Consider one-dimensional flux, with positive flux defined as in the direction of the x-axis. In this case, we consider diffusion that is driven only by a concentration gradient. If the gradient is negative (higher concentrations at lower values of x), then by Eqn [1.2.3], the flux is positive and directed to the right (middle panel). If the gradient is positive, then the flux is negative and directed to the left. In each case, the flux of solute is from the region of high concentration toward the region of lower concentration. Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow CONSERVATION OF MATTER OR ENERGY LEADS TO THE CONTINUITY EQUATION Here we consider that a concentration gradient exists and produces a solute flux as a consequence. We consider a cylindrical tube as shown in Figure 1.2.2, having a cross-sectional area A, which is intersected at right angles by planes at x 5 x and x 5 x 1 Δx. so that the tube is cut into three compartments, the left, middle, and right compartments. The concentration may vary with time and distance. We define the concentration in any volume element as Nðx; tÞ ½1:2:6 Cðx; tÞ 5 V where N(x, t) is the number of solute particles in the volume element and V is the volume element. We define J(x) as the net number of solute particles crossing the plane at x 5 x per unit time per unit area, with positive being directed along the x-axis, to the right. The number of particles entering the middle compartment from the left in time Δt is AJ(x)Δt and the number leaving the middle compartment by crossing the plane at x 5 x 1 Δx is AJ(x 1 Δx)Δt. Here the parenthesis means “function of” and not multiplication. If there is no chemical transformation of the solute particles, their number is conserved and we can write ΔN 5 AJðxÞΔt 2 AJðx 1 ΔxÞΔt ½1:2:7 ΔN 5 AJðxÞt 2 AJðx 1 ΔxÞ Δt Dividing by the volume element V 5 AΔx and rearranging, we have 1 ΔN AJðxÞ 2 AJðx 1 ΔxÞ 5 AΔx Δt AΔx ½1:2:8 N V Δ Δt 52 Similar continuity equations can be written for the flux of heat energy, charge, and volume. Their form is given in Eqn [1.2.10]: Heat: ρCP C @ψ @Je 52 V @t @x Charge: ½1:2:10 Solute: @T @JH 52 @t @x @CS @JS 52 @t @x Volume: C @P @JV 52 V @t @x where ρ is the density of matter through which heat flows and Cp is its specific heat capacity. In the next equation, C is the electrical capacitance, V is the volume. Next, Cs is the concentration. In the last line C is the compliance, and V is the volume. Here we have the unfortunate situation in which single variables denote different quantities: C stands for electrical capacitance (in farads), concentration (usually in molar) and compliance (5ΔV/ΔP). Generally the meaning of the variable is clear from its context. STEADY-STATE FLOWS REQUIRE LINEAR GRADIENTS AJðx 1 ΔxÞ 2 AJðxÞ AΔx In the limit of Δt-0 and Δx-0, this becomes ½1:2:9 This equation is called the continuity equation. What this equation says is that if the flux of solute is not the same everywhere, then the amount of solute must be building up or becoming depleted somewhere, and this buildup or depletion changes the concentration of solute. It is a straightforward consequence of the conservation of material. This equation is not true if the diffusing chemical undergoes chemical transformation. In this case, it is not conserved. @Cðx; tÞ @JðxÞ 52 @t @x Cross-sectional area, A In homeostasis, there is a steady supply of nutrients and removal of wastes and a steady withdrawal of nutrients and a steady production of wastes by the tissues. This steady state in which all flows are constant is more easily amenable to mathematical analysis. What “steady state” means is that each of the variables on the lefthand side of Eqn [1.2.10] is zero because at the steady state there are no changes in temperature, charge, concentration, or pressure with time: Heat: ρCP J(x) J(x +Δx) C @ψ @Je 52 50 V @t @x Charge: ½1:2:11 x x +Δx x axis FIGURE 1.2.2 Fluxes as a function of distance in the presence of a concentration gradient. Solute: @T @JH 52 50 @t @x @CS @JS 52 50 @t @x Volume: C @P @JV 52 50 V @t @x 17 18 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Substituting in from Eqn [1.2.3] for JH, Je, JS, and JV, we have @2 T 50 @x2 ½1:2:12 @2 ψ 50 @x2 @2 C 50 @x2 capacitance is an important concept for physiologists as well, because membrane potential derives from a separation of electrical charges across the membrane, and the membrane itself acts like a tiny capacitor with two conducting plates, separated by a dielectric. We will discuss this further in the sections on membrane potential, action potential, and the cable properties of nerves (Chapters 3.13.3). The other relationships completely analogous to the relation between charge, capacitance, and voltage, are H 5 CP MT: Heat energy 5 heat capacity 3 mass 3 temperature @ P 50 @x2 2 This condition is met only if the gradient of T, ψ, C, or P is constant; thus, the slope of T, ψ, C, and P at steady state is constant, and each of these intensive variables varies linearly with distance. HEAT, CHARGE, SOLUTE, AND VOLUME CAN BE STORED: ANALOGUES OF CAPACITANCE The steady state is often approximated in the body but rarely achieved. At rest heat production balances heat exhausted to the environment. When we begin exercising, heat production rises rapidly and the temperature of the body rises accordingly until, once again, heat production matches heat transfer to the environment, achieved by using other forces besides the conduction described in Fourier’s law. This new steady state of temperature during exercise is achieved at different operating conditions than at rest. In another example, transport of blood through the cardiovascular system is pulsatile, because the pressure that drives transport comes from the heart, and the heart produces force rhythmically. Each of the main four variables we have been discussing, heat, charge, amount of chemicals, and volume, can be temporarily stored or depleted. Electrical charge can be stored in capacitors. The constitutive relation between charge, voltage, and capacitance is given as ½1:2:13 Q 5 CV where Q here stands for charge, C is the capacitance, and V is the voltage. Here we are victims of the use of the same variables to denote entirely different quantities. We will use Q most often to signify a flow, but here it signifies charge, in coulombs. In physiology, we often use C to denote concentration, but here it means capacitance, in farads (5CV21); in physiology, V usually signifies volume, but here it means electrical potential, in volts. Electrical M 5 VC: Amount 5 volume 3 concentration V 5 CP: Volume 5 compliance 3 pressure ½1:2:14 where the capacitance-like elements include electrical capacitance (C in Eqn [1.2.13]), thermal mass (CpM, the specific heat capacity times the mass), volume (V), and compliance (C). Note again the multiplicity of uses of a single notation. C variously stands for capacitance, heat capacity, concentration, or compliance. The capacitances are all expressed as the ratio of an extensive variable and an intensive variable and are all themselves extensive variables. Table 1.2.1 summarizes the four kinds of capacitances. PRESSURE DRIVES FLUID FLOW In the case of fluid or air flow, pressure differences drive the flow. The SI unit of pressure is the pascal, Pa, equal to 1 N m22. However, physiologists still use other units, notably the atmosphere and mmHg. The atmospheric pressure is the weight of a column of air equal to the height of the atmosphere in the earth’s gravitational field per unit area of the earth’s surface. The actual pressure in the atmosphere decreases as you ascend, but the unit of 1 atm is defined for a standard condition of the air and standard altitude at sea level. The conversion between atmospheres and mmHg is an observed phenomenon. Atmospheric pressure can be measured in units of mmHg as described in Figure 1.2.3. Figure 1.2.3 illustrates that atmospheric pressure supports a column of 760 mmHg high. The pressure of this column of Hg is equal to atmospheric pressure, and the pressure of the column is simply its weight divided by its area. The weight is the force of gravity acting on the column, and is given as ½1:2:15 W 5 F 5 mg 5 ρVg 5 ρAhg TABLE 1.2.1 Four Kinds of Capacitances Capacitance Expression Units Application Electrical capacitance, C Q/V Nerve conduction, membrane potential Farads 5 coulombs/volt 21 Thermal capacitance, CpM H/T JK Chemical capacitance, V, volume M/C Volume 5 moles/molarity, L Mechanical capacitance, C, compliance V/P 5 joules/temperature Heat production/loss temperature regulation Metabolism, filtration Compliance 5 volume/pressure, m3 Pa21 Blood pressure, breathing Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow High vacuum High vacuum pump h = 760 mm Atmospheric pressure supports a column of Hg Weight of column of Hg = mg Pressure force up = PA Hg At equilibrium, the weight of the column of Hg equals the force of atmospheric pressure: mg = P A FIGURE 1.2.3 Measurement of atmospheric pressure. A closed vertical tube is connected to a high vacuum pump and evacuated air. When inverted into a dish of mercury, the atmospheric pressure forces mercury up the tube until mechanical equilibrium is achieved when the weight of the column of mercury exerts a pressure equal to the atmospheric pressure. At sea level in dry air, 1 atm will support a column 760 mmHg high. TABLE 1.2.2 Conversion Between Pressure Units atm mmHg Pa 1 760 1.013 3 105 1 133.29 0.00750 1 0.00132 26 9.87 3 10 The pressure is just the force per unit area. Dividing Eqn [1.2.15] by the area, we get F 5 ρgh A Thus, the height of the column of mercury in equilibrium with the atmospheric pressure is independent of its area. We need to specify only the height of the column of mercury. Thus, at sea level, the atmospheric pressure supports a column of 760 mmHg high and we say that 760 mmHg 5 1 atm. ½1:2:16 P5 The value of atmospheric pressure in pascals 5 N m22 can be calculated from 760 mmHg by using the density of Hg (13.59 g cm23) and the acceleration due to gravity (9.81 m s2). Inserting these values into Eqn [1.2.16], we get P 5 13:59 g cm23 3 9:81 m s22 3 0:76 m 5 13:59 g cm23 3 ð100 cm m21 Þ3 3 1023 kg g21 3 9:81 m s22 3 0:76 m 5 13:59 3 103 kg m23 3 9:81 m s22 3 0:76 m 1 atm 5 101:3 3 103 kg m s22 m22 5 1:013 3 105 N m22 5 1:013 3 105 Pa We can therefore complete a conversion table for pressure units (Table 1.2.2). POISEUILLE’S LAW GOVERNS STEADY-STATE LAMINAR FLOW IN NARROW TUBES In 1835, Jean Leonard Marie Poiseuille experimentally established the relationship between flow through narrow pipes and the pressure that drives the flow. The relationship is ½1:2:17 πa4 ΔP Qv 5 8η Δx where QV is the flow, in units of volume per unit time, π is the geometric ratio, a is the radius of the pipe, η is the viscosity, ΔP is pressure difference between the beginning and end of the pipe, and Δx is the length of the pipe. This equation describes the relationship between flow and pressure difference only for laminar flow. Laminar flow is steady, streamlined flow, and it is distinguished from turbulent or chaotic flow. This equation is often applied to problems in physiology even though the conditions for its valid application are missing. Its application requires us to understand viscosity. Consider two parallel plates separated by a fluid, as shown in Figure 1.2.4. The top plate can be moved at a constant velocity relative to the stationary bottom plate only if the plate is subjected to a force that continuously overcomes the frictional resistance on the plate caused by its contact with the adjacent fluid. The viscosity is the resistance of a fluid to shear forces. It is defined by the equation ½1:2:18 F dv 5η A dy 19 20 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Moving plate Force applied to the moving plate at constant v is opposed by an equal drag force of the fluid on the plate F FD Thin fluid layer The layer of fluid adjacent to the plate has the same velocity as the plate The layer of fluid adjacent to the stationary plate has zero velocity Stationary plate FIGURE 1.2.4 Definition of viscosity. Two plates are separated by a fluid. The top plate moves with constant velocity, v, with respect to the stationary bottom plate. The fluid adheres to the plates and a thin layer of fluid immediately adjacent to the plates has the same velocity as the plates. This results in a velocity profile in the fluid. The steepness of this velocity profile, dv/dy, is the velocity gradient. where F is the shear force, A is the area, v is the velocity, and y is the dimension perpendicular to the plate. The ratio F/A is called the shear stress and the quantity dv/dy is called the velocity gradient. F/A in this equation has the units of pressure, Pa 5 Nm22 and dv/dy has the units of m s21 m21 5 s21, so the units of η in SI are Pa s. Thus, the viscosity is the ratio of the shear stress, F/A, dvided by the velocity gradient, dv/dy. In older texts viscosity is sometimes given in units of poise 5 1 dyne cm22 s. These can be converted to Pa s by using the definition of Pa 5 1 N m22, 1 N 5 1 kg m s22 and 1 dyne 5 1 g cm s22: Thus, 1 N 5 105 dyne. 1 Pas 5 1 N m22 s 5 1 kg m s22 m22 s 5 1 kg 3 103 g kg21 m 3 100 cm m21 3 s22 3 m22 3 ð0:01 cm m21 Þ2 s 5 103 g 3 100 cm 3 s22 3 1024 cm22 s 5 10 g cm s22 2 cm22 s 5 10 dyne cm22 s 5 10 poise So 1 Pa s 5 10 poise EXAMPLE 1.2.1 Ultrafiltration in the Kidney The kidney has a structure called the glomerulus that consists of combined layers of cells and extracellular matrix—bundles of fibers in the extracellular space—that together form an ultrafilter (see Chapter 6.2 for further description). It is called an ultrafilter because the combined layers retain proteins while letting most small solutes pass into the ultrafiltrate. We model the membrane here as a flat membrane that is pierced by many identical right cylindrical holes, or pores. Assume that the radius of the pores is 3.5 nm and the pore length is 50 nm. The viscosity of the fluid is taken to be the same as plasma, 0.02 poise. The aggregate area of the pores makes up 5% of the total surface area of the membrane. The total pressure on the input side, the side of the blood, averages 60 mmHg and on the ultrafiltrate side the total pressure averages 45 mmHg (see Chapter 6.2 for a discussion of the origin of these pressures). The total available area of the membrane is 1.5 m2. What is the filtration rate in cm3 min21? The situation is depicted schematically in Figure 1.2.5. We use Poiseuille’s equation here. The total flow, QV, is the sum of the flow through all of the pores: Qv 5 Nqv 5 Nπa4 =8η ðΔP=ΔxÞ Pore Pultrafiltrate Pblood qV n = N pores /unit area Δx FIGURE 1.2.5 Model of the kidney ultrafilter. where N is the number of pores and qV denotes the flow through a single pore. Here a is the radius, a 5 3.5 3 1029 m, η is the viscosity (η 5 0.02 poise 3 1 Pa s/10 poise 5 0.002 Pa s), ΔP is the pressure difference 5 15 mmHg 3 133.3 Pa mmHg21 5 1999.5 Pa, Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow and Δx 5 60 3 1029 m. Now that all the units are compatible, we plug them into the equation and get: 29 qV 5 π3ð3:53 10 236 5 471:44310 23 mÞ =83 23 10 4 23 m =16310 4 29 Pa s3 ð1999:5 Pa=60310 mÞ 21 Pa s 333:33310 Pa3m 9 5 982:13 10224 m3 3s21 3ð100 cm3m21 Þ3 3 60 s3 min21 N 3 πa2 5 0:05 3 1:5 m2 Knowing that a 5 3.5 3 1029 m, we solve for N: N 5 0:075 m2 =π 3 ð3:5 3 1029 mÞ2 5 1:95 3 1015 which is a lot of pores! Multiplying N 3 qV, we get Qv 5 N 3 qv 5 1:9 3 1015 pores 3 5:89 3 10214 cm3 min21 pore21 5 5:893 10214 cm3 3 min21 5 111:9 cm3 min21 This is the flow through a single pore. We need to know how many of them are there. If their aggregate area is 5% of the total, then the number of pores can be calculated from THE LAW OF LAPLACE RELATES PRESSURE TO TENSION IN HOLLOW ORGANS Rearrangement of this equation gives the Law of Laplace for a cylinder: ½1:2:20 The blood vessels maintain a pressure difference across their walls. These vessels approximate hollow cylinders. The gallbladder, urinary bladder, heart, and lung alveoli also maintain a pressure difference, and approximate hollow spheres. The Law of Laplace can be derived for these ideal geometries by considering mechanical equilibrium. Consider first a circular cylinder of radius r subjected to an external pressure Po and internal pressure Pi, as shown in Figure 1.2.6. In Figure 1.2.6, the internal pressure acting on the upper half of the cylinder by the lower half cylinder produces a force on the upper half that is given as Pi 3 2r 3 L. The external pressure acting over the surface of the upper half of the cylinder produces a force that is equal to Po 3 2r 3 L, as can be deduced from the situation where Po 5 Pi in a mechanically stable fluid with no walls. The balance of the internal and external pressures produces a net force of ΔP 3 2r 3 L where Δ P 5 Pi 2 Po. This is balanced by the force of the walls of the lower half acting on the upper half. These forces are the wall tension, in units of N 3 m21, acting along the length of the wall. These forces are directed downward, as shown. For mechanical equilibrium to occur ½1:2:19 Fnet 5 0 5 T2L 5 ΔP2rL r T T T L Pi This is a reasonable approximation to the filtration rate in an adult human. ΔP Po FIGURE 1.2.6 Right circular cylinder of radius r and length L is subjected to a transmural pressure difference. For mechanical equilibrium, the net force on any part of the cylinder must be zero. We split the cylinder in half down its axis. The sum of forces on the upper half must sum to zero. ΔP 5 T r An exactly analogous argument can be made for a sphere. In this case, the pressure is distributed over the crosssectional area of the hemisphere, and the wall tension is distributed around the circumference. The result gives ½1:2:21 Fnet 5 0 5 T2πr 5 ΔPπr 2 which is easily rearranged to ½1:2:22 ΔP 5 2T r SUMMARY Flow consists of the movement of something from one place to another in the body. We identify four classes of flow of major importance in physiology: flow of heat energy, electric current, solute, and volume (gas or body fluids). Flow is an extensive variable, depending on the extent of the system. Flux is the flow divided by the area through which the flow occurs. Fluxes are driven by forces. Heat flux is driven by a temperature gradient, electrical current by a potential gradient, solute flux by a concentration gradient, and volume flux by a pressure gradient. Gradients are vector quantities. Here we consider one-dimensional flow and flux and suppress the vector nature of these gradients. In each of these cases, flux occurs “downhill,” meaning from regions of high temperature, potential, concentration or pressure, to regions of low temperature, potential, concentration, or pressure. Each flux has the form Jx 5 2K dψx dx The continuity equation for each of these fluxes states that a gradient of flux produces a time dependence of the driving force. If the flux is not the same everywhere, then temperature, voltage, concentration, or pressure varies with time. The continuity equation is a consequence of conservation of energy, electric charge, chemical species, or volume. 21 22 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Steady-state flows require linear gradients of temperature, potential, concentration, or pressure. Heat energy, electric charge, concentration, and volume can all be stored in the body. The ability to store these things is quantified by capacitances. These include thermal capacitance, electrical capacitance, chemical capacitance, and mechanical capacitance. The thermal capacitance is MCp, where Cp is the specific heat capacity; the units of MCp are in J K21; electrical capacitance is Q/V, in F; chemical capacitance is V, in L; mechanical capacitance is compliance, in L Pa21. Pressure drives flow. Pressure is measured in pascals 5 1 N m22. This unit is equal to 9.87 3 1026 atm or 0.0075 mmHg. Steady-state pressure-driven flow through narrow pipes is described by Poiseuille’s law: πa4 ΔP QV 5 8η Δx where a is the radius of the pipe, η is the viscosity, in Pa s, π is the geometric ratio of circumference to diameter of a circle, ΔP is the pressure difference in the pipe, and Δx is the length of the pipe. Viscosity is the resistance of a fluid to shear forces and is defined mathematically as the ratio of the shear stress to the velocity gradient. A transmural pressure within hollow organs or tubes requires tension in the walls of these organs. So far as these organs approximate thin-walled cylinders or spheres, the wall tension obeys the Law of Laplace: T r 2T ΔP 5 r Cylinder: ΔP 5 Sphere: where ΔP is the transmural pressure difference, T is the wall tension, in N m21, and r is the radius of the cylinder or sphere. REVIEW QUESTIONS 1. Is density an intensive or extensive variable? Is temperature intensive or extensive? 2. What drives current flow? What drives solute flow? If solute is charged, would its movement make a current? 3. What does it mean to say that fluxes “move downhill”? 4. What does capacitance mean for charge? Solute? Heat? Volume? 5. How is Ohm’s law like Fick’s law of diffusion? 6. What are the units of viscosity? 7. What are the assumptions in the derivation of Poiseuille’s law? Are these reasonable? 8. What is the relationship between velocity and flux in fluid flow? 9. What is the relationship between capacitance, charge, and voltage? 10. What is the relationship between volume, amount, and concentration? 11. What is the relationship between compliance, volume, and pressure? 12. What is the relationship between heat capacity, heat energy, and temperature? 13. What is the relationship between wall tension and pressure in a sphere? In a cylinder? APPENDIX 1.2.A1 DERIVATION OF POISEUILLE’S LAW POISEUILLE’S LAW DESCRIBES PRESSUREDRIVEN FLOW THROUGH A CYLINDRICAL PIPE Consider two fluid compartments which are joined by a right cylindrical pipe, as shown below, of area A. Since pressure is defined as a force per unit area, the total force acting on the fluid in the pipe on the left-hand surface is just PLA, and the total force acting on the fluid in the pipe on the right-hand side is PRA. If PL and PR are not equal, then the fluid within the pipe will be subject to a net force and therefore this volume of fluid will be accelerated. The result will be a movement of fluid in the direction of the net force. What we wish to establish is the quantitative relationship between the resulting flow and the pressure difference which drives this flow. SHEAR STRESS IS THE VISCOSITY TIMES THE VELOCITY GRADIENT The movement of fluid through the pipe shown in Figure 1.2.A1.1 encounters resistance along the cylindrical surfaces of the pipe. This resistance is key to deriving an expression for the flow as a function of pressure. It is due to the viscosity of the fluid, as shown in Figure 1.2.4. We consider that the flow of fluid occurs in layers, or laminae, and that each layer exerts a viscous drag on the layer immediately above and below. The fluid remains in contact with the walls on either side of the tube, and so has a velocity v 5 0 at the walls. In order to achieve a constant velocity, we need to apply a continual force, F, in order to overcome the viscous drag of the fluid. The shear stress is the force exerted by one lamina on an adjacent one, and is given by ½1:2:A1:1 F dv 5η A dy PL Area PR FIGURE 1.2.A1.1 Flow is driven by pressure differences. Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow where F is the force, A is the area, η is the coefficient of viscosity of the fluid, and dv/dy is the gradient of velocity. This equation makes good common sense: the force necessary to keep the lamina moving depends on how much area is exposed to the resistance, it depends on how sticky the fluid is, and it depends on how fast it is going. Insertion of this result into Eqn [1.10.A1.2] gives ½1:2:A1:4 0 5 ðPðxÞ 2 Pðx 1 ΔxÞÞ2πr dr dv 1 2πrΔxη dr r 1 2πðr 1 drÞΔxη FLOW THROUGH A PIPE We consider here the flow through a tube or pipe of radius a and length Δx. The pressure at the left end of the pipe is P(x) and the pressure at the right end is P(x 1 Δx). We consider here a constant flow which therefore has a constant velocity. This means that the fluid is not subjected to any net force. Consider the forces experienced by a hollow shell of inner radius r and outer radius r 1 dr, as shown in Figure 1.2.A1.2. The fluid in contact with the walls of the pipe does not move, so that v 5 0 at r 5 a. The velocity of the fluid increases as one approaches the center of the tube. The actual velocity profile will be solved on the way to deriving an expression for the total flow through the pipe. Because the fluid flows through the pipe at a constant velocity along the pipe (but which depends on the distance from the walls of the pipe), the sum of the forces on the hollow shell shown in Figure 1.2.A1.2 must be zero. The forces to the right include the force of the pressure on the left and the drag force of the inner layer of fluid, and the forces to the left include the pressure on the right and the drag force of the outer fluid. Thus, we have ½1:2:A1:2 F1 P(x) which can be re-written as 0 5 2ðΔPÞ2πr dr dv 1 2πrΔxη dr r dv dv 2 2πrΔxη 2 2π drΔxη dr r1dr dr r1dr ½1:2:A1:5 where ΔP 5 P(x 1 Δx) 2 P(x). We next approximate the gradient of velocity at (r 1 dr) by the first two terms of a Taylor’s series: ½1:2:A1:6 r F2 P(x + dx) r + dr x x + dx FIGURE 1.2.A1.2 Balance of forces on a hollow cylinder of inner radius r and outer radius r 1 dr. F1 is the drag force of the lamina with inner radius r 1 dr on the hollow cylinder; F2 is the drag force on the hollow cylinder by the lamina immediately inside, with outer radius r. 2 dv dv d v C 1 dr dr r1dr dr r dr 2 r Substitution of Eqn [1.2.A1.6] into Eqn [1.2.A1.5] gives ΔP 2πr drΔx 052 Δx dv 1 2πrΔxη dr r 0 5 PðxÞA 1 F2 2 Pðx 1 ΔxÞA 2 F1 The area of the hollow sphere on which the pressures act is 2πrdr. The drag forces F1 and F2 are given by Eqn [1.2.A1.1] as dv F1 5 2πðr 1 drÞΔxη dr r1dr ½1:2:A1:3 dv F2 5 2πrΔxη dr r dv dr r1dr dv dv 2 2πrΔxη 2 2π dr Δxη dr r dr r 2 2πr drΔxη 2 2 d v d v 2 2 2πr dr Δxη dr 2 r dr 2 r ½1:2:A1:7 Canceling out the like terms, and factoring out the term 2πrdr Δx, we obtain 2 ΔP η dv d v η d2 v 2η 2 dr 052 2 Δx r dr r dr 2 r r dr 2 r ½1:2:A1:8 In the limit as dr-0, the last term in Eqn [1.2.A1.8] vanishes. Multiplying through by 21, we are left with the differential equation 2 ΔP η dv d v 2 05 1η ½1:2:A1:9 Δx r dr dr 2 This second-order differential equation can be solved by converting it to a first order differential equation with the substitution y 5 dv/dr, and then multiplying through 23 24 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION by an integrating factor, ρ. Re-arranging Eqn [1.2.A1.9], we have dy ρ ρ ΔP ½1:2:A1:10 1 y1 50 ρ dr r η Δx We choose ρ so that the first two terms are an exact differential. This is true if ρ 5 r. Thus we have dy dðryÞ r ΔP 52 r ½1:2:A1:11 1y5 dr dr η Δx Integrating Eqn [1.2.A1.11], we obtain ð ð r ΔP dðryÞ 5 2 dr η Δx ½1:2:A1:12 1 ΔP 2 ry 5 2 r 2η Δx Canceling the r factor on both sides of Eqn [1.2.A1.12] and recalling that y 5 dv/dr, we integrate Eqn [1.2.A1.12] again: dv 1 ΔP 52 r dr 2η Δx ð ð 1 ΔP dv 5 2 r dr ½1:2:A1:13 2η Δx 1 ΔP 2 v52 r 1C 4η Δx where C is a constant of integration which can be evaluated from the boundary conditions. The boundary conditions are that v 5 0 when r 5 a. That is, the velocity of the fluid immediately adjacent to the walls of the pipe is zero. Insertion of v 5 0 and r 5 a into Eqn [1.2.A1.13] gives a2 ΔP C5 4η Δx ½1:2:A1:14 1 ΔP v5 ða2 2 r 2 Þ 4η Δx This equation gives the velocity of fluid flow as a function of the radial distance from the center (r 5 0) to the edge (r 5 a) of the pipe. This equation says that the velocity profile is parabolic. What we wanted to do at the outset was to calculate the total flow through the pipe, QV. This is the volume of fluid which crosses the total cross-section of the pipe per unit time. The volume flux, JV, is the volume moving through a small increment of unit area per unit time. In fact JV is the velocity of fluid movement. To see this, consider a block of fluid moving at constant velocity, v. The block has a cross-sectional area, A. In time t the block moves a distance vΔt. The volume of fluid moving in this time is AvΔt. The volume flux, JV, is the volume of fluid moved per unit area per unit time. This is ½1:2:A1:15 JV 5 AvΔt 5v AΔt Thus, the volume flux is the velocity of fluid movement. To find the total fluid flow, we integrate the flow as a function of distance from the center of the pipe. Thus ða ða ða QV 5 dQV 5 JV dA 5 JV 2πr dr 0 0 ð0a 5 v2πr dr 0 ða ½1:2:A1:16 1 ΔP 5 ða2 2 r 2 Þ2πr dr 4η Δx 0 ða 2π ΔP 5 ða2 2 r 2 Þr dr 4η Δx 0 The last integral can be evaluated by making the substitution u 5 (a2 2 r2), so du 5 22rdr, to obtain ð r5a ð r5a du 2 2 ða 2 r Þr dr 5 u 2 2 r50 r50 ð u50 1 52 u du ½1:2:A1:17 2 u5a2 5 a4 4 Inserting this result of the integration into Eqn [1.2.A1.16], we obtain πa4 ΔP QV 5 ½1:2:A1:18 8η Δx This last equation gives the total flow through the pipe, QV. This equation is called Poiseuille’s Law, in honor of the French physician Jean Leonard Marie Poiseuille, who experimentally established the law in 1835. The relation shows that the total flow is linearly dependent on the driving force for the flow, the pressure difference between the left and right ends of the pipe. Further, it is inversely related to the length of the pipe, Δx, and to the viscosity, η. Most importantly, the flow is proportional to the fourth power of the radius of the pipe. One of the points of the derivation given above was to show the mathematical origin of this very steep dependence on the size of the pipe. APPENDIX 1.2.A2 INTRODUCTORY STATISTICS AND LINEAR REGRESSSION INTRODUCTION Statistics have two functions: (1) to describe the variation in some data and (2) perform tests on sets of data to determine the cause of the variation in that data. As an example, we know that people of the same age are not the same height or weight. The set of heights for a population can be described by a statistic that describes the center (mean, median or mode) and by a statistic that describes the variation around that center. Typically the mean is used to describe the center and the standard deviation is used to describe the variation, or spread around that center. These are examples of descriptive statistics. Suppose we wanted to know the causes of Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow variation in human height. Possible causes might include diet, genetics, sleep patterns. It is extremely difficult to determine the cause of variation in humans because these possible causes are not controlled. However, there are populations that differ in a variety of ways. We may ask the question, do two different populations have different average heights? These kinds of questions involve statistical tests. n P ½1:2:A2:5 s2 5 i51 Xi2 2 2X Suppose we have a set of variables {X1, X2, . . ., Xi, . . ., Xn}. The mean is defined as n P n X ½1:2:A2:6 X 5 μX 5 n P i51 Xi 5 n X n P i51 2 σ 5 2 i51 n ðXi 2XÞ2 i51 The sample variance is a measure of the spread of a population when some sub-set of the population is used so that the mean is estimated rather than known. Its formula is similar to that of the population variance, except that one “degree of freedom” is used for the estimation of the mean. The sample variance is calculated as s2 5 i51 n21 2 here s is used as a symbol for the sample variance, whereas σ2 is used as a symbol for the population variance. Note that the formula for s2 is very similar to the formula for the variance except that n 2 1 is used as the divisor instead of n, the number of measurements used to estimate the mean and variance of the sample. The term (n 2 1) is the degrees of freedom that refers to the number of independent pieces of information that goes into the estimation of the statistic. Because the mean is used in the calculations, each measurement contributes to the mean and, if you know the mean, the last value can be calculated from all of the other values. Thus the degrees of freedom when the data is used to estimate the mean is n 2 1. Expanding the square term in Eqn [1.2.A2.3], we have s2 5 i51 i51 Xi2 2 n X 2 n ½1:2:A2:8 s2 5 n P i51 Xi2 2 n2 X 2 n ðn 2 1Þ Again making use of Eqn [1.2.A2.6] we arrive at the computational formula for the sample variance: n 2 n P P 2 n Xi 2 Xi i51 i51 ½1:2:A2:9 s2 5 n ðn 2 1Þ 2 ðXi2 2 2X Xi 1 X Þ n21 The sample standard deviation, denoted by the statistic s, is defined to be the positive square root of the sample variance: vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi n 2ffi u n P u P 2 un Xi 2 Xi t i51 i51 ½1:2:A2:10 s5 n ðn 2 1Þ THE STANDARD ERROR OF THE MEAN ðXi 2XÞ2 n P n P THE SAMPLE STANDARD DEVIATION n n P 2 Xi2 2 2 n X 1 n X Þ 5 n21 n21 Multiplying both numerator and denominator by n, we reach THE SAMPLE VARIANCE ½1:2:A2:4 2 Inserting this into Eqn [1.2.A2.5] gives ½1:2:A2:7 s2 5 Xi The variance is a measure of the spread of a population. It is the average of the squared deviations from the mean. The square is taken so that all deviations from the mean contribute to the variance; otherwise, negative deviations would cancel positive deviations: the average deviation from the mean is zero. For a population the variance is given by ½1:2:A2:3 Xi 1 n X n21 Recognizing from Eqn [1.2.A2.1] that THE POPULATION VARIANCE ½1:2:A2:2 n P i51 THE MEAN ½1:2:A2:1 This is rewritten as The standard deviation of a sampling distribution of a statistic is called the standard error of that statistic. A sampling distribution of a statistic is its probability distribution. What this means is that if we estimate the mean of a population several times using different random samples, we will not get identical answers. There will be some distribution of the mean. In the same way, there will be some variation in our sample variance calculated from the different samples. The standard deviation of the distribution of means is the standard error of the mean. Note that the sample standard deviation does not change with the sample size, because each additional member of the sample contributes both to n and to the squared deviation from the mean. On the other hand, we ought to expect that our estimate of the mean improves with the number of observations in the sample. Thus, the standard error of the mean gets smaller as the number of observations in the sample increases. The formula for the standard error of the mean is ½1:2:A2:11 s SEM 5 pffiffiffi n 25 26 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION PROBABILITY Probability theory began with the analysis of games of chance. The probability of an event happening, such as drawing a particular card, or a card that belongs to a particular suit, or the appearance of a particular face on a die after a roll, is defined as the ways in which the event can happen compared to the total possible number of outcomes. For example, there are six faces on a die and the likelihood that any particular face ends upright is one out six, or 1/6 5 0.1625. We can write this as P ð4Þ 5 0:1625; Pð5Þ 5 0:1625; P ð6Þ 5 0:1625 where P(i) refers to the probability of the face with i dots (pips) on it. If we have two dice, the probability of having a total number of dots depends on how many outcomes produce that number of dots. For example, we can find 12 dots only by having 6 on one die and 6 on the other. There are a total of 36 possible outcomes as shown in Table 1.2.A2.1 below. Since there are 36 possible outcomes, the probability of getting a 12 is 1/36 5 0.0278. The probability of getting a total of 7, however, is higher. We can get a 7 from the following combinations: {(4,3), (3,4), (5,2), (2,5), (6,1), and (1,6)}. The probability of getting a 7 is thus 6/36 5 0.1625. Note the probability of getting some outcome is 1.0, and this is the sum of each of the probabilities for the individual outcomes: i5n X The validity of these conclusions depends on fair dice that are independent. A die is fair if and only if the probabilities of all six outcomes are the same, and they are independent if and only if the outcome of one die does not influence the outcome of the other. HYPOTHESIS TESTING Pð1Þ 5 0:1625; Pð2Þ 5 0:1625; P ð3Þ 5 0:1625; ½1:2:A2:12 there are two ways to get (4,3), without regard to order. In this case P(4,3) 5 P(3,4) and the probability of getting a 7 with a 4 and a 3 is P{4,3} 5 P(4,3) 1 P(3,4) 5 2/36. PðiÞ 5 1:0 i51 where there are a total of n distinguishable outcomes. The probability of observing a particular outcome for a roll of the dice is the product of the individual outcomes for each die. Thus, the probability of a {4,3} outcome is given as Pð4; 3Þ 5 Pð4Þ 3 Pð3Þ 5 1=6 3 1=6 5 1=36 Here we are making the distinction of which die shows 4 and which shows 3. For example, if the dice were colored this would correspond to a red die with a 4 and a white die with a 3. If they are not distinguishable, then Often in scientific investigations or engineering it is necessary to obtain data to test whether or not some treatment had an effect, or whether or not the data fit some equation whose derivation is based on some theory. The derivation of the equation may have required some assumptions that limit the conditions under which the equation is valid. An excellent fit of the data to the predicted values based on the theory gives us confidence that the theory is valid and the assumptions used were met. There is no guarantee that this is so. The philosophy of science tells us that all we can do is disprove hypotheses. If the data do not agree with the theory, one of two things must be true: either there is something “wrong” with the data or there is something wrong with the theory. The data must be taken at face value but perhaps the data is not what you think it is. For example, suppose you want to investigate the relationship between pressure and flow at steady-state and you make measurements of the flow and pressure in a system. You make some measurements, but do not realize that the system is not at steady-state. Thus the data cannot be analyzed using an equation that assumes steady-state. The data are not “wrong,” but you are wrong in thinking that the data represent the steady-state condition. Failure of the equation to match your data is not a problem with the data or the validity of the data, it is a problem of trying to fit data to conditions that do not pertain to how the data was actually obtained. Other possible problems exist. For example, perhaps you measured the pressure with an incorrectly calibrated meter, so the pressures that you report are not the true pressures. There is generally far more ways of making mistakes than there are ways of making the measurements correctly. TABLE 1.2.A2.1 Possible Outcomes from a Roll of Two Dice Outcome of First Die Outcome of Second Die 1 2 3 4 5 6 1 (1,1) (2,1) (3,1) (4,1) (5,1) (6,1) 2 (1,2) (2,2) (3.2) (4,2) (5,2) (6,2) 3 (1,3) (2,3) (3,3) (4,3) (5,3) (6,3) 4 (1,4) (2,4) (3,4) (4,4) (5,4) (6,4) 5 (1,5) (2,5) (3,5) (4,5) (5,5) (6,5) 6 (1,6) (2,6) (3,6) (4,6) (5,6) (6,6) Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow TABLE 1.2.A2.2 Type I and Type II Errors Inference Null Hypothesis True False Reject H0 Type I error Correct inference Accept H0 Correct Inference Type II error So one kind of hypothesis is this: the derived equation is a good fit for the data. We can call this the “null hypothesis,” H0: there is no difference between the data and what is predicted from the equation. We suppose that the data is not a perfect fit. There are differences between the data and the predicted value for the data. If the differences are small, we may think that the fit is good and we are likely to accept the null hypothesis as being true. If the differences are large, we begin to think something is “wrong” with either the data or the equation, and we reject the null hypothesis as being false. It is possible to make two distinct types of errors in this testing of hypotheses. The Type I error is the incorrect rejection of a true null hypothesis. The Type II error is the failure to reject a false null hypothesis. These types of errors are clarified in Table 1.2.A2.2. Another way of describing these is that the Type I error is a false positive: you conclude there is an effect when there is none present. A Type II error is a false negative. You fail to detect an effect when it is actually present. The probability of making a Type I error is designated α. The probability of making a Type II error is designated as β. α is generally accepted as the level of significance of a statistical test of H0. This is to say, we have to agree on how different the data can be from what is predicted from the null hypothesis before we can reject the null hypothesis. There is some probability that the differences that you obtained arose by random chance, and this is the probability of making a Type I error. In other types of experiments we may want to know if some treatment or variable has some effect in a population of individuals. Suppose we do an experiment in which we measure some variable X for a set of individuals who are not treated, the control group, and we measure the same variable for a set of individuals who receive some treatment. Did the treatment affect the value of X? What we generally have is a set of measurements of variable X from the control group G0 5 {X01, . . ., X0n} and a second set of measurements from the experimental group Ge 5 {XE1, . . ., XEn}. The mean for G0 is designated μ0 and the mean for the experimental group is μE. The null hypothesis is H0: μ0 5 μE. Generally the means will not be exactly equal. There is some probability that a given difference μ0 2 μE will arise by chance. The t-statistic for comparison of two means is given as μ0 2 μE ffi ½1:2:A2:13 t 5 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi SEM20 1 SEM2E Clearly, the larger the value of t the more likely that the difference in the means does not arise by chance. If the difference arises by chance, then we would be wrong to reject the null hypothesis. The probability that the difference arises by chance, and we reject the null hypothesis, is the probability of the Type I error, and this is called the level of significance of the test. Typically it is accepted that α 5 0.05, which means there is a 5% chance that we will reject the null hypothesis even though it is true. Values of t for given values of α are published, and these depend on the degrees of freedom for the t-statistic. The degrees of freedom are the number of independent pieces of information that contribute to the estimation of the statistic, which is the t-value in this case. In a two-sample comparison such as the one described here, the degrees of freedom (symbolized as df or ν) is given as n0 1 nE 2 2. We subtract 2 because the means are derived from the data and so the means and the values themselves are not independent, and we have used 2 means in the calculation of the t-statistic. We could apply these or other types of statistical test to groups of people to compare them and to make statistical statements about them, such as “men are more prone to heart attacks than women”, or “black american males have a higher incidence of hypertension than white american males,” or “obese white females with multiple children are most prone to gallstones.” These statistical statements mean nothing when you are faced with a single incidence of the population: population averages and trends tells you nothing about any specific member of the group. You don’t know, without making some measurements, how far away from average this specific person will be. However, we can make general statements about the populations that may guide us in understanding or treating a condition or setting public health policy. THE NORMAL PROBABILITY DENSITY FUNCTION A probability density function is one in which the probability of an outcome being in some interval is the product of the density function and the interval. Mathematically, it is given as ðb ½1:2:A2:14 Pða # x # bÞ 5 pðxÞdx a where P(a # x # b) is the probability of x being between a and b and p(x) is the probability density function. There are several kinds of probability distribution functions. We will derive the normal or gaussian probability density distribution by considering a random throw of a dart aimed at the origin of a Cartesian plane. We make some basic assumptions of the probability of the dart landing in a particular area. These are: G the distribution is symmetrical: the probability of being high by some distance is the same as being low by the same distance, and this is equal to the probability of being right by the same distance, or by being left by the same distance. In fact, the probability of being off-center by some distance, r, is independent of the angle θ from the origin. 27 28 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION This differential equation is true for all x and y, and x and y are independent. This can happen only if the ratio defined on each side of the equation is constant. We let y dpðxÞ dx 5 C pðxÞx Δy r θ Δx dpðxÞ 5 C x dx pðxÞ ð ð dpðxÞ 5 C x dx px x ½1:2:A2:20 FIGURE 1.2.A2.1 Cartesian coordinate and polar coordinate systems for the analysis of the normal probability distribution function. The probability of a dart hitting the gray area, dxdy, is the probability of it landing in the vertical stripe of width Δx times the probability of it landing in the horizontal stripe with width Δy. G G G errors in particular directions are independent. That is, the probability of being off-center by x in the horizontal direction does not alter the probability of being off-center by y in the vertical direction. large errors are less likely than small errors. our aim is unbiased. That is, on average, the distance from the origin, taking x and y as positive and negative, is zero. Consider the Cartesian coordinate system shown in Fig. 1.2.A2.1. The probability of the dart landing in the vertical stripe between x and x 1 Δx is given by p(x)Δx, where p(x) is the probability density function. Similarly, the probability of the dart landing in the horizontal stripe between y and y 1 Δy is p(y)Δy. What we want to do is determine the mathematical form of the function p. From the independence assumption, the probability of the dart landing in the shaded region is the product of the probabilties of landing in the horizontal or vertical stripes: p (x)Δxp(y)Δy. Because of symmetry, the probability of the dart landing in the area the size of ΔxΔy located r distance from the origin does not depend on θ and we can write ½1:2:A2:15 gðrÞΔxΔy 5 pðxÞΔx pðyÞΔy which gives gðrÞ 5 pðxÞ pðyÞ Differentiating with respect to θ, we obtain dgðrÞ dpðyÞ dpðxÞ 5 0 5 pðxÞ 1 pðyÞ dθ dθ dθ dg(r)/dθ 5 0 because g(r) is independent of θ. Inserting x 5 r cosθ and y 5 r sinθ, we can re-write Eqn [1.2.A2.17] as 0 5 pðxÞ ½1:2:A2:18 ½1:2:A2:19 dpðyÞ dr sinθ dpðxÞ dr cosθ 1 pðyÞ dy dθ dx dθ dpðyÞ dpx r cosθ 2 pðyÞ r sinθ 0 5 pðxÞ dy dx dpðxÞ dpðyÞ dx 5 dy pðyÞy pðxÞx x2 1c 2 Taking the exponent of e of both sides of the last equation gives k 2 pðxÞ 5 Ae2 x ½1:2:A2:21 Because large x is less likely than small x, we know that C must be negative and we can re-write this equation as k 2 pðxÞ 5 Ae2 2 x ½1:2:A2:22 We can evaluate A by the requirement that the total probability for all outcomes is 1.0. This is expressed mathematically as ðN ½1:2:A2:23 k 2 2N Ae2 2 x dx 5 1:0 Since A is a constant, it can be removed from the integrand. Due to the symmetry of the problem, this integral is twice the integral from zero to infinity. We write this as ðN ½1:2:A2:24 k 2 e2 2 x dx 5 0 ½1:2:A2:16 ½1:2:A2:17 ln pðxÞ 5 C 1 2A We can combine this with the distribution about y, as these are symmetrical: ½1:2:A2:25 ðN k 2 e2 2 x dxU 0 ðN k 2 e2 2 y dy 5 0 1 4A2 Since x and y are independent, we can re-write this product of integrals as the integral of their products: ½1:2:A2:26 ðN ðN 0 0 k 2 2 1 e2 2 ðx 1y Þ dx dy 5 4A2 This can be converted to polar coordinates, recognizing that (x2 1 y2) 5 r2 and dxdy 5 rdrdθ. We obtain ½1:2:A2:27 ðπ ðN 2 0 0 k 2 e2 2 r r dr dθ 5 1 4A2 Phys ical Foundat ions of P hysiology I: P ressure-Dr iven Flow Evaluation of the interior integral is 21. Eqn [1.2. A2.28] thus becomes ðπ 2 1 π 1 dθ 5 5 ½1:2:A2:29 2k 4A2 0k The value of A is thus given as rffiffiffiffiffiffi k A5 ½1:2:A2:30 2π The probability distribution in Eqn [1.2.A2.22] becomes sffiffiffiffiffiffi 2 k x2 2 k ½1:2:A2:31 pðxÞ 5 e 2π We can evaluate k from the variance of the probability distribution. The average value for x is given as ðN μ5 ½1:2:A2:32 x pðxÞ dx 2N The variance, similar to our earlier definition in Eqn [1.2.A2.2] is given as ðN ½1:2:A2:33 ðx2μÞ2 pðxÞ dx σ2 5 2N Because of the symmetry of the coordinates, and the fact that p(x) 5 p(2x), we know that the mean, μ, is zero. With Eqn [1.2.A2.31], and μ 5 0, the variance is given as k 2 sffiffiffiffiffiffi k ÐN 2 22x 2 ½1:2:A2:34 σ 52 x e dx 2π 0 We use symmetry as we did before to integrate from 0 to N: sffiffiffiffiffiffi k 2 k ÐN 2 22x 2 ½1:2:A2:35 x e dx σ 52 2π 0 We can integrate this by parts by identifying ½1:2:A2:36 k 2 22x u 5 x dv 5 x e k 2 1 22x dx v 5 2 e k Substituting these into Eqn [1.2.A2.35] we obtain sffiffiffiffiffiffi ð N k 2 22x k 2 x2 e dx σ 52 2π 0 sffiffiffiffiffiffi ð N k 2 σ 52 udv 2π 0 sffiffiffiffiffiffi N ð N k 2 uv 2 vdu σ 52 2π 0 0 3 sffiffiffiffiffiffi2 k 2 ð N 2 k x2 2 2 x N 2 k x 1 1 42 e σ2 5 2 e dx5 2π k 0 k 0 ½1:2:A2:37 0.05 μ=0 0.04 σ = 10 μ = 50 0.03 P(x) We can evaluate the integrals by making the substitution u 5 2kr2/2. It follows that du 5 2krdr, and thus 2du/k 5 rdr. Making these substitutions in the interior integral, we derive ð 2N ðπ 2 21 1 ½1:2:A2:28 eu du dθ 5 4A2 0 k 0 σ = 15 0.02 σ = 25 0.01 σ = 50 0.00 –150 –100 –50 0 50 100 150 X FIGURE 1.2.A2.2 Gaussian probability distribution functions for varying values of σ and μ. Different values of μ move the distribution to the right or to the left; different values of σ influence the shape of the curve, with smaller values resulting in sharper curves and larger values creating more spread out curves. The first term in brackets is evaluated from x 5 0 to x 5 M in the limit as M-N, and it is zero. The second term in the brackets has already been done in this derivation (Eqn [1.2.A2.24] and Eqn [1.2.A2.30]). This last equation in Eqn [1.2.A2.37] becomes rffiffiffiffiffiffi pffiffiffiffiffiffi k 1 2π 1 2 pffiffiffi 5 ½1:2:A2:38 σ 52 01 2π k2 k k 2 This last equation gives k 5 1/σ . Inserting this into the probability density function (Eqn. [1.2.A2.31]) gives 1 2 pðxÞ 5 2 pffiffiffiffiffiffi e22σ ½1:2:A2:39 σ 2π This is the normal probability distribution centered at μ 5 0. The general equation with mean μ is achieved by a horizontal shift in x: ½1:2:A2:40 2 1 pðxÞ 5 2 pffiffiffiffiffiffi e σ 2π ðx2μÞ2 2 σ2 Calculated Gaussian or normal probability distribution functions are shown in Figure 1.2.A2.2. Note that the value of σ determines the spread of the distribution. Smaller values of σ result in a sharper distribution. LINEAR REGRESSION For linear regression, the data we’re interested in is generally numerical and comes in sets of ordered pairs: {(x1,y1), . . ., (xn,yn)}. What we desire to know is what is the best linear fit to the ordered pairs? There are many different ways of doing this. What we are going to go over is the least squares linear regression. The first thing we are going to do is to make a few assumptions. These are: 1. there is no error in the values of Xi. All of the error is in the values of Yi 2. the values of Yi are distributed normally. That is, they obey the normal probability distribution. This is the same as the Gaussian probability distribution and is commonly referred to as the “bell curve”. This curve has been described in Eqn [1.2.A2.40] 29 30 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION We introduce a model equation Y^ 5 m X 5 b ½1:2:A2:41 where Ŷ is the predicted value of Y by this model equation. Since we have assumed that the values of X are determined perfectly, with no error, all of the error is in the set of Yi and how it differs from the predicted values, Ŷi. What we want to do is minimize the squared error between the set of Yi and Ŷi. The total squared error is X Error 5 ½1:2:A2:42 ðYi 2 Y^ i Þ2 i Substituting in from Eqn [1.2.A2.41] we have X Error 5 ½1:2:A2:43 ðYi 2mXi 2bÞ2 i The error estimated as the square of the deviations from the observed values, the set of Yi, from the predicted values, mXi 1 b, varies with m and b used in the calculations. We require this sum of square errors to be minimized in order to have the best fit line. We achieve this by looking for minima in the square errors as we vary m and b. That is, we look for the m and b that has the least squared error. This occurs when the partial derivatives of the error with respect to m and b are minima: P @ i ðYi 2mXi 2bÞ2 50 @m ½1:2:A2:44 P @ i ðYi 2mXi 2bÞ2 50 @b These two form a pair of simultaneous equations in m and b; all of the pairs of (xi, yi) are known. Expanding the square term, we get P @ i ðYi2 2 2mYi Xi 2 2Yi b 1 m2 Xi2 1 2mbXi 1 b2 Þ 50 @m P @ i ðYi2 2 2mYi Xi 2 2Yi b 1 m2 Xi2 1 2mbXi 1 b2 Þ 50 @b canceling out the common factors of 2, this is recognized as a system of two simultaneous equations in two unknowns: X X X Xi2 m 1 Xi b 5 Yi Xi ½1:2:A2:47 i i X Xi m 1 n b 5 i i X Yi i This system of simultaneous equations in two unknowns (m and b) can be solved by application of Cramer’s Rule: P P i Xi Y i i Xi P n i Yi P P ½1:2:A2:48 5m i Xi2 i Xi P X n i i and the expression for b is P 2 P i Xi i Xi Yi P P i Xi i Yi P 2 P 5b ½1:2:A2:49 i Xi i Xi P n i Xi Evaluating the determinants in the numerator and denominator for m in Eqn [1.2.A2.48], we get the computational formula for the slope of the least-squares best fit line: P P P n Xi Y i 2 Xi Yi ½1:2:A2:50 m 5 Pi 2 P i P i n i Xi 2 i Xi i Xi ½1:2:A2:45 Evaluating the determinants from Eqn [1.2.A2.49] we get P P P 2P Xi Yi 2 Xi Xi Y i b5 i P i2 P i P i ½1:2:A2:51 n i X i 2 i Xi i Xi Performing the partial differentiation gives X X X 2m Xi2 1 2b Xi 2 2 Y i Xi 5 0 By algebraic manipulation, it can be shown that this last equation for the intercept of the least-squares line is given also as ½1:2:A2:46 i 2m X i i Xi 1 2nb 2 2 X i i Yi 5 0 ½1:2:A2:52 b 5 Y 2 mX Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current Learning Objectives G G G G G G G G G G G G G Write Coulomb’s law for electrostatic forces Define electrical potential at x as the work done moving a unit positive charge from infinity to x Write three equivalent but different descriptions of a conservative force Define electric field as the electric force per unit charge Describe the electric field as the negative gradient of the potential Recognize Gauss’s law Write the formula for capacitance in terms of charge and voltage Write the formula for capacitance in terms of area, dielectric constant, and plate separation Describe how capacitance varies with area, dielectric, and plate separation Be able to calculate the capacitance of biological membranes given k, δ, and physical dimensions Be able to calculate electric field intensity and force on a charged particle given V(x, y, z) Write Kirchhoff’s Current Law and Kirchhoff’s Voltage Law Be able to calculate the time constant for a simple RC circuit. COULOMB’S LAW DESCRIBES ELECTRICAL FORCES Electric charge is a fundamental property of some subatomic particles. Electrons have negative charge and protons have positive charge. These designations of positive and negative are arbitrary but rigidly accepted by convention. Separated electrical charges in a vacuum (see Figure 1.3.1) experience a force that is described by Coulomb’s law: F5 ½1:3:1 F1 on 2 5 q1 q2 4πε0 r 2 q1 q2 r12 5 2F2 on 1 4πε0 r 3 The bold face symbols signify vector quantities. F is the force; q1 and q2 are electrical point charges, in coulombs, that are separated by the distance r, in meter; ε0 is a constant, the electrical permittivity of space, which has the value of 8.85 3 10212 C2 N21 m22. The lower © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00003-3 1.3 equation signifies that the direction of the force is along the line between the two charges. The magnitude of the force is proportional to the product of the charges and inversely proportional to the square of their separation. Its sign depends on the signs of the point charges. Two positive charges, or two negative charges, result in a positive force which is repulsive and directed away from their center as indicated in Figure 1.3.1. Two charges of opposite sign experience a negative force which is attractive along a line connecting them. If the intervening space is not vacuum, but some medium, the equation is altered slightly by the inclusion of a dielectric constant, κ, whose value depends on the medium: ½1:3:2 F1 on 2 5 q1 q2 r12 κ4πε0 r 3 The dielectric constant for the vacuum is 1.0. For all other materials, κ . 1.0. The reduced force in the presence of a dielectric material is due to charges present in the material that reorient themselves in the presence of the external point charges and thereby screen the charges from each other. Materials with asymmetric charge distributions within their materials typically have large dielectric constants. THE ELECTRIC POTENTIAL IS THE WORK PER UNIT CHARGE Suppose there is a positive charge of magnitude qfixed fixed in space at some location. We have another charge, a unit positive charge, located infinitely far away so that the force between the charges initially is effectively zero. If we bring the unit positive charge qtest toward the fixed charge at a constant velocity, then its kinetic energy does not change. As we approach the fixed charge, the repulsive force becomes larger and larger and we must apply an external force to keep the qtest at constant velocity. Because our applied external force, Fext, has moved through a distance, we have performed work on the body, given as ðf ½1:3:3 Worki.f 5 Fext Uds i This work is the amount of energy we have expended in moving the positive qtest toward the positive qfixed. Where did that energy go? If we release qtest, we find that it moves away from qfixed and gains kinetic energy which is exactly equal to the energy we used to move 31 32 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION qtest toward qfixed. We say that the energy we used to move qtest was stored as potential energy, U. We define the potential at any place A in space as being the work done to bring a positive unit charge from infinite separation to point A: ðA WorkN.A Fext Uds ½1:3:4 5 UA 5 qtest N qtest The unit of potential here is joules coulomb21 5 volts. The usefulness of the potential is that the work can be determined easily by multiplying the potential times the charge. Fext in this equation is the external force required to move the positive test charge with no change in velocity. It is exactly equal to and opposite in sign to Fint, the interacting electrostatic force. This is, in turn, given by Coulomb’s law (see Eqn [1.3.1]). Since it is directed along ds, we can write ðA ðA Fint Uds qqtest 52 dr UA 5 2 2 q 4πε test 0 r qtest N N ½1:3:5 1 q UA 5 4πε0 rA F2 on 1 F1 on 2 + q1 r12 F2 on 1 + q1 THE IDEA OF POTENTIAL IS LIMITED TO CONSERVATIVE FORCES THE CONSERVATION OF ENERGY THEOREM STATES THAT ENERGY MAY BE CONVERTED BUT NOT DESTROYED The First Law of Thermodynamics is the conservation of energy theorem. It states that in ordinary mechanical events, the total energy is constant. It is written in differential form as + q2 F1 on 2 r12 From this definition of the potential, it should be clear that the potential surrounding a positive charge is positive: it takes work to bring a positive charge toward it. The potential surrounding a negative charge is negative, as we can get energy out of bringing a positive charge toward it. These lead us to an important conclusion: a separation of charge produces an electric potential. The potential defined in this way is a scalar quantity, having magnitude but not a direction, whereas the electrical force is a vector. This comes about from integrating the dot product of Fint with ds where ds is the distance increment that points along the pathway taken from infinite separation to point A. The dot product means that we add only those components of the force that are directed on the line connecting the centers of the charges. These conclusions are illustrated in Figure 1.3.2. – q2 ½1:3:6 FIGURE 1.3.1 Electrical forces between separated point charges. q indicates charge. Like charges repel, so that the force of q1 on q2 is directed away from q1 on a line connecting their centers. The force of q1 on q2 is exactly opposite to the force of q2 on q1. Unlike charges attract with forces opposite but in line with the vector connecting their centers. dE 5 dq 2 dw where E is the total energy, q in this case is the heat energy, and w is the work. This is another unfortunate case where variables are used to denote completely different quantities. In thermodynamics, q symbolizes The potential at A is the work required to bring a unit positive charge from infinity to A; it is independent of path For a positive fixed charge, the energy required to bring a positive charge to A is positive: the potential around a positive charge is positive A + + + + + + + qfixed FIGURE 1.3.2 Definition of the electrical potential. The potential at a point A is defined as the work required to bring a unit positive charge (qtest) from infinite separation to point A. If there is a fixed positive charge near A, it takes work to bring qtest to A (we must apply a force to overcome the repulsive force and we move that force through a distance) and the potential is positive. If there is a fixed negative charge near A, then qtest is attracted to it and it takes energy (work) to slow qtest—the work is negative because the applied force is opposite to the direction of movement. - - - - - qfixed Fapplied Felectrical + For a negative fixed charge, the energy required to bring a positive charge to A is negative: the potential around a negative charge is negative A qtest Felectrical Fapplied + qtest Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current heat, and in electrostatics it symbolizes charge. The appearance of heat in this equation is extremely important, because it turns out that there are theoretical limits in the conversion of heat energy to useful work, and this gives rise to the concepts of entropy and free energy. This equation assumes that heat and work are alternate forms of energy. We take the equivalence of mechanical, thermal, chemical, and electrical energies for granted but historically this idea took some time to develop. In writing that any energy change in the system is the balance between work output and heat input, it is assumed that work is equivalent to heat. The equality of mechanical work and heat was established in 1845 by Joule. The signs of dq and dw in this equation are important, and are consequences of the definitions of heat and work. The quantity dq is defined as the heat absorbed by the system from its surroundings, and dw is defined as the work done by the system on its surroundings (see Figure 1.3.3). The “system” here is anything we have drawn a conceptual line around, usually in agreement with some physical boundary, that sets part of the universe off from the rest of it. In the case of electrostatics, the system is the set of charges distributed in space. THE WORK DONE BY A CONSERVATIVE FORCE IS PATH INDEPENDENT The work done by moving qtest toward qfixed is the work done by the surroundings (us) on the system of interacting charged particles. It is equal but opposite in sign to the work done by the interacting force. There may be heat generated by the necessity to apply more force than the interacting force in order to overcome friction, if the charges move through some medium, but this is separate from the interacting force (the coulombic or electrostatic force) itself. The coulombic force itself generates no heat at all. It belongs to a class of forces called conservative forces that do not dissipate energy as heat. Conservative forces are characterized by three equivalent statements: 1. The work done by a conservative force depends only on the initial and final positions, and not on the path (see Figure 1.3.4). 2. The potential difference between two points depends only on the end points and not the path. 3. The total work done by a conservative force acting around a closed loop is zero. The positive increment in heat, dq, is defined as the heat taken up by the system from its surroundings Heat energy Work energy dq + dw + System dE = dq – dw Surroundings The positive increment in work, dw, is defined as the work done by the system on its surroundings FIGURE 1.3.3 Theorem of the Conservation of Energy. The system is any part of the universe that we have enclosed by some boundary, real or imagined. Positive heat flow is defined as heat energy that is absorbed by the system from its surroundings. Similarly, positive work is defined as work that is done by the system on its surroundings. By these definitions, conservation of energy means that dE 5 dq 2 dw. In order to write this equation, it is assumed that heat and work have the same units, that of energy. Here work can be electrical, mechanical, or chemical. Components of path normal to r require no force and entail no work + q2 + q1 Components of the path along r contribute F dr to the work A ... so the work required to get to point A depends only on the radial separation, and not the path FIGURE 1.3.4 Potential is independent of the path. Any path from start to finish can be successively approximated by a series of paths oriented either parallel to the vector connecting the charges or perpendicular to it. Those components of the path perpendicular to the vector require no force and therefore contribute nothing to the potential at point A, which is defined as the work necessary to bring a unit positive charge (here shown as q2) from infinite separation to point A. Components of the path oriented parallel to the vector connecting the point charges contribute F dr to the force. Therefore, the total work (potential) moving the charge depends only on the radial separation and not the path taken. 33 34 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION POTENTIAL DIFFERENCE DEPENDS ONLY ON THE INITIAL AND FINAL STATES If the potential depends only on the position, then it is a state function, one that is independent of path and dependent only on the state of the system. Thus we can associate a potential with any position, A and B: ðA Fint Uds UA 5 2 N qtest ðB Fint Uds UB 5 2 ½1:3:7 N qtest ðB Fint Uds UA.B 5 2 5 UB 2 U A A qtest If A 5 B, where A is the initial state and B is the final state, we can write I final Uinitial.final 5 0 5 2 ½1:3:8 Fint Uds initial This is the mathematical statement that the work performed by the system around any closed loop is zero. This turns out to be equivalent to the statement that the potential is a function of position (state) only and not of the path used to get to that position. THE ELECTRIC FIELD IS THE NEGATIVE GRADIENT OF THE POTENTIAL The electric field intensity is defined as the electric force per unit charge: E5 ½1:3:9 Fint qtest Insertion of this into Eqn [1.3.7] and differentiating, we get ½1:3:10 E52 dU dS This equation is not correct as written yet, because we have a vector (the electric field intensity) on one side and a scalar on the other! We need to take a particular kind of derivative, the gradient, to convert the scalar potential into a vector force. Equation [1.3.10] is correct as written as long as the axis of ds corresponds with the direction of F. The full three-dimensional vector equation is @U @U @U 2j 2k @x @y @z E 5 2 rU where i, j, and k are unit vectors in the x, y, and z directions. The expression on the right-hand side of Eqn [1.3.11] is called the gradient of the function U. It is a vector whose components on each axis are ½1:3:11 E52i the slope of the potential projected onto that axis. Generally, the gradient is a vector that does not align with any axis. Instead, it points in the direction of the steepest slope of the potential surface in three dimensions. The force points down this slope. It is the negative of the gradient of the potential. The last equation shows the gradient written in operator notation. The operator r is called del and is defined as ½1:3:12 r5i @ @ @ 1j 1k @x @y @z FORCE AND ENERGY ARE SIMPLE CONSEQUENCES OF POTENTIAL The usefulness of potential is that it simplifies the idea of force and energy. The force on a charge q is given easily by multiplying it times the electric field, which is 2 grad U. The energy cost in moving a charge from one potential to another is just qΔU: ½1:3:13 F 5 2 qrU Δ Energy 5 qΔU where r is the del operator and Δ signifies the difference between final and initial states. The electric potential, U, here is in units of volts. If charge is in coulombs, the force is in units of coulomb-volt per meter; a voltcoulomb is a joule 5 1 N m 5 1 kg m2 s22; therefore, the force is in units of 1 N m/m 5 N. The energy is in units of joules. From now on we will abandon use of U as a symbol for the potential; physiologists typically use V, E, or ψ as symbols of potential. GAUSS’S LAW IS A CONSEQUENCE OF COULOMB’S LAW Gauss’s law is written as I q EUds 5 ½1:3:14 ε0 where the integral is taken over any closed surface of the dot product of the electric field and the area vector, equal to the area increment ds and oriented perpendicular to the surface. What this says is that this dot product, summed over any closed surface, is equal to the charge enclosed by the surface divided by ε0, the electrical permittivity of space. If there is no enclosed charge, the surface integral is zero. This equation is a variant of Coulomb’s law (see Eqn [1.3.1]). To see how this equation works, we consider a spherically symmetrical distribution of positive charges as shown in Figure 1.3.5. The evaluation of the surface integral is simplified by choosing an appropriate surface. In this case, we choose a sphere centered on the symmetrical charge. By symmetry, the electric field is directed radially outward, pointing along the vector ds. Similarly, the electric field is everywhere constant in magnitude at a Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current Area = A Electric field lines are radially directed and symmetrical E E = kq r2 ds Gaussian surface + + + + + ++ + + + δ r E FIGURE 1.3.6 A parallel plate capacitor. Two plates, each of area A, are separated by a distance δ. They are charged by connecting them to a battery that produces a capacitance current until the potential difference between the two plates is equal to that across the two poles of the battery, so that the net potential difference across the entire circuit loop is zero. At this point, there is no more current flow. The separation of charges produces a uniform electric field between the two plates. FIGURE 1.3.5 Electric field surrounding a spherically symmetrical distribution of positive charge. ds is a vector having a magnitude of the area increment and directed normal to the closed surface. In this case, we take the Gaussian surface, indicated here by a dashed line, to be a sphere centered on the symmetrically distributed charge. The electric field vector and the surface normal vector are pointing in the same direction, so that the angle between them, θ, is zero and the dot product of E and ds is E ds, because cos θ 5 1. In the plate, E = 0 Area = A δ prescribed distance, r, from the center of the charged body. Thus we can write I E q EUds 5 E4πr 5 ε0 q E5 4πr 2 ε0 2 ½1:3:15 which is the magnitude of the electric field (E 5 F/q, the electric force per charge) from Coulomb’s law (see Eqn [1.3.1]). THE CAPACITANCE OF A PARALLEL PLATE CAPACITOR DEPENDS ON ITS AREA AND PLATE SEPARATION As described in Chapter 1.2, the ability to store electric charge is characterized by the capacitance, defined as ½1:3:16 C5 Q V where C is the capacitance, Q is the charge, and V is the potential, in volts. We now consider a particular type of device to store charge, a parallel plate capacitor, as described in Figure 1.3.6. The two charged plates will be attracted to each other and so must be held apart by some dielectric material that insulates the plates and keeps the charges separated. There will be some fringing of the electric field FIGURE 1.3.7 Parallel plate capacitor with a Gaussian surface. The Gaussian surface is the box indicated by the dashed lines. The electric field is constant within the capacitor and oriented as shown. The integral of E ds in the plate is zero because E is zero there. The integral of E ds in the dielectric between the plates is EA. The integral of E ds on the sides of the enclosed surface is zero because E and ds are orthogonal in this region. around the edges of the plate, which we shall ignore. The resulting electric field within the capacitor is uniform, which can be proved by integrating the Coulomb force over the uniformly distributed charge on a plane, which we will not do here. We draw a rectangular closed surface, one side of which is in the middle of the dielectric and the other in the middle of the plate. Since the plate is a good conductor, the electric field within the plate is zero—the voltage difference in the plate is zero. The closed surface integral is just the constant electric field times the area of the surface in the dielectric. The situation is illustrated in Figure 1.3.7. Application of Gauss’s law according to the description in the legend of Figure 1.3.5 gives I q EUds 5 EA 5 ε0 ½1:3:17 q E5 ε0 A 35 36 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Since the electric field is uniform, its relation to V from Eqn [1.3.10] is given as ½1:3:18 E52 dV V1 2 V2 5 dx δ The capacitance is calculated as ½1:3:19 C5 q V1 2 V2 Substituting in for q from Eqn [1.3.17] and for V1 2 V2 from Eqn [1.3.18], we have ½1:3:20 C5 ε0 A δ The presence of a dielectric between the plates reduces E for a given charge, and therefore increases the capacitance. The formula for parallel plates with a dielectric is ½1:3:21 κε0 A C5 δ where κ is the dielectric constant, a dimensionless ratio. According to Eqn [1.3.21], the capacitance increases linearly with the area and inversely with the separation between the plates, and is increased by materials with high dielectric constants. BIOLOGICAL MEMBRANES ARE ELECTRICAL CAPACITORS Biological membranes share some of the features of parallel plate capacitors and act as electrical capacitors. Their structure is detailed in Chapter 2.4. Briefly, biological membranes consist of an asymmetric bilayer of lipid molecules that assemble to form an interior insulating core. This effectively separates two plates—the surfaces of each bilayer—from each other. The separation distance is typically quite small, on the order of 7 nm. This bilayer structure is shown in Figure 1.3.8. From it, you can see the resemblance of the bilayer to a parallel plate capacitor. The dielectric constant of some materials is shown in Table 1.3.1. This constant varies with temperature and the chemical make-up of the dielectric. Materials that are polar and mobile, such as water, can orient their partial charges with the electric field, and reduce the field within the dielectric. In this way, more charge can be added to the surfaces of the plates and therefore these dielectrics have a high dielectric constant. TABLE 1.3.1 Dielectric Constant of Some Materials Material Dielectric Constant, κ Air 1.00059 Water 80 Glycerol 43 Acetic acid 6.2 Benzene 2.3 CCl4 2.2 Oleic acid 2.46 Hexanol 13.3 Hexane 1.89 Stearic acid 2.29 Monopalmitin 5.34 Aqueous layer Insulating hydrocarbon layer FIGURE 1.3.8 Lipid bilayer membrane consisting of various lipid molecules arranged with their hydrocarbon tails toward the interior of the bilayer and their water-soluble parts facing the water phase. Aqueous layer Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current EXAMPLE 1.3.1 Capacitance of Planar Lipid Bilayers Dr. Alexandre Fabiato at VCU drilled a narrow and clean hole with a diameter of 250 μm 5 0.25 mm into a Lexan partition that separated two electrolyte solutions. He “painted” some phospholipids over the hole using a Teflon stick cut at an angle and dipped into a solution of lipids dissolved in hexane. After “thinning” (passive removal of the hexane through the aqueous phase), the membranes form a planar lipid bilayer (see Figure 1.3.9). Dr. Fabiato measured the capacitance of the membrane using an AC signal connected to electrodes immersed in the solutions. He derived a capacitance of 350 pF (350 3 10212 F). Calculate the specific capacitance of the membrane Dr. Fabiato made, in F cm22. The specific capacitance is just the capacitance per unit area of membrane: Cm 5 C/A. The measured capacitance is 350 3 10212 F and the area A is πr2, where r 5 0.0125 cm; therefore, Cm 5 340 3 10212 F=4:9 3 1024 cm2 5 0:71 µF cm22 Barrier What Is the Approximate Thickness of the Bilayer? Equation [1.3.19] allows us to calculate the thickness as δ 5 κε0/Cm, where Cm is the specific capacitance. We do not know κ, but the dielectric constant for lipid-like substances has been determined, as examples shown in Table 1.3.1. Here we use ε0 5 8.85 3 10212 C2 J21 m21 and Cm 5 0.71 3 1026 C V21 cm22 that we calculated earlier. Using the dielectric constant for n-hexane as an example, we calculate δ 5 1:89 3 8:85 3 10212 C2 J21 m21 =0:71 26 3 10 CV21 cm22 3 ð100 cm m21 Þ2 5 2:36 3 1029 m The calculated values of δ (Table 1.3.2) are of the same order as expected from electron micrographs of membranes. If the potential across a membrane is 80 mV, and its thickness is 7 nm, What is the electric field intensity? The field is uniform inside a capacitor, and so is given by E 5 2ΔV/Δx, where ΔV is the potential difference and Δx is the separation of the plates. Thus the electric field is E 5 2 80 3 1023 V=7 3 1029 m 5 2 11:4 3 106 V m21 Water Water TABLE 1.3.2 Calculated δ for Various κ Narrow aperture κ δ (nm) 1.89 2.36 2.29 2.85 2.46 3.07 5.34 6.67 FIGURE 1.3.9 Planar lipid bilayer formed in a narrow hole between two aqueous compartments. ELECTRIC CHARGES MOVE IN RESPONSE TO ELECTRIC FORCES As mentioned in the section “Force and Energy Are Simple Consequences of Potential,” the usefulness of the concept of potential lies in the ease of calculating the force on a charged particle or the energy needed to move from one region to another. The electrical force on a charged particle is given as ½1:3:22 F 52qrU 5 qE 5 zeE where U is the potential, often written as V, q is the charge, E is the electric field, z is the valence (1/ 2 integral number of charges per particle), and e is the unit charge of the electron. Thus, a charged particle, of either sign, in an electric field is subjected to an accelerating force. Ions in solution are subjected to these forces and accelerate on account of them. These ions accelerate until they reach a terminal velocity, v, at which point the electrical force is matched by a drag force on the particle by the surrounding solution. Figure 1.3.10 illustrates this situation. MOVEMENT OF IONS IN RESPONSE TO ELECTRICAL FORCES MAKES A CURRENT AND A SOLUTE FLUX The drag force on a particle moving through a solution is proportional to its velocity and directed opposite to it. Further, the electrical force given in Eqn [1.3.22] at 37 38 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION E Fe Fd + V ½1:2:3 q FIGURE 1.3.10 Forces on a charged particle in solution subjected to a constant electric field. The electrical force, Fe, is the product of the charge, q, on the particle and the electric field, E. This electrical force accelerates the charged particle and it moves through the solution. This movement produces a drag force, Fd, which is proportional to the velocity, v. The particle reaches a terminal velocity when the net force on the particle is zero: Fe 1 Fd 5 0. the terminal velocity is equal but opposite to Fd. Therefore, we can write Fd 5 2βv Fe 5 2Fd Fe 5 βv ½1:3:23 where β is a drag coefficient or frictional coefficient. Thus ions subjected to a constant electrical field will move at a constant terminal velocity, v, and that velocity will be proportional to the electric field. This movement of charged particles constitutes a movement of charge from place to place, and so it is an electrical current. Further, because solute particles carry the charge, the movement also forms a solute flow. The solute flux is related to the velocity by ½1:3:24 Js 5 vC where J and v are written as vectors and C is the concentration of the solute (see Figure 1.3.11). Because each solute particle carries the charge ze, the current density is i 5 zeJs ½1:3:25 This expression can be converted into Ohm’s law by using Eqns [1.3.22][1.3.24]: ½1:3:26 i 5 zeJs 5 zevC Fe i 5 zeC β i5 The last equation is an analogue of Eqn [1.2.3] for the one-dimensional form of Ohm’s law: z2 e2 C ð2rVÞ β Je 5 2σ EXAMPLE 1.3.2 Forces on Charged Particles Consider the planar lipid bilayer in Example 1.3.1, which has a potential difference of 80 mV across it. What would the electric force be on a Na1 ion in the middle of the bilayer? The electric force on a charged particle is given as Fe 5 qE. We calculated the electric field intensity, E, in this case to be 211.4 3 106 V m21. The charge on any ion is ze, where z is the valence or integral number of charges on the particle and e is the charge on the electron. In this case, z 5 11 and the charge on the electron is given in various units. The most useful unit here is the coulomb: 1e 5 1.6 3 10219 C. The force is thus given as Fe 5 1:6 3 10219 C 3 2 11:4 3 106 V m21 5 2 1:82 3 10212 V C m21 5 2 1:82 3 10212 J m21 5 2 1:82 3 10212 N m m21 5 2 1:82 3 10212 N THE RELATIONSHIP BETWEEN J AND C DEFINES AN AVERAGE VELOCITY Consider the right cylindrical tube shown in Figure 1.3.11 that contains solute particles moving to the right at average velocity v. In time Δt, the particles travel a horizontal distance Δx 5 vΔt. All of the solute particles in the volume AΔx will have crossed a crosssectional plane in the cylinder. Thus the flux will be the total number of solute particles in that volume, per unit area per unit time. The number of solute In time Δt the solute particles move Δx = vΔt; the flux is therefore C A Δx / AΔt = CΔx/Δt = Cv Cross-sectional area, A J(x) FIGURE 1.3.11 Relationship among J, C, and v. If solutes have an average velocity v, they sweep out a distance vΔt in time Δt, and this corresponds to an entire volume of solute, equal to AvΔt, moving to the right. The number of solute particles in this volume is CAvΔt. The flux is this number per unit area, per unit time: J 5 CAvΔt/AΔt 5 Cv. dψ dx v Δx = vΔt x axis Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current particles in the volume AΔx is CAΔx. Thus the flux is given as ½1:3:27 CAΔx AΔt J 5 Cv J5 From this equation, it is clear that the ratio of J/C defines an average velocity for solute particles. OHM’S LAW RELATES CURRENT TO POTENTIAL According to the discussion earlier, a difference in potential produces a force on charged particles, and the force is proportional to the negative gradient of the potential. The movement of charges in response to a potential makes a current, a flow of charges. It is given as ½1:3:28 I5 Δq Δt The current I is given in amperes 5 coulomb s21. Current is defined as the movement of positive charge, so that the movement of cations (positive ions) constitutes a current in the direction of the flow; the movement of anions (negative ions) makes a current in the opposite direction to the flow. The current due to an ion is related to its flow by ½1:3:29 Ix 5 z ℑ Qx where Ix is the current of ion x, in coulombs s21, Qx is the flow of ion x in mol s21, z is the integral charge per ion (1/ 2 1, 2, ...) and ℑ is the Faraday (9.649 3 104 coulombs mol21 5 6.02 3 1023 electrons mol21 3 1.6 3 10219 coulombs electron21); zℑ converts mol to coulombs. Note that current is an extensive variable, while current density is an intensive variable. The movement of charge through matter-filled space encounters resistance from the matter. Those materials that offer little resistance are called conductors. Other materials, such as membrane lipids and the myelin sheath that surrounds nerve axons, offer high resistance and are called insulators. The current is greater if the potential driving it is greater and is less according to the resistance of the material through which the current flows. This is Ohm’s Law: ½1:3:30 I5 Δψ E 5 R R where I is the current, Δψ is the potential difference, often symbolized as E or V, and R is the resistance. Resistance has the units of ohms 5 volts/amps, symbolized as Ω. Ohm’s law can also be written as ½1:3:31 I5g Δ ψ where g 5 1/R is the conductance. The SI unit for conductance is the siemen 5 amp/volt. A battery is a device for using chemical reactions to create a voltage difference. In effect, chemicals trap electrons, with their negative charges, at fixed distances from their positively charged nuclei. Different chemicals will then have different potentials for their electrons. In chemistry these are called oxidation potentials. These refer to the energy required to remove the electron from the chemical. The movement of electrons from one chemical to another can then release energy equal to the difference in the oxidation potentials times the number of charges that move. If we hook up a battery in series with a resistance, we can produce a current which is given by Ohm’s Law (see Eqn [1.3.30]). This situation is shown schematically in Figure 1.3.12. KIRCHHOFF’S CURRENT LAW AND KIRCHHOFF’S VOLTAGE LAW The circuit shown in Figure 1.3.12 is a simple circuit in which a resistor is placed over the terminals of a battery. When the circuit is completed, current flows and this can be measured with an ammeter. Kirchhoff’s Voltage Law states that the total voltage differences around any loop must be zero. This is a restatement of the conservative nature of the electric force: the work done in any loop is zero. Since the only resistance is between nodes 2 and 4, the voltage differences around the loop are ΔV23 and ΔV71, where ΔV71 is the voltage across the battery from the negative electrode to the positive electrode. Thus, ΔV71 5 2ΔV17 5 2E, the voltage provided by the battery. The total voltage drop around the loop is given as 0 5 ΔV23 1 ΔV71 5 ΔV23 2 E. Solving this for ΔV23, we find ΔV23 5 E and therefore the current through the resistor is I 5 E/R. Kirchhoff’s Current Law states that the sum of current into any node must be zero. Thus if there is a current I23 5 E/R that enters node 3, there must be a current I34 5 I23 that leaves node 3. The total current into node 3 is I23 2 I34 5 0. I34 is negative in this last equation because it leaves node 3—it is the opposite direction (with respect to node 3) of the current I23, but it is equal in magnitude. THE TIME CONSTANT CHARACTERIZES THE CHARGING OF A CAPACITOR IN A SIMPLE RC CIRCUIT Suppose now that we include a capacitor in the circuit shown in Figure 1.3.12. This expanded circuit is shown in Figure 1.3.13. Initially, with the switch open, there will be no potential across the capacitor, and no current in the circuit. If we flip the switch, current will begin to flow because there will be potential differences in the circuit. But the capacitor is filled with a dielectric that disallows current flow! How can current flow across the 39 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Kirchoff's Voltage Law: ΣV R 3 Ammeter – I=0 + 1 – I = E/R 4 E + 5 – I E – 5 2 + + + + R 3 Ammeter Ammeter 4 0 The sum of the voltage drops around any loop is zero R 1 = loop Closing the circuit produces current I = E/R When the circuit is open, there is no current – 40 E – I = E/R I I56 7 6 6 7 I67 Kirchoff's Current Law: ΣI node = 0 The sum of the currents into a node is zero FIGURE 1.3.12 Ohm’s Law, Kirchhoff’s Current Law, and Kirchhoff’s Voltage Law. When the circuit is completed by closing the switch between nodes 1 and 2, current flows as described by Ohm’s Law: I 5 ΔV23/R, where ΔV23 is the voltage difference across the resistor (between nodes 2 and 3) and R is the resistance. In this circuit, the only resistance is between nodes 2 and 3 and all other resistances are negligible. Kirchhoff’s Current Law (KCL) states that the sum of currents into a node is zero. At node 6, for example, the sum of the currents is I56I67, the negative sign indicating that the current is out of the node. Thus, I56I67 5 0 or I56 5 I67. Kirchhoff’s Voltage Law (KVL) states that the sum of voltage drops around any closed loop is zero. For the circuit shown, the voltage drops are from nodes 2 to 3, where ΔV23 5 I R 5 E; and from nodes 7 to 1 where ΔV71 5 2ΔV17 5 2E. Thus the voltage drop around the loop is ΔV23 1 ΔV71 5 E 2 E 5 0. All of the other voltage drops between nodes (ΔV45, ΔV56, ΔV67, and ΔV12) are zero because the resistances in this part of the circuit are negligible. capacitor? The current in this case is a capacitive current, not a resistive current, like the kind shown in Figure 1.3.12. The flow of positive ions onto the top plate of the capacitor produces an electric field (a force per unit charge) that repels positive charges from the bottom plate. This movement of charges away from the bottom plate is the capacitive current. As the charges move away, there is a separation of charges on the capacitor and it now has a potential difference given by V 5 q/C. Charges continue to move until the potential across the capacitor is equal but opposite to the potential across the battery, E. To analyze the time course of current and voltage in the circuit, we make use of Kirchhoff’s Voltage Law that says the sum of the voltage drops in the circuit must be zero. We write ΔV61 1 ΔV23 1 ΔV45 ½1:3:32 2E1 dq q R1 dt C 5 0 5 0 where q is the charge on the capacitor and dq/dt is the current through the resistor, which is also the current across the capacitor. We can separate variables in Eqn [1.3.32] and integrate to solve this equation for q as a function of t. The rate of charging of the capacitor is given as t q 5 EC 1 2 e2RC ½1:3:33 where t is the time and the combined terms RC is called the time constant because it describes the time taken to charge the capacitor. The current can be obtained from differentiation of q to give ½1:3:34 I5 E 2t e RC R The time course of charge and current is shown in Figure 1.3.14. In Eqn [1.3.33], if t 5 RC, then q 5 EC (1 2 1/e), so the time constant is the time required for the charge to be 1/e 5 0.37 of its final value. SUMMARY Some particles in nature either repel or attract other particles, and the force developed between them varies inversely with the square of their separation. These particles are said to be “charged,” and there is no more basic description or explanation of their interaction than Coulomb’s law that quantifies it. Charges have two types: positive and negative. Like charges repel, unlike charges attract. This electrostatic force is a conservative force, meaning that the work performed in moving a charge around any closed loop is zero. Conservative forces also mean that the work done in moving a particle around depends only on the initial and final states, and not the path. Equivalently, the potential energy associated with a distributed set of charges depends only on the position and not on the path it takes to get there. Physical Foundations of Physiology II: Electrical Force, Potential, Capacitance, and Current V=0 voltmeter Area = A - R 3 + 2 1 4 + E * The capacitance is defined as: C 5 Q/V, where Q is the charge and V is the potential difference across the capacitor. The capacitance of a parallel plate capacitor depends on several physical characteristics of the capacitor and is given as E 5 V = V (t) voltmeter Area = A - + + 6 R 3 4 2 1 C5 I + + + + E + * - E - - V=E voltmeter Area = A - + 6 R 3 2 1 4 + + + + + + + + ++ + + + + + + + + * - - - E - - - - E - 5 kε0 A δ where κ is the dielectric constant that depends on material between the plates, ε0 is a physical constant, A is the area of the plates, and δ is the spatial separation of the plates. Biological membranes form capacitors with very small δ. Ions move in response to electrical forces. This movement forms both a solute flux and an electrical current. Movement in response to electrical forces accelerates ions in solution until a terminal velocity is reached. At this point, the net force on the ion is zero and is the balance between the electrical force and the drag or frictional force on the solute particle by the solution. The average terminal velocity is the flux divided by the concentration. Ohm’s Law is given as I 5 E/R. I 5 is measured in volts. The charge is measured in coulombs. Energy is measured in joules or volt-coulombs. Because the potential is defined as the integral of the work to move a unit positive charge from infinity to A, the electric force per unit positive charge is the negative derivative of the potential. The potential is a scalar whereas the force is a vector. The derivative here is the gradient, which converts the scalar potential into a force vector. The electric force per unit positive charge is the electric field. 6 FIGURE 1.3.13 Charging of a capacitor. The capacitor consists of two parallel conducting plates separated by a dielectric, or insulating, material. At the start, top, the circuit is broken by a switch and there is no potential difference across the capacitor. When the switch is closed, middle panel, charge begins to move, making a current. The positive charges on the top plate repel positive charges on the bottom plate, which move back to the battery, completing the circuit for the current. This is a capacitive current, because there is no flow across the dielectric but there is a flow in the circuit. The separation of charges across the capacitor creates a potential difference related to the capacitance of the capacitor: V 5 qC. This builds up as current continues to flow until V is exactly opposite to E. At this point, current flow stops and the capacitor is fully charged. The electrical potential at a point, A, is defined as the work necessary to bring a unit positive charge from infinite separation to that point. Therefore, positive fixed charges are associated with positive potential and negative fixed charges are associated with negative potential. Separation of charge produces a potential. The potential Conservation of charge and the conservative nature of the electric force give rise to Kirchhoff’s Current Law and Kirchhoff’s Voltage Law. Kirchhoff’s Current Law states that the sum of all currents into any node of a circuit must be zero. Kirchhoff’s Voltage Law states that the potential differences around any closed loop must sum to zero. When a capacitor is connected in series with a resistor and a voltage source (a battery), the capacitor gradually becomes charged until the potential across it exactly opposes the potential of the battery. The time course of charging depends on RC, the product of the resistance and the capacitance. When the capacitor is fully charged, current in the circuit goes to zero. The value of RC is the time constant for the circuit, which is the time that the charge takes to get to within 1/e of its final value. REVIEW QUESTIONS 1. What do we mean by “electric potential”? 2. What makes an electrical potential difference between two points? 3. How much energy is gained by charge q over a potential difference V? 4. What is meant by “electric field intensity”? 5. What is the relationship between charge and voltage for a parallel plate capacitor? 41 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION FIGURE 1.3.14 Charging of a capacitor in series with a resistor. Here the specific capacitance was taken as 1 μf cm22 and the specific resistance was taken as 1000 Ω cm2 and the time constant RC 5 1 ms. The charge rises exponentially to reach a maximum and the current simultaneously decreases exponentially from an initial value to zero. The time constant is the time required for q to rise to with 1/e of its final value, which is the same as the time required for I to decrease to within 1/e of its final value or for I to decrease to 1/e I0. 1.20e-4 1.2e-7 1.00e-4 1.0e-7 8.00e-5 8.0e-8 6.00e-5 4.00e-5 q (coulombs cm-2) I (amps cm–2) 42 Charge rises exponential to a maximum when the switch is closed; voltage on the capacitor follows the charge q ( e–1) = 0.37 q At t = RC, q rises to within 1/e of its final value 6.0e-8 q ( 1– e–1) = 0.63 q 4.0e-8 2.00e-5 2.0e-8 0.00 0.0 6. Why do biological membranes act as tiny capacitors? 7. How does capacitance depend on membrane thickness and surface area? What is the dielectric constant? 8. What is the relationship between average solute velocity and solute flux? 9. What is the frictional coefficient and how does it relate to velocity and flux? 0.000 0.002 0.004 0.006 0.008 0.010 Time (s) Current is initially maximum and falls exponentially to zero as the capacitor charges 10. What is Kirchhoff’s Current Law? What is Kirchhoff’s Voltage Law? 11. What is meant by “resistive current”? “Capacitive current”? 12. Why does current stop in an RC circuit? 13. What is meant by the term “time constant”? What is the time constant in a simple RC circuit? Problem Set Physical Foundations: Pressure and Electrical Forces and Flows 1. Identify whether the following variables are intensive or extensive and explain your reasoning. A. Temperature B. Heat content C. Volume D. Density E. Mass F. Concentration G. Moles H. Pressure I. Area J. Flow K. Flux L. Viscosity 2. Normal systolic blood pressure is about 120 mmHg A. Convert this to atmospheres. B. Convert this to pascals 3. Normal diastolic blood pressure is about 80 mmHg A. Convert this to atmospheres B. Convert this to pascals 4. The density of whole blood is typically 1.055 g cm23. The density of Hg is 13.6 g cm23. A. Derive an equation to express the hydrostatic pressure of a column of blood as the height of Hg that would produce the same hydrostatic pressure as that column of blood B. Use the equation to determine the pressure of a 20-cm column of blood, expressed in mmHg C. If systolic blood pressure is 120 mmHg, what is the height of blood that this could support (if it was constant)? 5. For dialysis membranes, the Lp was determined to be 6.34 3 1027 cm min21 mmHg21. A cylindrical hole 1 cm in diameter was cut into two Lexan pieces that were then bolted together with the membrane between. Fluid entered one side through a pump and a pressure transducer was affixed. The flow was adjusted until the pressure (above atmospheric) was 20,000 pascals. The pressure on the opposite side was atmospheric (zero). Assume steady-state flow. What was the flow through the membrane? 1.1 6. The viscosity of water at 25 C is about 0.00089 Pa s. The inner diameter of a PE160 polyethylene tubing is 1.14 mm. Assume steadystate laminar flow. A. What pressure is necessary to get a flow of 5 mL min21 through a 20-cm length of this tubing? B. What is the velocity of the flow? C. What pressure is necessary to get a flow of 5 mL min21 through the same PE160 tubing if plasma is used, with viscosity of 0.002 Pa s? D. What pressure would be needed for the same flow of water through a 20-cm length of PE60 tubing with i.d. 5 0.76 mm? 7. After the end of a normal inspiration, the volume of air in the lungs is about 2.8 L. Normally quiet inspiration is driven by a pressure difference of about 2 mmHg. The air in the lungs is at 37 C and after normal expiration it is at atmospheric pressure. Quiet inspiration is driven by the expansion of the chest cavity by contraction of the diaphragm, which expands the air in the lungs. How much is the air expanded to produce an decrease of 2 mmHg in pressure? Use the ideal gas equation, PV 5 nRT, where P is the pressure, V is the volume, n is the number of moles, R 5 0.082 L atm mol21 K21 is the gas constant, and T is the absolute temperature. 8. A burette is a vertical, right circular cylinder that is open at the top and has a stopcock valve at the bottom to let fluid out. Assume that you have a burette that has an inner diameter of 1 cm and a height of 100 cm. A. Assume that you fill the burette with water to some height, h. What is the relation between h and the pressure at the base of the burette? B. What is the relation between volume of water in the burette and the pressure? C. How does this relation map onto the relation between charge and voltage on a capacitor? D. What is the hydraulic analogue of voltage? E. What is the hydraulic analogue of charge? 43 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00004-5 44 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 9. 10. 11. 12. 13. F. Suppose you open the stopcock fully. The fluid will drain out. Assume a constant diameter of the opening of the stopcock that provides a constant resistance, R. Derive an equation that describes the time course of draining the burette. G. Identify the time constant for the burette emptying. You are given a long vertical tube filled with fluid of viscosity η at sea level where the acceleration due to gravity is g 5 9.81 m s22. You have a steel ball or radius r and density ρsteel that you carefully drop into the fluid. Assume that the drag force on the ball obeys Stokes’ Law: the drag coefficient is 6 πrη, where Fdrag 5 26 πrη v. Remember that the steel ball is subjected to a buoyant force equal to the volume of the ball times the density of the fluid, ρfluid, times the acceleration of gravity. A. Derive an expression for the time of approach of the steel ball to terminal velocity. B. Derive an expression for the viscosity of the fluid as a function of the terminal velocity. (This is how you can determine the viscosity of a fluid.) The Poiseuille equation that relates flow in narrow tubes to the pressure difference is analogous to Ohm’s Law for current flow. A. What is the resistance to flow in terms of the parameters of the tube? B. What happens to resistance if the radius of the tube is halved? C. What happens to resistance if the radius of the tube is doubled? Suppose that you begin to exercise and as a consequence you begin to produce more heat. A. What do you suppose will happen to the body temperature if, when environmental temperature does not change, the rate of heat production increases. B. How is this change in body temperature described by the continuity equation applied to heat? C. How do you suppose the body will shed this excess heat production? You are out camping and it is very cold outside. You are losing heat faster than you can produce it. A. What happens to body temperature? B. How is this change in body temperature described by the continuity equation applied to heat? C. How do you suppose you can prevent hypothermia? Consider a capacitor with a capacitance of 10 μF. You connect it to a variable DC voltage source with a switch in the circuit and a resistor of 1000Ω in series with the capacitor. A. What is the relation between the steady-state voltage across the capacitor and the charge? 14. 15. 16. 17. 18. B. Before you close the switch, there is no charge on the capacitor. When you close the switch, current begins to flow from the voltage source at E volts. What is the relation between current and the voltage drop across the resistor in terms of the current and the resistance? C. What is the voltage drop across the capacitor in terms of charge and capacitance? D. Kirchoff’s voltage law says that the voltage drop around any loop must be zero. Write the equation for the voltage drop across the resistor, capacitor, and voltage source. E. Solve the equation in part D to derive the time course of charging of the capacitor. F. Solve the equation in part C to derive the time course of the current. G. Identify the time constant for the charging of the capacitor. An unmyelinated axon can be considered to be a long right circular cylinder. Consider that an axon is 10 cm long with a diameter of 1.0 μ (1 μ 5 1026 m). A. If the specific capacitance of the membrane is 1 μF cm22, what is the capacitance of the axon membrane? B. How much charge is separated by this membrane to give a potential of 70 mV? A muscle cell approximates a right circular cylinder 10 cm long and 70 μ in diameter. The specific capacitance of the membrane (the capacitance per unit area) is 1 μF cm22. A. What is the capacitance of the muscle membrane? B. How much charge is separated by this membrane to give a potential of 285 mV? A bubble is held at a radius of 250 μm. The transmural pressure difference is 2 mm Hg. What is the tension in the wall? The thickness of a single membrane is about 7 nm with a specific capacitance Cm 5 1 μF cm22. A myelin sheath consists of multiple membranes produced by coils of Schwann cell or oligodendroglia cells. Suppose a myelin sheath results from 100 membranes stacked on top of each other. A. What is the specific capacitance of the myelin? B. If the myelin sheath is a right circular cylinder 1 mm long with a radius of 3 m, what is its total capacitance? C. How much charge does it take to produce a voltage of 270 mV across this capacitor? D. If the myelin were just 1 membrane—i.e., not myelin—what would be its capacitance? How much charge would it take to produce a voltage of 270 mV across the single membrane? The transmural pressure difference across a small vein is 20 mmHg. The radius is 1 mm. What is the wall tension? Problem Set 19. The heat capacity of the human body is about 3500 joules kg21 C21. Suppose that a person weighs 75 kg. How much energy does it take to raise the body temperature from 98 C to 102 C? 20. The charge on the electron is 1.602 3 10219 coulombs. How much energy, in joules, is gained when an electron is accelerated across a potential difference of 1 v in a vacuum? This amount of energy is called the electron-volt. 21. Suppose that the average diameter of the aorta is 1.1 cm. The flow through the aorta is nearly the entirety of the cardiac output. A. If the cardiac output is 5 L/min, what is the average flow through the aorta? B. Suppose further that the cardiovascular system is nearly closed to fluid transfer. That is, that on a short-term basis the volume of the blood does not change. This means that ALL of the blood that leaves the heart goes through the arteries, and then capillaries, and then returns to the heart through the veins. Using the continuity equation, what is the flow through the aggregate capillaries? C. If the average diameter of the capillary is 4 3 1024 cm, and flow through the capillary is 0.1 cm s21, how many capillaries are there? 45 1.4 Chemical Foundations of Physiology I: Chemical Energy and Intermolecular Forces Learning Objectives G G G G G G G G G G G G G List the chemical elements that make up the organic part of the body List the major chemical elements found as electrolytes in the body List the chemical elements found in trace quantities and used as cofactors for enzymes Explain why single CC bonds rotate easily whereas double CC bonds do not Define and give examples of structural isomerism, geometric isomerism, and optical isomerism Write correct estimates of bond length and energy (within a factor of 2) for covalent bonds Define electronegativity Describe what is meant by a polar bond Distinguish between covalent and ionic bonds Define dipole moment and be able to calculate the energy of dipoledipole interactions Describe the hydrogen bond and recognize its typical energy Describe what is meant by London dispersion forces Draw the Lennard-Jones Potential and label the axes ATOMS CONTAIN DISTRIBUTED ELECTRICAL CHARGES In ordinary chemical reactions, atoms are the fundamental particles. The word atom derives from the Greek atomos, which means indivisible. The atoms themselves are composed of simpler subatomic particles, the neutrons, protons, and electrons. These particles are characterized by their rest mass and electrical charge, as shown in Table 1.4.1. Ernest Rutherford showed that all of the positive charges in an atom are concentrated within a very small volume, called the nucleus, and was awarded the 1908 Nobel Prize in Chemistry for the work. The nucleus has dimensions on the order of 10215 m! This requires some new thinking: if positive charges repel each other according to Coulomb’s law, how can they be concentrated in the nucleus? The answer is that there are other fundamental forces at work here, the strong 46 nuclear force and weak nuclear force, that have effects only over very short distances (,10214 m) and account for the stability of atomic nuclei. Each atom has a definite number of neutrons, protons, and electrons. The number of protons in the nucleus is called the atomic number, Z, and this number defines the chemical element that describes the atom. In a neutral atom, the number of electrons is equal to Z, and these orbit the nucleus. The behavior of the electrons defines the chemical reactivities of the elements, and the concentrated positive charge of the nucleus, in turn, determines the behavior of the surrounding electrons. ELECTRON ORBITALS HAVE SPECIFIC, QUANTIZED ENERGIES Although we refer to electrons as particles, in fact they have wave-like characteristics such as constructive and destructive interference. The structure of the atom cannot be explained using classical physics. Instead, it requires quantum mechanics. Quantum mechanics posits that the “orbit” of electrons around the nucleus is described by a wave function, which has been interpreted as being related to the probability of finding the electron in some volume. The wave function has quantum numbers in that it allow electron orbitals to have only specific energy levels, and transitions between them can occur only when the exchange of energy is exactly equal to the difference in the two energy levels. A set of quantum numbers uniquely describes the energy state of each individual electron in an atom. One quantum number describes the electron “shell,” a second describes the “orbital” within that shell, and a third describes the spin of the electron. The Pauli exclusion principle states that no two electrons can share the same set of quantum numbers within an atom. These orbitals are generally described as “clouds,” indicating the distributed nature of the orbital electrons. TABLE 1.4.1 Mass and Charge of Subatomic Particles Particle Rest Mass (g) 224 Electrical Charge (C) Neutron 1.6747 3 10 0 Proton 1.6726 3 10224 11.602176 3 10219 Electron 9.132 3 10228 21.602176 3 10219 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00005-7 Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces Group 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Period 1 1 H 2 3 Li 3 4 11 Na 19 K p Orbital s Orbital d Orbital 4 Be 12 Mg 17 Cl 33 As 34 Se 35 Br 35 Kr 47 Ag 48 Cd 49 In 50 Sn 51 Sb 52 Te 53 I 54 Xe 76 Os 77 Ir 78 Pt 79 Au 80 Hg 81 Tl 82 Pb 83 Bi 84 Po 85 At 86 Rn 65 Tb 66 Dy 67 Ho 68 Er 69 Tm 70 Yb 71 Lu 40 Zr 41 Nb 42 Mo 43 Tc 72 Hf 73 Ta 74 W 75 Re 88 Ra 16 S 46 Pd 39 Y 87 Fr 15 P 44 Ru 26 Fe 7 13 Al 10 Ne 18 Ar 31 Ga 25 Mn 56 Ba 9 F 30 Zn 24 Cr 55 Cs 8 O 29 Cu 23 V 6 7 N 28 Ni 22 Ti 38 Sr 6 C 27 Co 45 Rh 21 Sc 36 Rb 5 B 14 Si 32 Ge Transition elements 20 Ca 5 2 He 57 La 58 Ce 59 Pr 60 Nd 61 Pm 62 Sm 89 Ac 90 Th 91 Pa 92 U 93 Np 94 Pu 63 Eu 64 Gd FIGURE 1.4.1 Periodic table of the elements. Each element is distinguished by the atomic number, the number of protons in the nucleus of each atom. Each chemical element is symbolized by a one or two letter abbreviation, as indicated in the figure. Some elements have special biological significance. C, H, O, and N (carbon, hydrogen, oxygen, and nitrogen, respectively) form the basic organic structures of the body. Phosphorus and sulfur (P and S, respectively) are less common parts of the organic substance. Sodium (Na), potassium (K), magnesium (Mg), calcium (Ca), and chlorine (Cl) highlighted in gray make up the electrolytes of the body fluids and are necessary in large amounts. Other elements, particularly transition elements, bind to organic structures and enable their activities. These are required in trace amounts and include iron (Fe), manganese (Mn), cobalt (Co), nickel (Ni), copper, (Cu), zinc (Zn), selenium (Se), molybdenum (Mo), and iodine (I). HUMAN LIFE REQUIRES RELATIVELY FEW OF THE CHEMICAL ELEMENTS As noted above, each chemical element consists of atoms whose nuclei contain a definite number of protons and some number of neutrons, which typically is about the same as the number of protons, and an equal number of electrons distributed among the atomic orbitals. There are 94 naturally occurring elements, but relatively few of these are essential to human life, as illustrated in Figure 1.4.1. ATOMIC ORBITALS EXPLAIN THE PERIODICITY OF CHEMICAL REACTIVITIES There are eight main “shells,” referring to the principal quantum number, n 5 (1,2,3,4,5,6,7,8) that describes atomic orbitals. There are four major subshells: s, p, d, and f, whose names derive from spectroscopic descriptions of sharp, principal, diffuse, and fundamental. These orbitals are described by the azimuthal quantum number, l 5 (0,1,2,3) for (s,p,d,f), respectively. Each subshell has a structure and a capacity for electrons that is described by the magnetic quantum number, m, and the spin quantum number, s. The s subshell is spherically symmetrical and holds only 2 electrons; each set of p orbitals holds 6 electrons, the d orbitals hold 10, and the f orbitals hold 14. The sequential filling of these Principal quantum number n Azimuthal quantum number, l l=0 l=1 l=2 l=3 1 1s 2 2s 2p 3 3s 3p 3d 4 4s 4p 4d 4f 5 5s 5p 5d 5f 6 6s 6p 6d 7 7s 7p FIGURE 1.4.2 Order of filling of atomic orbitals. Electronic orbits are characterized by a principal quantum number that determines the main shell, an azimuthal quantum number that determines the subshell, a magnetic quantum number that determines the orbital, and the spin quantum number that determines the spin of the electron. There are four subshells: s, p, d, and f. These have 1, 3, 5, and 7 orbitals that each can hold up to two electrons of opposite spin. The order of filling with increasing number of electrons follows the blue diagonal arrows in the diagram: 1s fills first, followed by 2s and 2p; next is 3s followed by 3p and 4s, followed by 3d, 4p, and 5s; next is 4d, 5p, and 6s; then 4f, 5d, 6p, and 7s. orbitals accounts for the periodic chemical behavior of the elements with their atomic number. This order of filling is shown in Figure 1.4.2. Each subshell (s, p, d, f) is typically filled with the requisite number of electrons 47 48 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION before filling the remaining subshells. Each electron has a spin quantum number, s, that is represented as “up” or “down.” The orbitals in the subshells are typically filled singly with electrons of parallel spin before double occupancy begins. This is the so-called “bus seat rule,” analogous to the filling of a bus where double seats tend to fill with single individuals before double occupancy occurs. is entirely different from the two elements themselves. A classic example is water. Two volumes of hydrogen gas will combine with one volume of oxygen gas to produce water, which at the same temperature is a liquid and behaves altogether differently from either the hydrogen or oxygen gas. Such combinations of elements are called compounds, and the fundamental unit of them is the molecule. Molecules consist of atoms that are bonded together through the sharing of electrons in their outer atomic orbitals. The electrostatic shielding and energy involved in the orbital electrons overcome the repulsive forces between the positively charged nuclei. The resulting molecule is generally more stable than the starting materials. In order to break apart the molecule, energy must be supplied. This energy is called the bond energy, and its magnitude depends on the compound. Full orbitals are inherently stable, because they have low energy, and atoms having full orbitals are chemically unreactive. These correspond to the noble gases, helium (He), neon (Ne), argon (Ar), krypton (Kr), xenon (Xe), and radon (Rn). The electronic structure of some of these stable atoms is shown in Figure 1.4.3. All of the other elements can react with other atoms, in order to become more stable by attempting to fill their orbitals. They do this by sharing electrons, a process that constitutes chemical bonding. This sharing can be equal or very unequal, corresponding to the extremes of covalent bonding and ionic bonding. COMPOUNDS HAVE CHARACTERISTIC GEOMETRIES AND SURFACES Carbon has an atomic number of 6. Its electronic structure is 1s2 2s2 2p2: there are two electrons in the 1s orbital, two in the 2s orbital, and two in the 2p orbitals, as shown in Figure 1.4.4. Carbon can achieve the stable neon configuration of 1s2 2s2 2p6 by sharing electrons with four hydrogen atoms. The resulting ATOMS BIND TOGETHER IN DEFINITE PROPORTIONS TO FORM MOLECULES Two or more elements can combine to form a compound, and the resulting character of the compound Helium (2) 1s2 s Neon (10) 1s2 2s2 2p6 Argon (18) 1s2 2s2 2p6 3s23p6 Krypton (36) 1s2 2s2 2p63s23p64s23d10 4p6 s s s p p d p Xenon (54) 1s2 2s2 2p63s23p64s23d10 4p65s24d105p6 s d 1 1 1 1 1 2 2 2 2 2 3 3 3 3 3 4 4 4 4 4 p d 5 6 FIGURE 1.4.3 Electronic structure of the inert gases. These inert gases are chemically unreactive because their orbitals are already filled. Helium, with n 5 2 protons in its nucleus, fills the 1s orbital with 2 electrons of opposite spin. Spin is indicated in the drawing by an arrow pointed upward or downward. Neon (n 5 10) fills the 2s and 2p orbitals with a total of 8 electrons. Each orbital in the subshells carries at most two electrons. The order of filling of the orbitals corresponds to that shown in Figure 1.4.2. Carbon (6) 1s2 2s2 2p2 s p 1 2 3 4 FIGURE 1.4.4 Electronic structure of carbon. Carbon has 6 protons in its nucleus and 6 electrons that occupy the orbitals, 2 in the 1s orbital, 2 in the 2s orbital, and 2 more in the p orbitals. Only two of the three p orbitals are occupied by electrons. Carbon requires four more electrons to reach the stable configuration of Neon. Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces compound is methane, written as CH4 to convey the definite and fixed proportion of 1C for 4H atoms. All molecules in methane have this compositional stoichiometry. The close approach of the H nuclei and C nucleus alters the electronic structure of both the carbon and hydrogen. In molecular orbital theory, both the 2s and 2p orbitals of carbon participate in bonding by forming four hybrid molecular orbitals, termed sp3, meaning the hybrid of the 2s orbital with 3p orbitals (see Figure 1.4.5). The angle between the neighboring CH bonds is 109 280 . The shape-filling model of methane shows the edges of the carbon and hydrogen atoms as if they were hard spheres, but really the orbitals do not have such definite boundaries. The electron orbitals define these soft edges. All compounds are defined by the relative locations of their atomic nuclei and the three-dimensional distribution of their electronic charges. These make up a three-dimensional surface that can interact with other three-dimensional surfaces. The bedrock of all of chemistry and physiology is the interaction of these surfaces. SINGLE CC BONDS CAN FREELY ROTATE Carbon can also form bonds with other carbon atoms. Ethane has the compositional stoichiometry of C2H6 (see Figure 1.4.6). It is two methane molecules in which two CH bonds are replaced by a single CC bond. In the single CC bond, the sp3 hybrid orbitals overlap along their axis and form a circularly symmetric sigma bond. There is relatively free rotation around the axis of symmetry of this single bond, with three dips of about 12 kJ mol21 for each rotation when the H atom from one methyl group aligns with the space between the H atoms of the opposite methyl group. These ideas are shown schematically in Figure 1.4.6. 1s orbital from H sp3 from C H H 0.11 nm H C H C 109.5o H H H H 0.1 nm Chemical structure Ball-and-stick model Orbital picture Space-filling model FIGURE 1.4.5 Structure of methane. The compositional stoichiometry of methane is CH4—one carbon atom bonded to four hydrogen atoms. It arises from the sharing of the 1s electron of H with the 2s and 2p electrons of carbon. The bonding arises from overlap of the 1s H electron with electrons with hybrid C orbitals called sp3—formed from one s orbital and three p orbitals. Sigma bond forms by overlap of carbon sp3 orbitals Rotation about the C—C bond is relatively free, except for slight energy variations due to positions of H atoms in the two methyl groups H H H H C H C H C H H C H 109.3° H H H Chemical structure Orbital picture Ball-and-stick model Space-filling model FIGURE 1.4.6 Structure of ethane. Here 1C atom binds to another C and 3H atoms. The CC bond forms by overlap of the sp3 orbital along its axis to form a sigma bond that has circular symmetry. This bond can rotate about its axis, with some resulting configurations having just a little more stability than others. The most stable arrangement is shown, with the H atom of one methyl group aligned with the space between the H atoms in the opposing methyl group. 49 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION DOUBLE CC BONDS PROHIBIT FREE ROTATION of bonds that connect atoms together to make molecules, and each of these bonds has definite average bond lengths, bond angles, and bond energies. Table 1.4.2 summarizes approximate values for some of these bonds. These are approximate because atoms bound to other parts of the molecule can influence the molecular orbitals some distance away, thereby altering the angle, or length, or energy of any particular bond. Carbon can form double bonds by altering its molecular orbitals. Ethylene is ethane in which another pair of CH bonds converts to a second CC bond (see Figure 1.4.7). Instead of combining 2s2 with 2px1, 2py0, and 2pz1 to form the 4sp3 orbitals, it can arrange the electrons in a planar trigonal geometry by hybridizing the 2s2 with the 2px1 and 2py0 orbitals to form three sp2 orbitals with one electron each, and another pz1 orbital that can form a second bond, a pi bond, out of the plane of the sp2 orbitals. This bond resists twisting and a 90 twist breaks the overlap of the p orbitals, and hence breaks the second bond. Thus double bonds such as that in ethylene, shown in Figure 1.4.7, produce a somewhat rigid plane in any molecule in which they are found. BOND ENERGY IS EXPRESSED AS ENTHALPY CHANGES Earlier we wrote the conservation of energy theorem as ½1:3:6 where dE was the incremental change in the energy of a system, dq is defined as the heat absorbed by the system from its surroundings, and dw is defined as the work performed by the system on its surroundings. The total energy content of the system, E, does not depend on the path taken to get to its configuration. It is a state variable. If we conduct a change of state at constant pressure, then Eqn [1.3.6] can be written as CHEMICAL BONDS HAVE BOND ENERGIES, BOND LENGTHS, AND BOND ANGLES ½1:4:1 So far we have discussed CH bonds, CC single bonds, and CC double bonds. There are a host of other kinds Pi bond forms by overlap of carbon py orbitals H dE 5 dq 2 dW dE 5 dqp 2 P dV Rotation about the C—C double bond is restricted. H H C H C C C H H H H H H H pi bond C C H H C H H sigma bond d H C C C C Space-filling model H C H Ball-and-stick model Orbital picture Chemical structure H 50 H Carbon 1s orbitals Carbon sp2 orbitals H H Carbon pz orbitals H H Hydrogen 1s orbitals FIGURE 1.4.7 Structure of ethylene. The double CC bond is stronger than the single bond (it requires more energy to break) and locks the C atoms and all of the bonded groups into a single plane. This plane resists twisting because twist along the CC axis rotates the p orbitals away from overlap, breaking the pi bond. The space-filling model is not oriented the same way as the ball-and-stick model; the plane of the space-filling model is parallel to the plane of the paper. The lower part of the diagram shows the component orbitals. The 1s orbitals from carbon do not participate in molecular bonding. The sp2 orbitals bond to the H 1s orbitals through sigma bonds and form the CC sigma bond. The pz orbitals perpendicular to the plane of the sp2 orbitals form a second CC bond, a pi bond. Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces TABLE 1.4.2 Typical Bond Length and Bond Energy for Some Bonds Important in Physiology Bond Length (pm) Energy (kJ mol21) Bond Length (pm) Energy (kJ mol21) Bond Length (pm) Energy (kJ mol21) CC 154 346 CN 147 308 CO 143 360 CS 182 272 HC 109 411 HN 101 386 HO 96 459 HvS 134 363 PO 163 335 SS 205 226 CC 134 602 CO 120 799 PO 150 544 CC 120 835 Bond lengths are given in pm (510212 m) and bond energy is reported in kilojoule per mole (kJ mol21). Many tables of bond energies report them in kcal mol21. The calorie is defined as the amount of heat energy required to raise the temperature of 1 g of water from 14.5 C to 15.5 C. The calorie is readily converted to the joule using the conversion factor 1 cal 5 4.184 J. Calorie derives from the Latin calor, meaning to heat. if the only work is pressurevolume work. Note here that ½1:4:2 F dx 5 F A dx 5 P dV A clarifies that a pressure moving a volume has the same work units as a force moving through a distance. Integrating Eqn [1.4.1] between two states, we obtain Ðf Ðf ÐV 5 i dqp 2 Vif P dV i dE ½1:4:3 Ef 2 Ei 5 qp 2 PðVf 2 Vi Þ ðEf 1 PVf Þ 2 ðEi 1 PVi Þ 5 qp Hf 2 Hi 5 qp ΔH 5 qp Here we make the definition ½1:4:4 H 5 E 1 PV where H is the enthalpy. Since E, P, and V are all state variables, depending only on the state and not the path taken to that state, enthalpy is also a state variable. The bond energies are differences in the enthalpy of formation of the products and reactants in any chemical reaction. THE MULTIPLICITY OF CX BONDS PRODUCES ISOMERISM Some chemical compounds with identical compositional stoichiometry behave differently. Their different behaviors can be obvious and large, or quite subtle, depending on how the compounds differ. These compounds with identical composition differ in the way the atoms are arranged in the molecule. They are called isomers, meaning same weight. There are three major classes of isomers, structural, geometric, and optical, which are described below and shown schematically in Figures 1.4.81.4.10: G G G G Structural isomers differ in the connectivity of the atoms in the molecule. Stereoisomers have identical connectivity but the atoms are arranged differently in space. These consist of two classes: G Geometric isomers, involving a double CC bond that does not allow free rotation G Optical isomers, existing in two types. Enantiomers are mirror images of each other. They have identical physical characteristics such as melting point and density but are not superimposable. This requires an asymmetric carbon atom in which four nonidentical groups are bonded. They differ in their ability to rotate the plane of polarized light. Diasteriomers are optical isomers that are not enantiomers. They typically differ in the spatial distribution about one or more asymmetric carbons, while not being mirror images. UNEQUAL SHARING MAKES POLAR COVALENT BONDS The electrons that are shared in covalent bonds distribute themselves according to the charges on the nuclei within the molecule and the other electrons in the unshared orbitals that shield that charge from the electron. In almost all cases, the electrons are not shared equally but tend to spend more time near one or the other of the nuclei involved in the bond. The ability of an atom in a molecule to attract shared electrons is called its electronegativity. It generally increases going up the periodic table and going to the right, so that F has the highest electronegativity and Fr has the lowest. The electronegativity of some common elements is shown in Table 1.4.3 in arbitrary Pauling units scaled to F at 4.0. 51 Structural isomerism: same formula, different connectivity An aldehyde is a carbonyl group —O) bonded to an H and a C (C— Hydroxyl group is an —OH group bonded to another atom H H H C H A ketone is a carbonyl group —O) bonded to two C atoms (C— C O O C O P H C OH H O _ H _ O O C OH O C O P H O O _ _ Glyceraldehyde 3-phosphate Dihydroxyacetone phosphate Phosphate groups bind H2PO4 to C with high energy store FIGURE 1.4.8 An example of structural isomerism. Both glyceraldehyde 3-phosphate and dihydroxyacetone phosphate contain the same number of each type of atom. However, the connectivity of the atoms differs. Glyceraldehyde 3-phosphate contains an aldehyde group, which is defined as a C atom with a double bond to O and a single bond to H. Dihydroxyacetone phosphate contains a ketone group, which is a C atom with a double bond to O and the remaining bonds to C atoms. Conversion of the two chemicals requires breaking and reforming chemical bonds. Geometric isomerism: same connectivity but different spatial distribution Carboxyl group gives acid character to organic compounds HO HO O O C H C C FIGURE 1.4.9 An example of geometric isomerism. Both elaidic acid and oleic acid belong to a class of organic compounds called fatty acids, characterized by a carboxyl group at one end of an unbranched hydrocarbon chain. In these molecules, the chains are 18 carbons long. Both have one double bond beginning at the ninth carbon from the carboxyl end or the ninth carbon from the terminal methyl end. The spatial arrangement of the carbon chain and hydrogens at the double bond can be achieved in two ways: either cis or trans. In the cis arrangement, found in oleic acid, the two hydrogens are on the same side of the double bond, meaning that the hydrocarbon chain toward the carboxyl end and the hydrocarbon chain toward the methyl end are also on the same side, opposite to the hydrogens. This produces a kink in the hydrocarbon chain. In the trans arrangement, in the case of elaidic acid, the two H atoms on the doubly bonded C atoms are on opposite sides of the double bond. This has the effect of keeping the hydrocarbon chain straighter. Both compounds have the same composition but different physical properties due to this geometric isomerism. H C C C C H C H H H H H C H C H H C H H H C H C H H H H Oleic acid C H H C H C H C H C C H H H C H H C H H H H H H H C H H C C H H H H Elaidic acid H H H C H C C H Trans double bond has H groups on opposite sides of the double bond H H C H C H H H C C H H H Cis double bond has H H groups on the same side of the double bond; connectivity is identical to the trans configuration, but spatial difference is great C C H H H H C C H H H Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces Optical isomerism: same connectivity, not superimposable L-alanine D-alanine H O H H C O H C H H C C H C C N H H O H H N H H The amino group gives a basic character to the amino acid H The α carbon is asymmetric; it has four different groups and no plane of symmetry; it can form mirror image arrangements that cannot be superimposed WATER PROVIDES AN EXAMPLE OF A POLAR BOND TABLE 1.4.3 Electronegativity of Common Atoms Atoms H O C N O F Na Mg P Electronegativity 2.2 2.5 3.0 3.5 4.0 0.9 1.2 S FIGURE 1.4.10 An example of optical isomerism. Amino acids consist of a backbone of a carboxyl group bonded to a central C atom, called the α carbon, and an amino group bonded to the other side of the α carbon. Typically the α carbon has two more bonds, one to an H atom and the other to a variable group, called R for residue, whose composition determines the kind of amino acid. There are 20 different R groups and 20 different amino acids. The R group for alanine, shown here, is a methyl group. Because the α C atom is asymmetric—there is no plane of symmetry—the four groups can be arranged in two nonequivalent ways. Our bodies use only the L-amino acids. In the L-amino acids, starting from the carboxyl group and moving toward the amino group, the R group is to the left. This is very important, as D-amino acids in proteins would have their R groups pointing the wrong way. Cl 2.1 2.5 3.0 Source: L. Pauling, The Nature of the Chemical Bond. Cornell University Press, 1960. Atoms with similar electronegativities will share bonding electrons equally and will produce nonpolar bonds. Examples include the CC bond, CH bond, and HS bond. Bonds such as OH will be polar bonds. IONIC BONDS RESULT FROM ATOMS WITH HIGHLY UNEQUAL ELECTRONEGATIVITIES If the difference in electronegativity is too great, an ionic bond will form in which the strongly electronegative atom strips an electron from the weakly electronegative atom. Examples include NaCl, KCl, CaCl2, MgCl2, and a host of other physiologically relevant compounds. These are noted for their dissociation in water to form ions, Na1, Cl2, K1, Ca21, and Mg21, for example. These isolated ions are stabilized by their interaction with the polar water molecule, which presents a negative side towards the positive ions (cations) and a positive side towards the negative ions (anions). The bond angle defined by HOH is 104.5 , which is close to the tetrahedral angle. In this case, O forms 3 orbitals by hybridization of the 2s2 orbitals with the 2px and 2py orbitals. This leaves 2 O electrons in an sp2 orbital that are unshared and 2 electrons in a pz2 orbital that are also unshared. These form the lone electron pairs of water that participate in yet another kind of bonding, the hydrogen bond, discussed later. The electronegative O atom attracts the electrons away from the H nuclei, forming a partial separation of charge in the molecule itself. Thus the bond is said to be polar (see Figure 1.4.11). The estimated charge separation is about 20.67 on the O atom and about 10.33 on each of the H atoms. The dipole moment is defined as ½1:4:5 p 5 qd where p is a vector, the dipole moment, q, is the charge divided into equal q2 and q1, and d is the vector pointing from q2 to q1. The dipole moment of a single water molecule is 1.855 debye (1 debye 5 3.33564 3 10230 C m), but the dipole moment of water varies with the size of the water cluster, because nearby water molecules rearrange themselves in the presence of an electric field. Dipoles themselves produce an electric field and therefore interact with electric charges in its vicinity. These electrostatic interactions are part of the forces that govern the interaction of surfaces. In an electric field, the dipole experiences a torque given by 53 54 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION δ+ H 0.096 nm H O O δ– 104.5o H δ+ H Chemical structure Ball-and-stick model Orbital picture Space-filling model FIGURE 1.4.11 Structure of water. The HOH angle is 104.5 and the bond length is about 96 pm. The O atom is far more electronegative than the H atoms, thereby causing an asymmetric redistribution of the electrical charges away from the H atoms and surrounding the O atom. These partial separations of charges produce a polar compound and a net dipole moment. δ– δ+ δ– δ+ O H δ+ δ+ H δ– H O H δ+ FIGURE 1.4.12 The hydrogen bond in water. The polar OH bond involves a separation of charge. The partial positive charge on the H atom in water is attracted to the lone electron pairs on the opposite side of the molecule on adjacent molecules. This forms a weak bond that is easy to form and easy to break. In many situations, the number of hydrogen bonds significantly stabilizes large structures. In water, there are many hydrogen bonds because each water molecule has the potential of participating in four of them: two because of the H atoms and two more because of the two lone electron pairs. ½1:4:6 τ5p3E where τ is the torque, p is the dipole moment, and E is the electric field strength. INTERMOLECULAR FORCES ARISE FROM ELECTROSTATIC INTERACTIONS Chemical bonds join atoms together to form molecules. Molecules can also be attracted to each other through a variety of intermolecular forces that include: G G G Hydrogen bonding Dipoledipole interactions London dispersion forces. HYDROGEN BONDING OCCURS BETWEEN TWO ELECTRONEGATIVE CENTERS The hydrogen bond involves the sharing of the positive charge of hydrogen between two electronegative centers such as oxygen and nitrogen. It requires proximity and proper orientation of the two electronegative centers. In the case of water, the bond is strongest when the OH bond on one water molecule aligns with the lone electron pair orbital of the adjacent water molecule, as shown in Figure 1.4.12. The hydrogen bond requires only 840 kJ mol21 to break, compared to much higher values for covalent bonds such as CC (346 kJ mol21) or CH (411 kJ mol21). This low bond energy makes it useful, because it means that the bond can form and break under the influence of normal thermal agitation. At the same time, a large number of hydrogen bonds can stabilize structures such as proteins and DNA. DIPOLEDIPOLE INTERACTIONS ARE EFFECTIVE ONLY OVER SHORT DISTANCES By virtue of their spatial separation of charges, dipoles produce an electric field surrounding them whose magnitude is given by ½1:4:7 Uðr; θÞ 5 p cos θ 4πε0 κr 2 An elementary proof of this result is given in Appendix 1.4.A1. The angle θ is defined as the angle between the point at which potential is given and the midpoint between the two separated charges within the dipole. Note that this is a potential, not the force. In Coulomb’s law, the potential varied with 1/r (see Eqn [1.3.5]) but here it varies with 1/r2. Replacing one ion with a dipole causes the interaction to be shorter range. At short range, Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces a nearby charge “sees” both charges; at longer range the charge “sees” two opposing dipole charges that tend to neutralize each other; the interaction becomes weaker at longer distances. Because they both produce an electric field, pairs of dipoles interact with each other. The energy of interaction is given by 2 pA pB 2 1 ½1:4:8 UðrÞ 5 2 3kT 4πε0 κ r 6 where k is a new physical constant, Boltzmann’s constant, which is the gas constant per molecule. Its value is 1.38 3 10223 J K21; T is the temperature in K, pA and pB are the dipole moments of the two dipoles, ε0 is the electrical permittivity of space, given earlier as 8.85 3 10212 C2 N21 m22, and r is the separation of the centers of the dipoles. LONDON DISPERSION FORCES INVOLVE INDUCED DIPOLES Water has a permanent dipole moment. Symmetrical compounds such as methane and H2 have no permanent dipole moment, but these can be induced to form a dipole by the presence of an externally applied electric field. A polarizable atom redistributes its internal charge in response to an electric field to form a dipole moment aligned with the applied field. For small electric field strength, the induced dipole is approximately proportional to the applied field: ½1:4:9 pind 5 αE where pind is the induced dipole moment, E is the electric field, and α is the polarizability. The SI unit for α is C m2 V21, but it is often converted to units of volume, cm3, by multiplying by 1/4πε0 3 1026, where ε0 is the electrical permittivity of space. This effect, shown in Figure 1.4.13, results in the attraction of a neutral molecule to a charged molecule. The totality of forces between atoms or molecules due to dipoledipole, dipole-induced dipole, and instantaneously induced dipoles (London dispersion force) is called the van der Waals force. It excludes the interaction due to covalent bonds or electrostatic interactions. CLOSE APPROACH OF MOLECULES RESULTS IN A REPULSIVE FORCE THAT COMBINES WITH THE VAN DER WAALS FORCES IN THE LENNARDJONES POTENTIAL Imagine two atoms or molecules separated by a large distance. Because of the large distance, their interaction is very small—there is little force between them. Even if they have very little dipole moment, as they approach one another they will experience attractive forces due to London dispersion forces, and these are attractive. As the distance between their atomic nuclei shortens, they begin to experience repulsive forces due to the interpenetration of their atomic or molecular orbitals. This repulsive force varies quite steeply with separation. The overall potential energy for the interaction of nonbonding particles has been mathematically approximated by the LennardJones potential. It is given as r 6 σ 12 σ6 rm 12 m 2 22 5ε UðrÞ 5 4ε r r r r ½1:4:10 where U(r) is the potential energy of the interaction, ε is depth of the potential well (a measure of the strength of the attractive forces), and σ is the distance at which the intermolecular potential is zero; rm is the distance of separation when the potential is 2 ε (see Figure 1.4.14). E + Even molecules that do not have a permanent dipole moment can transiently produce dipole moments that result in their attraction. We imagine that electrons orbit their nucleus in an “electron cloud” but this picture is an average distribution. At any instant the electrons are separated from their nucleus, producing a transient dipole. When nearby atoms synchronize the distribution of electrons in their clouds, they can produce attractive forces first described by F. London in 1937 and called London dispersion forces. r – – – – – + + + + + q– q+ Δr FIGURE 1.4.13 Electrical polarizability. The presence of a charge establishes a local electrical field. Electrons within nearby molecules can respond to this field by redistribution of charge, causing an induced dipole. Because of the separation of charge, the induced dipole experiences unequal forces from the fixed charge. The result is a net movement of the induced dipole. ATOMS WITHIN MOLECULES WIGGLE AND JIGGLE, AND BONDS STRETCH AND BEND The bond lengths, angles, and energies listed in Table 1.4.2 are averages. Two atoms involved in a bond actually oscillate back and forth around the average bond length. In addition, the angles defined by, say, HCH are not fixed, but the three atoms oscillate around the average bond angle. The molecules also translate through the solution or gas, and rotate. Some of these motions affect others, as rapid rotation about an axis perpendicular to a bond tends to 55 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION stretch the bond. All of these modes of action within molecules store kinetic and potential energy on a submolecular scale. When a moving body stops because of friction, the kinetic energy does not disappear; it is converted from the macromolecular to the molecular scale—the kinetic energy appears as heat and the temperature of the surfaces increases. This conversion is a one-way street: we can never fully recover that molecular energy back into macromolecular action. The temperature is, in one sense, a measure of this dynamic motion. These various modes of thermal motion are shown in Figure 1.4.15. Intermolecular potential U(r) 56 rm 0 σ r ε F >0 F=0 Repulsive force F<0 Attractive force FIGURE 1.4.14 The LennardJones potential. At far separation, there is little force and no potential. Because of the inverse 6th power dependence on the separation, this force becomes larger as particles approach one another, reaching a minimum at U(r) 5 ε and r 5 rm. The separation at which U(r) 5 0 is r 5 σ. The equilibrium for the particles occurs at the minimum potential. SUMMARY Atoms are the fundamental units of the elements and are defined by the number of protons in the atomic nucleus. Only relatively few of the chemical elements have active roles in the body. The electrons orbit the nucleus in electron orbitals that have definite energy levels. These electron orbitals describe the distributed nature of the electrons within their atoms. Electrons populate the orbitals in a defined sequence that gives rise to the periodic behavior of the chemical elements within the periodic table. The chemical elements can combine to form compounds in fixed ratios to each other, often by sharing the outer orbital electrons. These bonds form by combination of their atomic orbitals or by the formation of molecular orbitals in which the electrons are distributed among the nuclei that make up the molecule. Most organic compounds form covalent bonds in which the electrons are shared approximately equally among the nuclei. These bonds have energies in the range 200500 kJ mol21 and have bond lengths on the order of 0.10.2 nm. The bond angles are determined by the kind of bonds that form. Many of the biochemicals within the human body are organic compounds that have CC, CC, CN, COH, CO, and CH bonds. The single bonds typically freely rotate about the axis connecting the two nuclei. Double bonds do not permit free rotation because the rotation would break the second bond and this requires energy. The rigid form of the CC bond gives rise to geometric isomerism. The different arrangement of the same nuclei around the CC bond can cause large differences in the overall shape of molecules. Isomerism also arises from the different spatial arrangement of chemical groups around an asymmetric C atom. Such isomers can be mirror images of each other (enantiomers) or not. All of the amino acids (except glycine) exist in enantiomeric forms, but only the L-type of amino acids are used to make proteins. Rotation about the y axis Bending of the bond Stretching of the bond Rotation about the x axis Rotation about the z axis Three translational degrees of freedom FIGURE 1.4.15 Degrees of freedom of motion in water. The HOH bond can bend, changing its angle; the OH bond can oscillate, stretch, and compress its bond length; the entire molecule can rotate around several independent axes; the molecule can translate in three independent directions. Each of these modes of movement is independent and each carries some kinetic energy. The energy distributed amongst the various modes of motion increases with increase in temperature. Che mical Foundations of Physiology I: Chemical E nergy a nd Inter mol ecular Forces Polar bonds involve unequal sharing of bonding electrons between the two nuclei involved in the bond. A good example is the OH bond in which the oxygen atom is more electronegative. Electronegativity refers to the ability of a nucleus to attract shared electrons. Because O is more electronegative than H, the resulting OH bond is polar, with more negative charge around the O atom than the H atom. This partial separation of charge produces a dipole which is described by its dipole moment, equal to the charge times its separation directed from the 2 to the 1 charge. The dipole moment produces an electric field that can interact with other nearby atoms. The OH group is very important because it forms hydrogen bonds with other electronegative atoms. This low energy bond is easy to form and easy to break and can stabilize biological structures. Ionic bonds occur when two atoms or chemical groups differ greatly in their electronegativities. The more electronegative atom or group effectively “steals” an electron from the less electronegative atom. When dissolved in water, these types of compounds typically dissociate into a cation and anion. Molecules can interact with each other through ionion interactions, described by Coulomb’s law, hydrogen bonding, dipoledipole interactions, and London dispersion forces. The London dispersion forces arise from transient dipoles in atoms that induce transient dipoles in nearby atoms, producing an attractive force. The LennardJones potential describes the overall interaction between nonbonded particles. Atoms and molecules form surfaces that interact with other surfaces. Almost all of physiology is about how these surfaces interact to produce catalysis or tight binding or loose binding or specificity of binding. REVIEW QUESTIONS 1. What is the fundamental unit of an element? What is the fundamental unit of a chemical compound? 2. Why do atoms form compounds in definite and fixed proportions? 3. What forms the surfaces of atoms and molecules? 4. Why are the CH bonds in methane arranged to point toward the vertices of a tetrahedron? 5. What is the tetrahedral angle? 6. What are typical energies for CH and CC bonds? 7. What are typical bond distances for CH and CC bonds? 8. What is meant by structural isomerism? 9. What is meant by geometric isomerism? 10. What is optical isomerism? 11. Why does water form polar covalent bonds? 12. What is hydrogen bonding? How much energy is in a hydrogen bond? 13. What is an electric dipole? 14. What are dipoledipole interactions? 15. What are London dispersion forces? APPENDIX 1.4.A1 DIPOLE MOMENT A DIPOLE CONSISTS OF TWO EQUAL CHARGES SEPARATED BY A DISTANCE, D, AND IS DESCRIBED BY ITS ELECTRIC DIPOLE MOMENT An electric dipole consists of two equal charges, q1 and q2, separated by a distance d, as shown in Figure 1.4. A1.1. These are typically molecules whose separation distance is small compared to the distance at which electrical effects are noted. As we shall see, cardiomyocytes can also act as electric dipoles. The electric dipole moment is defined as ½1:4:A1:1 p 5 q1 d where d is a vector pointing from q2 to q1, as shown in Figure 1.4.A1.1. The net force on a unit positive test charge at any point surrounding the dipole can be determined by the vectorial sum of the component forces from q1 and q2, as shown in Figure 1.4.A1.1. We can determine a set of points surrounding the dipole that has the same magnitude of force, but in varying directions. This set forms a curve. The family of lines for a set of force magnitudes, shown in Figure 1.4.A1.2, represents the electric field surrounding the dipole. Moving a positive unit charge from a large distance away (N) to any point within the electric field entails p = q+ d d q+ q– θ r+ r––r+ r– F– F– Ftot qtest F+ Ftot qtest F+ FIGURE 1.4.A1.1 Origin of the electrical forces near an electric dipole determined at two different locations. The electric dipole consists of two equal charges (q1 and q2) separated by a distance, d. Charges placed nearby experience a force as a result of the electric dipole. The net force is the vector sum of the forces exerted by the two charges, as shown for two different positions. Because the force declines as 1/r2, the direction of the force changes with distance and angle, θ, made between the line joining the center of the dipole and the test charge, and the electric dipole moment. 57 58 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION V=0 in Figure 1.4.A1.2. These lines intersect the lines of electric force at right angles. The electric potential surrounding a dipole is just the sum of the potentials associated with each charge. Thus we write: - + + ½1:4:A1:2 FIGURE 1.4.A1.2 Electric field and electrical potential surrounding a dipole. The electric field lines are shown in dashed lines with arrows. They represent the lines of equal force that would be felt by a positive unit test charge as it approaches the dipole. The electric field is a vector which, when magnified by the size of a charge, gives the magnitude and direction of the force. The electric potential contour lines are shown in solid. The negative gradient of these contour lines are the electric field lines. Thus the electric field is the steepest slope down the potential surface. The lines of equal potential intersect the electric field lines at right angles, much like the steepest descent off a hill intersects its altitude contour lines at right angles. expending energy that is stored as potential energy, which is identified as the electrical potential. The potential at any point is the scalar sum of the potential energy associated with each charge. That is, the potential at any point is just the sum of the potential energy of q1 and that of q2. Contour lines of equal potential are shown Vtotal 5 V1 1 V2 q1 q1 5 1 4πε0 r1 4πε0 r2 q1 1 1 5 2 r2 4πε0 r1 q1 r2 2 r1 5 4πε0 r1 r2 q1 d cos θ p cos θ 5 4πε0 r 2 4πε0 r 2 where the quantity q1d appears. We have identified this as the electric dipole moment, p 5 q1d. Here d cos θ enters the equation as an approximation of r2r1, as suggested by the geometry of Figure 1.4.A1.1. This assumption is generally made when rcd. Thus the voltage at any point is inversely proportional to the square of the distance and varies with the relative position of the point with respect to p, the electric dipole moment. The point of these calculations is to show that the electric dipole creates an electric field and a potential field that can be measured some distance away. The value of the potential depends on the distance and the relative position compared to the dipole. Chemical Foundations of Physiology II: Concentration and Kinetics Learning Objectives G G G G G G G G G G G G Be able to calculate the molar concentration given the mass of solute, formula weight, and volume of solution Be able to calculate the number of moles of solute in an aliquot of solution given the volume and concentration Be able to determine the gram molecular weight of simple compounds Know the prefix notations for scientific notation Be able to calculate dilutions of stock solutions to form working solutions Be able to calculate the volume of distribution using Fick’s dilution principle Define rate constant Be able to derive a first-order rate equation and calculate its half-life Define equilibrium constant Describe how an enzyme can change the rate of a reaction without changing its equilibrium Write the MichaelisMenten equation and draw a saturation plot. Identify Km and Vmax Draw a double-reciprocal plot and identify Km and Vmax on this plot mixtures of isotopes that differ slightly in their atomic mass. These are atoms that possess the characteristic number of protons (the Z number) but differ in their numbers of neutrons. Each molecule has a definite mass that depends on the atoms that make it up. If we add up the atomic masses of the constituent atoms, we get the molecular weight in daltons. If we express this molecular weight in grams, we will have the gram molecular weight (just scratch out “daltons” and replace it with “grams”). A pile of molecules whose mass in grams is the gram molecular weight is called a mole. Such a pile will consist of not one molecule but very many of them, and the number will be the number of daltons in a gram. To see this, suppose we take a substance with a molecular weight of Z daltons. How many of these molecules will we have to pile up in order to make up a pile with a mass of Z grams? Let N be the number of molecules in the pile of Z grams. The mass of one molecule is Z daltons. The mass of N of them is just N 3 Z daltons. This mass will be the gram molecular weight. Thus, we have Z daltons 3 N 5 Z grams N5 AVOGADRO’S NUMBER COUNTS THE PARTICLES IN A MOLE As described in Chapter 1.4, each chemical species is composed of a definite number of atoms of each element. This is Dalton’s law of fixed proportions, which states that a molecule contains integral numbers of each kind of atom. Under ordinary chemical reactions, these atoms cannot be converted into each other. Each of these atoms contributes a tiny but definite mass to the molecule. The atomic mass unit is defined as 1/12 of the mass of a carbon-12 atom, the one containing 6 protons and 6 neutrons in its nucleus. This unit of mass is also called the dalton, Da. The atomic weight of a carbon-12 atom is defined as being 12 daltons. The atomic weight of most elements, including carbon, is not exactly integral because most elements consist of 1.5 Z grams 1g 5 Z daltons 1 Da The actual mass of 1 Da is 1.66 3 10224 g. Thus, the number of molecules in a gram molecular weight is N5 1g 5 6:02 3 1023 1:66 3 10224 g This is, of course, Avogadro’s number. It is the number of molecules per mole. A mole is a pile of molecules of a single substance whose mass in grams is equal to its gram molecular weight and which contains Avogadro’s number of molecules. We can turn this definition around and define the mole as Avogadro’s number of particles. This definition is completely general. We did not specify how big the molecule is in the above calculation. No matter the size of the compound, one mole contains Avogadro’s number of molecules. 59 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00006-9 60 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION CONCENTRATION IS THE AMOUNT PER UNIT VOLUME In Chapter 1.2, we defined extensive and intensive variables as those that depend on the extent of the system and those whose value is independent of the extent of the system, respectively. The amount of a solute, and the volume in which it is distributed, are both extensive variables. If you have twice the volume of solution, you have twice the amount of solute dissolved in it. The concentration, on the other hand, is an intensive variable which is given as the ratio of two extensive variables: amount Concentration 5 volume ½1:5:1 m C5 V where C is the concentration, m is the amount of solute, and V is the volume in which that amount of solute is dissolved. SCIENTIFIC PREFIXES INDICATE ORDER OF MAGNITUDE Often concentrations of physiologically relevant materials in physiological fluids are quite small and must be expressed in scientific notation, as in the example of hemoglobin in Example 1.5.3. It is useful to know the established prefixes for units of volume, mass, or length. The standard units are the liter, L, the gram, g, and the meter, m, respectively. The standard prefixes are shown in Table 1.5.1. EXAMPLE 1.5.1 Calculating Molar Concentration Suppose we dissolve 18 g of NaCl in water in a 2-L volumetric flask and then add water to the mark so that the solution volume is 2 L. What is its concentration? We can report this concentration in mass per liter as 9 g L21. For many solutes, this is an acceptable way of expressing the concentration. For example, the concentration of hemoglobin in whole blood is 15 g dL21. Here dL means “deciliter” and is one-tenth of a liter or 0.1 L. Since 1 L 5 1000 mL, 1 dL 5 1000 mL/10 5 100 mL. Thus, a hemoglobin concentration of 15 g dL21 means that there is 15 g of hemoglobin in every 100 mL of whole blood. The concentration of many solutes is not reported in these units, however, but in units of moles per liter, or molar. We can convert from mass to moles by dividing by the gram molecular weight. In the case of NaCl given above, the atomic weight of Na is 22.99 daltons and the atomic weight of chlorine is 34.45 daltons, giving a molecular weight for NaCl of 22.99 1 35.45 5 58.44 daltons; its gram molecular weight is 58.44 g. Thus 58.44 g of NaCl constitutes 1 mole of NaCl. The number of moles of NaCl in 18 g is calculated as X moles 5 18 g NaCl 5 0:308 mol 58:44 g NaCl mol21 Since this amount was dissolved in 2 L of solution, its concentration 5 amount/volume. Concentration 5 0:308 mol 5 0:154 M 2L where M designates molarity or moles per liter. EXAMPLE 1.5.2 How to Make Up a Solution You need to make up 500 mL of a solution containing 0.3 M urea. How much urea do you need? Here we rearrange the equation C 5 m/V to read m 5 C 3 V. In this case, C 5 0.3 M and V 5 0.5 L. The number of moles of urea is thus m 5 0:3 M 3 0:5 L 5 0:15 mol We need to convert this into grams so that we can accurately weigh out the required amount of urea on a good balance. To do this, we need the gram molecular weight of urea. We can find this out several ways. We can look it up in a CRC handbook or similar source. Another way is to write out the formula and add up all of the atomic weights of all the atoms in the molecule times their compositional stoichiometry. This could be tedious for a large molecule. We could look on the bottle, because most chemical companies publish the formula weight on the bottle. This formula weight may be different from the gram molecular weight because the chemical might have waters of hydration with it or ions to counterbalance charges on the chemical. ATP, for example, is usually sold as Na2ATP 3H2O and its formula weight is not the gram molecular weight of ATP alone. Looking on the bottle seems like a winner, as we must go find the bottle to weigh out the urea. We find that the formula weight for urea is 60.08 g mol21. The mass of urea is found by multiplying the moles by the gram molecular weight: m 5 0:15 mol 3 60:08 g mol21 5 9:012 g This is the amount you need to weigh on an accurate balance. Chemical Foundations of Physiology II: Concentration and Kinetics EXAMPLE 1.5.3 Blood Concentration of Hemoglobin The blood content of hemoglobin is 15 g dL21. The molecular weight of hemoglobin is 66,500 daltons. What is its average molar concentration in the blood? If the molecular weight is 66,500 daltons, then 1 mole has a mass of 66,500 g. The question asks for the average concentration because hemoglobin is not uniformly distributed within the blood, but is contained within red blood cells only. Therefore, its concentration in the red blood cells exceeds its average concentration in the blood. The concentration in blood is given by ½Hb 5 15 g dL21 3 1 dL=0:1 L 5 0:00266 M 5 2:66 3 1023 M 66; 500 g mol21 These examples should enable you to calculate: G G G the number of moles in a given volume of solution of given concentration; the mass of material in a given volume of solution of given concentration; the concentration of solution containing a known mass of material in a given volume. EXAMPLE 1.5.4 Make a Dilution of a Stock Solution Suppose we have a stock solution of 0.1 M MgCl2 and we want to make up 50 mL of solution with a final concentration of 5 mM MgCl2. How much of the stock solution should we add to a 50mL volumetric flask? We use Eqn [1.5.3] directly here: 0:1 M 3 X mL 5 0:005 M 3 50 mL Solving for X, we find X 5 2.5 mL TABLE 1.5.1 Prefixes Used with Powers of Ten, in 103 Ratios Atto 10218 Femto 10215 Pico 10212 Nano 1029 Micro 1026 Milli 1023 100 Kilo 103 6 Mega 10 Giga 10 9 Thus we can write: Tera 12 10 DILUTION OF SOLUTIONS IS CALCULATED USING CONSERVATION OF SOLUTE 1 picoliter (1 pL) 5 10212 L 1 picogram (1 pg) 5 10212 g 1 picometer (1 pm) 5 10212 m 1 nanoliter (1 nL) 5 1029 L 1 nanogram (1 ng) 5 1029 g 1 nanometer (1 nm) 5 1029 m 1 microliter (1 μL) 5 1026 L 1 microgram (1 μg) 5 1026 g 1 micrometer (1 μm) 5 1026 m 1 milliliter (1 mL) 5 1023 L 1 milligram (1 mg) 5 1023 g 1 millimeter (1 mm) 5 1023 m Often it is necessary or easier to prepare solutions from more concentrated stock solutions by removing an aliquot (a fraction of the solution) of the more concentrated solution and placing it in a volumetric flask. We then add solvent to bring the volume up to the final solution volume, as shown in Figure 1.5.1. 1 liter (1 L) 5 100 L 1 gram (1 g) 5 100 g 1 meter (1 m) 5 100 m 1 kilogram (1 kg) 5 103 g 1 kilometer (1 km) 5 103 m Let V1 be the volume of the aliquot of the more highly concentrated solution, with concentration C1. The amount of solute in this aliquot is given by Eqn [1.5.2]: ½1:5:2 The prefixes mega-, giga-, and tera- are not typically used for units of volume, mass, or distance, but often find use with other units such as hertz or watts. There are additional prefixes in the SI. These include: centi 1022 deci 1021 deca 101 hecto 102 m 5 C 1 3 V1 This amount is also the amount in the final solution with volume V2 and final concentration C2. Since the amount in the aliquot is still in the final solution, we can write ½1:5:3 C1 V1 5 C2 V2 61 62 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 1 Remove aliquot of known 2 Transfer aliquot to empty 3 Add water to volume; volumetric flask; m1 = V1 C1 volume, V1; m1 = V1 C1 m1 = V2 C2 = V1 C1 Stock solution Flask 1 Flask 2 Flask 1 Flask 2 Flask 1 Flask 2 Flask 1 Flask 2 FIGURE 1.5.1 Dilution of a stock solution. Flask 1 contains a stock solution of concentration C1. We remove a known volume, V1, with a calibrated pipette and place it in an empty volumetric flask. We then add water, which adds solvent but no solute, to bring the volume up to the final volume. The amount of solute in the second flask is the same amount that we added in the aliquot, V1, but now it is distributed in the volume V2. CALCULATION OF FLUID VOLUMES BY THE FICK DILUTION PRINCIPLE The equation describing dilution of stock solutions can be used to determine the volume of distribution of materials in the body. Suppose we inject a known amount of a substance, Y, into a person and this substance Y remains in the plasma because it cannot get out of the vascular system and it cannot get into the cells suspended in the blood. We wait a few minutes for Y to become evenly distributed, and then take a sample of plasma and measure Y’s concentration. Then we can calculate the volume of the plasma as amount Volume 5 concentration ½1:5:4 m V5 C This is just a variant of Eqn [1.5.2] in which we solve for V instead of C or m. In our particular case in this example, we write V 5 mY =½Y where mY is the amount of substance Y and [Y] is its concentration. CHEMICAL REACTIONS HAVE FORWARD AND REVERSE RATE CONSTANTS Suppose that we observe a simple chemical reaction that can be described as ½1:5:5 A 1 B"C where A, B, and C denote different chemicals and the arrows indicate that the reaction proceeds in the direction of the arrow. The reaction actually consists of two separate reactions: A 1 B,C A 1 B&C ½1:5:6 The top reaction is the forward reaction and the bottom reaction is the reverse reaction. Both are characterized by a rate constant. The rate constant gives the rate of the reaction when the rate constant is multiplied by the concentration of reactants: Jf 5 kf ½A½B Jr 5 kr ½C ½1:5:7 EXAMPLE 1.5.5 Estimation of Plasma Volume Suppose that we inject a 50-kg person with 2 mL of a solution of Evans’ Blue Dye, at 5 mg mL21. Evans’ Blue Dye is restricted to the plasma because it tightly binds to a plasma protein, albumin, that ordinarily does not escape from the circulation and which only slowly is removed or added to the plasma. We wait 10 min and then obtain a sample of blood and measure the concentration of Evans’ Blue Dye in the plasma and find that it is 0.4 mg dL21. What is the person’s plasma volume? The amount of Evans’ Blue Dye that was injected is calculated as minj 5 Cinj Vinj 5 5 mg mL21 3 2 mL 5 10 mg The concentration of Evans’ Blue Dye after mixing was 0.4 mg dL21. We calculate its volume of distribution as V5 10 mg 5 25 dL 5 2:5 L 0:4 mg dL21 Chemical Foundations of Physiology II: Concentration and Kinetics where Jf is the forward reaction rate, Jr is the reverse reaction rate, kf is the forward reaction rate constant, and kr is the reverse rate constant, and [A], [B], and [C] are the concentrations of the indicated reactants. Although this form of the reaction rate is true for elementary reactions, more complicated reactions could have a form that appears to have little to do with the overall reaction as written—its stoichiometric relation or accounting of the number of each kind of molecule that participates in the reaction. This is because complicated reactions occur with intermediary steps that may involve chemicals that do not appear in the overall balanced reaction. Nearly all enzymatic reactions, for example, do not obey Eqn [1.5.7] because the reaction rate is determined largely by the concentration of enzyme. For the moment, we will consider only elementary reactions that obey Eqn [1.5.7]. The rate of a reaction is the number of completed reactions that take place per unit time. You should recognize this as an extensive property. If we had twice the volume of a solution, with the same concentrations of reactants and products, we would have twice the number of reactions taking place per unit time. So we often convert reaction rates into intensive variables by dividing by the volume. The forward rate is thus the number of completed reactions per unit time per unit volume. We express these numbers of reactions in terms of moles, which are related by Avogadro’s number to a real number of completed reactions. Thus the forward rate constant has units of M21 s21. By similar reasoning, the reverse rate constant in Eqn [1.5.7] has the units of s21. The rate of change of reactants A and C can be given as ½1:5:8 d½C 5 2 kf ½ A½B 2 kr ½C dt d½A 5 0 5 2kf ½A½B 1 kr ½C dt FIRST-ORDER RATE EQUATIONS SHOW EXPONENTIAL DECAY Some kinds of chemical reactions, such as decomposition reactions, obey the equation ½1:5:11 The rate constants, kf and kr, are characteristic of the reaction path involved and the experimental conditions such as ionic strength and temperature of the reactants. They typically do not vary with the concentrations of A, B, and C. Thus, at equilibrium, the middle expression in Eqn [1.5.9] is true for any set of [A], [B], and [C]. We rearrange this to get: A-B Often this reaction is strongly directed to the right, meaning that kf .. kr. The reaction rate is approximated as ½1:5:12 J 5 2 kf ½A Here we imagine that initially [B] 5 0 and there is no reverse reaction. Some reactions occur with such completeness that the reverse reaction is negligible. We can rewrite Eqn [1.5.12] as d½A 5 2 kf ½ A dt This is called a first-order rate equation because the rate of reaction is proportional to the first order or the concentration of reactant—it is proportional to the first power of its concentration. We can separate variables and integrate this equation, as follows: ½1:5:13 ðt ½1:5:14 0 ðt d½A 5 2k dt ½A 0 In½A 2 In½A0 5 2 kt ½A 5 ½A0 e2kt The last equation describes the concentration of A with time—it decays exponentially. This equation is described as a first-order exponential decay. Many reactions and processes are described by this type of analysis, such as radioactive decay and the disappearance of many different hormones from the circulation. This type of reaction is characterized by its half-life, the time required for [A] to fall from its initial value, [A]0, to one-half of its initial value, [A]0/2. In this case, Eqn [1.5.13] gives ½A0 5 ½A0 e2kt1=2 2 kf ½A½B 5 kr ½C Jf 5 Jr kf ½C 5 Keq 5 ½A½B kr where Keq is the equilibrium constant. From the units of kf as M21 s21 and kr as s21, we see that Keq has the units of M21. d½A 5 2 kf ½ A½B 1 kr ½C dt The negative sign before Jf(5kf[A][B]) in the above equation indicates that this reaction reduces the concentration of reactant A; the positive sign of Jr indicates that this reaction flux adds to the concentration of reactant A. Similar reasoning gives us the rate of change of [C]. At equilibrium, the concentrations of A, B, and C are no longer changing. The values of [A], [B], and [C] are altered from their original concentrations so that the forward and reverse rates are equal: ½1:5:9 ½1:5:10 ½1:5:15 1 ln 5 ln e2kt1=2 2 2ln2 5 2 kt1=2 ln 2 t1=2 5 k Thus the half-life of a first-order reaction is inversely proportional to the first-order rate constant. 63 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION RATES OF CHEMICAL REACTIONS DEPEND ON THE ACTIVATION ENERGY The values of kf and kr depend on the reaction path taken for the reaction, but Keq is not affected by the reaction path. We shall not prove this result here but will give some rationale for it. A more complete description is given in Appendix 1.5.A1. The reactants (A and B) and the product, C, can be viewed as possessing some degree of energy. This energy is a potential energy that consists of the potential energy of all of the interactions of their orbital electrons with the positive nuclei. The set of nuclei has some spatial arrangement which changes during the course of the reaction. This is the essence of a chemical reaction, in which the relative positions of the nuclei are altered. The orbital electrons, of course, follow the nuclei so that the energy of the ensemble changes during the course of the reaction. We call this energy the potential energy, and it includes the potential energy of the electrons and the nuclei and the kinetic energy of the electrons, but does not include the kinetic energy of the nuclei. We can plot this potential energy against the “reaction coordinate,” which is the actual distance along the minimum energy path from reactants to products. An example of such a graph is shown in Figure 1.5.2. The rate constants that govern the rates of reaction depend on the energy required to reach the activated complex intermediate between reactants and products. Large activation constants are associated with small rate constants. The relationship is expressed by the Arrhenius equation: –300 ½1:5:16 ln k 5 ln A 2 Ea RT where k is the rate constant for the reaction, A is a preexponential factor that has to do with the orientation of the reaction and not its temperature dependence, Ea is the activation energy in J mol21, R is the gas constant (58.314 J mol21 K21), and T is the absolute temperature (K). The equilibrium constant, however, depends only on the initial and final energy levels. Note that the Arrhenius equation means that a larger Ea would cause a lower rate constant, and a smaller Ea would increase the rate constant. ENZYMES SPEED UP REACTIONS BY LOWERING EA It is possible for the reaction to take a different path. For example, A and B could be absorbed onto the surface of an enzyme. The forces that aid in this binding have been discussed in Chapter 1.4: electrostatic interaction between ions in the substrate and enzyme; hydrogen bonding; dipoledipole interactions; and London dispersion forces. This binding to the enzyme changes the configuration of nuclei and alters the energy of the activated complex. In this way, the enzyme offers an alternative pathway for the reaction that involves far less activation energy. This increases the reaction rates without altering the final energetics of reactants and products. Thus the rates increase without changing the equilibrium constant. Figure 1.5.3 illustrates this idea. Activated complex A–B Activated complex –350 Uncatalyzed Ea Ea –400 –450 Initial state A+B Final state C Potential energy Potential energy (kJ mol–1) 64 Ea Initial state Intermediate steps ΔG –500 0 200 400 600 800 1000 Distance along the reaction coordinate (pm) Enzyme reduces Ea ΔG is unaltered by the enzyme Final state Distance along the reaction coordinate FIGURE 1.5.2 Potential energy along the reaction coordinate for the reaction A 1 B-C. Reactants A and B are at low potential energy. The activated complex is a form intermediate between reactants and products, and can be attained by converting kinetic energy into potential energy by a collision between A and B. The difference between the energy of the activated complex and the reactants is the activation energy that must be supplied for the reaction to proceed. The “reaction coordinate” is the distance along the minimum free-energy path from reactants to products. FIGURE 1.5.3 The effect of an enzyme on the overall activation energy for a reaction. The uncatalyzed reaction requires a large activation energy, Ea, and so the reaction occurs slowly. The enzyme offers an alternative path, often by breaking the reaction into a number of small steps, so that the reaction occurs more quickly. Enzymes do not change the overall energetics of the products with respect to the reactants (ΔG in the figure is the change in free energy in the transition between reactants and products) and so the equilibrium constant is unaffected. Chemical Foundations of Physiology II: Concentration and Kinetics THE MICHAELISMENTEN FORMULATION OF ENZYME KINETICS and dividing numerator and denominator by k1, we obtain Catalysis is the speeding up of a chemical reaction by a chemical species that does not enter into the stoichiometry of the reaction. Enzymes are catalysts because they speed up reactions without themselves being altered. Enzymes typically bind the reactants and alter their shape (the three-dimensional arrangement of all of the atomic nuclei in the reactants) by virtue of their being attracted to or bound to the surface of the enzyme. The enzyme changes the reaction from a homogeneous reaction in the solution to a heterogeneous reaction occurring on the surface of the enzyme. In this process, the enzyme itself is unchanged. It participates in the reaction but does not show up in an accounting of the reactants and products because it is unchanged. ½1:5:24 Figure 1.5.3 shows that enzymes speed up reactions by providing an alternate path for the reaction to take, and that this path requires less activation energy. Michaelis and Menten provided a simple analysis of this reaction path by imagining it to take place in two steps: k1 E 1 S " E 2 Sk3 E 1 P - k2 where E is the enzyme, S is the substrate, ES is the substrateenzyme complex, and P is the product. This is a very simple reaction mechanism. The rate of the enzyme-catalyzed reaction is defined by the rate of product release: ½1:5:18 J 5 k3 ½E 2 S If we know [ES] and k3, we can calculate the reaction rate. Here k3 is in units of s21. If we keep [S] and [P] constant, we can define a steady-state rate in which S is converted to P at a constant rate. Under these conditions, we can solve for [ES]. The rate of change of [ES] is given by ½1:5:19 ½1:5:20 k1 ½E½S 5 ðk2 1 k3 Þ½E 2 S Since the enzyme can exist in only two states, E and ES, we have a conservation relation ½1:5:21 ½Etotal 5 ½E 1 ½E 2 S where [E]total is the total concentration of enzyme. Inserting this relation for [E] in Eqn [1.5.20], we get ½1:5:22 k1 ð½Etotal 2 ½E 2 S½SÞ 5 ðk2 1 k3 Þ½E 2 S Solving for [ES], we find ½1:5:23 ½1:5:25 Jmax 5 k3 ½Etotal Inserting this definition into Eqn [1.5.24], we finally obtain ½1:5:26 J5 ½E 2 S 5 k1 ½Etotal ½S k2 1 k3 1 k1 ½S From Eqn [1.5.18], the rate of the reaction is just k3[ES]. Multiplying both sides of Eqn [1.5.23] by k3, Jmax ½S Km 1 ½S where Km is a newly defined constant called the MichaelisMenten constant. It is given as ½1:5:27 Km 5 k2 1 k3 k1 The units of k2 and k3 are both in s21, whereas the unit of k1 is M21 s21; thus, the unit of Km is M. If k3{k2, Km approximates the value of k2/k1, which is the dissociation constant for binding of S to the enzyme. It can be obtained experimentally as the value of the substrate concentration at which the enzyme exhibits one-half maximal velocity. This can be seen from Eqn [1.5.26] by inserting J 5 1/2Jmax and finding that [S] 5 Km when the rate of the reaction is one-half maximal. The saturation curve for a MichaelisMenten type reaction is shown in Figure 1.5.4. In this case, Jmax was 8 μmol min21 mL21. At one-half of this maximal velocity, the substrate concentration was 1.5 mM. The curve shown in Figure 1.5.4 often lacks sufficient points to accurately extrapolate the observed velocity Vmax 10 d½E 2 S 5 k1 ½E½S 2 k2 ½E 2 S 2 k3 ½E 2 S dt At steady state, d[E 2 S]/dt 5 0, so that k3 ½Etotal ½S ðk2 1 k3 Þ=k1 1 ½S The maximum velocity occurs when all of the enzyme is present as [ES]. The maximum velocity or rate of the reaction is thus given as Velocity (μmol mL–1 min–1) ½1:5:17 J 5 k3 ½E 2 S 5 8 1/2V max 6 Km is the concentration of S when V is 1/2 maximum 4 2 0 0 5 10 15 [S] (mM) 20 25 FIGURE 1.5.4 Saturation curve for an enzyme that obeys Michaelis Menten kinetics. The velocity at low [S] is nearly linear with [S] but quickly levels off. This behavior is called saturation kinetics. The concentration of substrate at one-half maximal velocity (5one-half saturation) is equal to the Km, the MichaelisMenten constant. 65 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION to the maximal velocity. To aid in this graphical determination of enzyme kinetics, the MichaelisMenten equation is transformed. If we take the inverse of Eqn [1.5.26], we obtain J5 Jmax ½S Km 1 ½S 1 Km 1 ½S 5 J Jmax ½S ½1:5:28 1 1 Km 1 5 1 J Jmax Jmax ½S This equation suggests that a plot of 1/J versus 1/[S] should be linear with an intercept on the 1/J axis of 1/Jmax and a slope of Km/Jmax. The intercept on the 1/[S]-axis is 21/Km. Such a plot is called a doublereciprocal plot, also known as a LineweaverBurk plot. An example for the results shown in Figure 1.5.4 is shown in Figure 1.5.5. Throughout this discussion of enzyme rates, we have used J as the symbol of the rate of product release, or the rate of the reaction. J has the units of moles per unit volume, per unit time. Often this is taken to be the same as the reaction velocity, and the usual symbol in the MichaelisMenten equation is V, and not J. Sometimes, the reaction velocity in these units is normalized by dividing by the enzyme concentration, so that the reaction velocity is given as a specific activity of the enzyme, in units of moles of reaction per unit enzyme per unit time. These various representations of the velocity all obey the same general form of equations and can be subjected to the same analysis. 1/V intercept = 1/Jmax 0.4 Slope = Km/Jmax 0.3 0.2 X-intercept = –1/Km 0.1 –1.0 –0.5 0.0 0.5 1.0 1/[S] PHYSIOLOGY IS ALL ABOUT SURFACES As noted above, enzyme catalysis derives from the reaction occurring on the surface of the enzyme rather than in homogeneous solution. Physiology is all about surfaces and their interactions. Proteins have complicated three-dimensional shapes that closely appose the surfaces of some materials and not others—they are specific. Sometimes the fit is very close and therefore the binding strength is very high—the two surfaces form a tight association. Which proteins stick to which other proteins, or which other substrates, or ligands, determines the activity of the proteins, which determines the activities of the cells, and then the organs, and, finally, the organism. Ultimately, physiology is all about what happens on the surfaces of molecules. SUMMARY The concentration of a solute in solution is the amount of that solute per unit volume of solution. It can be expressed as the mass of the solute per unit volume or the number of moles of solute per unit volume. A mole of any substance is Avogadro’s number of particles. Avogadro’s number originates in the ratio of the mass of a carbon-12 atom to 12 g. The dalton is defined as 1/12 of the mass of a carbon-12 atom. The number of daltons in 1 g is Avogadro’s number, 6.02 3 1023. The concentration in molar is the number of moles per liter of solution. Small concentrations use established prefixes to describe them, in increments of 1000. A millimolar solution is 1023 M, micromolar is 1026 M, nanomolar is 1029 M, and picomolar is 10212 M. These same prefixes are used for units of g, L, and m. The relationship among concentration, amount of solute, and volume of solution can be used to determine the volume of physiological fluids. Evans’ Blue Dye is an example of a solute that can be used to estimate plasma volume, because the dye enters the plasma but cannot leave it easily. 0.5 1/Velocity 66 1.5 2.0 2.5 FIGURE 1.5.5 LineweaverBurk plot. Values of 1/[S] are plotted along the abscissa while values of 1/J are plotted on the ordinate. Here J is the velocity, or rate, of the enzyme reaction and [S] is the substrate concentration. The slope of the line is Km/Jmax, where J is the enzyme flux enzyme velocity and Jmax is the theoretical maximal rate. Km is the MichaelisMenten constant. It is also determined as the X-intercept 5 21/Km. The Y-intercept is 1/Jmax. Deviations from linearity on the LineweaverBurk plot suggest that the enzyme does not obey MichaelisMenten kinetics. Elementary chemical reactions have forward and reverse rate constants that govern the rate of conversion in either the forward or reverse reaction. The rates of reaction have the units of moles per unit time per unit volume of solution. The ratio of the forward and reverse rate constants is the equilibrium constant. Conversion of reactants to products requires an activation energy, and because kinetic energy increases with temperature, the rate of the reaction also varies with the temperature. The relationship between temperature and rate is described by the Arrhenius equation, which incorporates the activation energy, Ea. Enzymes speed chemical reactions by altering the path of the reaction by allowing it to proceed on the surface of the enzyme. Thus enzymes convert homogeneous reactions in the fluid phase to heterogeneous reactions on the surface of the enzyme. The alternate path reduces the activation energy for the reaction, thereby allowing it to proceed quicker. Chemical Foundations of Physiology II: Concentration and Kinetics The MichaelisMenten formulation of enzyme kinetics describes the steady-state rate of enzyme reactions and is derived for a particular kind of enzyme mechanism. Nevertheless, the resulting equation often fits many enzyme-catalyzed reactions. It is given as J5 Jmax ½S Km 1 ½S where J is the reaction flux or velocity, in units of moles of completed reaction per unit time per unit volume, Jmax is the asymptotic maximal rate achievable extrapolated at infinite concentration, Km is the Michaelis Menten constant, equal to the concentration of substrate at half maximal velocity, and [S] is the substrate concentration. This equation describes saturation kinetics. It can be analyzed more easily using the inverse of the equation, rearranged, to give 1 1 Km 1 5 1 J Jmax Jmax ½S Plots of 1/J against 1/[S] yield 1/Jmax as the intercept on the abscissa and 21/Km as the extrapolated intercept on the ordinate. REVIEW QUESTIONS 1. What is a dalton? What is meant by molecular weight? 2. What is a mole? What is meant by the gram molecular weight? How would you determine the gram molecular weight of small compounds? 3. Write the relationship among concentration, volume, and amount. 4. What is meant by “first-order reaction”? 5. How do you calculate the half-life of a reaction? 6. In the plot of potential energy against reaction coordinate, what is meant by “reaction coordinate”? What units does it have? 7. What is the activation energy? 8. Why do reaction rates generally depend on the temperature? 9. How could you determine the activation energy for a reaction? 10. In general, how do enzymes speed up biochemical reactions? 11. How would you determine Km and Vmax for an enzyme? APPENDIX 1.5.A1 TRANSITION STATE THEORY EXPLAINS REACTION RATES IN TERMS OF AN ACTIVATION ENERGY TRANSITION STATE THEORY CALCULATES THE POTENTIAL ENERGY OF REACTANTS AS A FUNCTION OF SEPARATION Transition state theory gives an expression for the rate constant of a chemical reaction by applying statistical mechanics to quantum mechanical calculations of various configurations of reactant and product. In this procedure, a collection of nuclei with their attendant electrons is treated as a “supermolecule.” Quantum mechanical calculations are performed in which the potential energy of the supermolecule is calculated as a function of the relative positions of the nuclei. An exact solution requires solving for the total energy including the kinetic energy terms of all of the nuclei and electrons and the potential energy terms for all electronelectron, electronnucleus, and nucleusnucleus pairs. Since this solution is extremely difficult, some simplifying assumptions are made. One is that the electronic motion is extremely rapid compared to translation of the nuclei and that the electrons adjust instantly to any change in the positions of the nuclei. Thus the energy of the electrons and potential energy of the nuclei can be calculated as if the nuclei were at rest. The energy calculated in this way includes everything but the kinetic energy of the nuclei. It is called the potential energy even though it includes the kinetic energy of the electrons. POTENTIAL ENERGY CAN BE GRAPHED AGAINST SEPARATION IN A SINGLE MOLECULE The potential energy of a configuration of nuclei with their electrons may be represented as a point on an f-dimensional surface in an f 1 1-dimensional space, where f is the number of independent variables required to specify the relative positions of all nuclei. For a diatomic molecule (like H2 or HF), f 5 6 because we need six variables to specify the Cartesian coordinates of the two nuclei. However, three of these can be considered to locate the center of mass of the molecule and another two specify the orientation of the molecular axis. As these do not really concern us, we have only one remaining variable, the internuclear distance, to specify the relative positions of the two nuclei. Since f 5 1, we have a potential surface which is the one-dimensional potential energy curve in the twodimensional graph for a diatomic molecule (H2) as shown in Figure 1.5.A1.1. POTENTIAL ENERGY CAN BE GRAPHED AGAINST NUCLEI SEPARATION IN A CHEMICAL REACTION In a configuration of three nuclei, the potential energy surface can be shown as a fixed-angle surface, where the angle is defined as the angle between one bond and the approaching reactant, as shown in Figure 1.5.A1.2. This allows us to plot the potential surface on paper; otherwise we need another physical dimension. The entire potential surface consists of an infinite number of these fixed-angle surfaces, one for each angle of approach. Such a fixed-angle surface is shown in Figure 1.5.A1.3 for the reaction F 1 H 2 H0 -F 2 H 1 H0 The surface itself is three-dimensional. What Figure 1.5. A1.3 shows is the two-dimensional projection of the surface onto the plane of the paper. Here every point on a given line has the same potential energy. Thus these lines represent potential energy contours in much 67 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION broken. The horizontal cross-section at “Z” gives the potential energy dependence of FH on its bond distance. +300 +200 Potential energy (kJ mol-1) 68 THE ACTIVATED COMPLEX IS A METASTABLE COMPLEX OF REACTANTS +100 0 –100 –200 –300 –400 –500 0 100 200 300 400 rHH' Internuclear distance (pm) 500 FIGURE 1.5.A1.1 Potential energy diagram for a diatomic molecule, H2. Potential energy is plotted as a function of the separation of the two H nuclei. The average bond length corresponds to the minimum energy level, which in this case occurs at about 80 pm. The horizontal lines indicate quantum vibrational states. Rotation of the molecule alters the potential energy profile due to a centrifugal effect (rotation stretches the bond length). Adapted from J.W. Moore and R.G. Pearson, Kinetics and Mechanism, John Wiley & Sons, New York, NY, 1981. C 135° A B FIGURE 1.5.A1.2 Fixed angle of approach of reactant C to the diatomic molecule AB. The potential energy surface varies with the angle of approach. the same way as contour lines on a topographic map represent lines of equal altitude. In the contour map for the reaction F 1 H 2 H0 F 2 H 1 H0 , several regions are labeled. Region “W” corresponds to separated F, H, and H0 nuclei and this is a plateau region where potential energy varies little with the distance between the nuclei. At region “X,” F is far removed from HH0 and the potential energy, U, is affected little by changes in rFH, the distance between the F nucleus and the H nucleus. The effect of rHH on U is that shown earlier for the diatomic molecule, so that a vertical cross-section at “X” would show a deep valley with steep sides. At region “Y,” all three nuclei are close together and the potential energy has increased due to van der Waals repulsion. A collision of F with HH0 thus corresponds to the movement of the configuration of nuclei from region “X” to region “Y” where potential energy is increased. If the reaction is completed, then the configuration moves on to region “Z.” In this region, rFH is short and rHH’ is long, indicating that a bond has formed between FH and the HH is The movement of the configuration of nuclei from X to Y to Z in Figure 1.5.A1.3 traces a completed reaction from the initial reactants. At the point Y, there is a saddle point represented by higher potential energy, in this case, than for either reactants or products. The configuration of nuclei at this point is called the activated complex which is said to be in the transition state. Here all of the nuclei are close together, forming a complex. The path of steepest ascent from reactants to activated complex and on to products is called the reaction coordinate. This path is perpendicular to each contour it crosses. It is the minimum energy path from reactants to products. Recall that the potential energy surface is just that—a potential energy. Collisions between reactants can overcome the potential energy barrier between the reactant and the activated complex if there is sufficient kinetic energy of the molecules that collide. The potential energy can be plotted against the distance along the reaction coordinate. This is the familiar diagram encountered in general chemistry, without any kind of explanation of what the reaction coordinate actually is. For our example reaction, F 1 HH0 -FH 1 H0 , the potential energy versus reaction coordinate plot is shown in Figure 1.5.A1.4. This plot is the conceptual origin of Figure 1.5.2 in the text. REACTIONS GENERALLY DO NOT FOLLOW A MINIMUM ENERGY PATH Reactions in general do not follow the reaction coordinate. Two alternative trajectories are shown in Figure 1.5. A1.5. Because the HH0 bond vibrates, the configuration of all three nuclei while F approaches HH0 oscillates. In Figure 1.5.A1.5A, the reaction ascends near to the activated complex but does not go to completion. This trajectory represents a nonreactive, inelastic collision between F and HH0 . In this case, there is a transfer of energy from translational kinetic energy to vibrational energy. In Figure 1.5.A1.5B, a completed reaction is shown. Here the oscillations correspond to vibrations of reactant HH0 and product FH. THE TRANSITION STATE THEORY SAYS THAT THE RATE CONSTANT VARIES WITH THE EXPONENT OF THE ACTIVATION ENERGY The derivation of a rate equation from the transition state theory is a bit complicated and we will not attempt it in any detail. However, an elementary understanding of it can be provided by thinking of the activated complex as a separate, transient species. Then the rate of completed reaction will be proportional to the amount of activated complex, and the rate per unit volume will be proportional to its concentration. If we imagine that the reactants and +200 l mo (kJ y g ner –1) +300 W +100 0 ia l e ent Pot -100 W -200 -300 This point is H + FH -400 500 Potential energy isoclines (kJ mol–1) Z 400 -510 -470 W -440 This point is the activated complex +100 300 200 -290 -200 -100 0 100 0 100 200 300 400 rFH Internuclear distance (pm) -500 0 -400 X -300 Y -100 X y rg J +200 -320 +300 -410 -380 -350 -200 rHH' Internuclear distance (pm) This point is H + F + H Z -500 -1) ol m (k ne e al ti en t Po This point is the reactants, F + H-H FIGURE 1.5.A1.3 Potential energy contour diagram for a fixed angle (180 ) for the reaction F 1 HH0 -FH 1 H0 . The lines represent a set of points of equal potential much like contour lines on a topographic map represent a set of points of equal altitude. The value of the potential energy (in kJ mol21) for each line is indicated. Point W lies on a plateau of high potential energy corresponding to dissociated F, H, and H0 nuclei. X corresponds to the configuration in which F is far from HH0 ; it is at a potential energy minimum. Y is the saddle point at which all three nuclei are close; it corresponds to the activated complex. Z corresponds to the configuration in which H0 is far from FH. It represents the products of the completed reaction. The graph at the right is a vertical cross-section through the three-dimensional surface at point X. The graph at the top is a vertical crosssection (orthogonal to the one at the right) through the three-dimensional surface at point Z. Adapted from J.W. Moore and R.G. Pearson, Kinetics and Mechanism, John Wiley & Sons, New York, NY, 1981. –300 the activated complex are in equilibrium, then the rate constant for the overall reaction will be proportional to the equilibrium constant. Activated complex Y Potential energy (kJ mol–1) –350 ½1:5:A1:1 Ea –400 –450 Initial state X Final state Z –500 0 200 400 600 800 1000 Distance along the reaction coordinate (pm) FIGURE 1.5.A1.4 Potential energy along the reaction coordinate for the reaction F 1 HH0 -FH 1 H0 . Reactants F 1 HH0 are at a low potential energy. The activated complex is at higher potential energy, which can be attained by converting kinetic energy into potential energy through a collision of F and HH0 . The reactants are at lower potential energy. The difference between the energy of the activated complex and the reactants is the activation energy that must be supplied for the reaction to proceed. Kr 5 κ kT K h There are a lot of k’s in this equation. Kr is the rate constant for the reaction; κ is a transmission coefficient that tells us what fraction of activated complexes goes on to complete the reaction; k in kT is Boltzmann’s constant, which is equal to the gas constant per molecule, or R/No, the gas constant divided by Avogadro’s number; h is Planck’s constant; and K* is a constant that has the form of an equilibrium constant for the formation of the activated complex, but strictly speaking it is not an equilibrium constant. The equilibrium constant is related to the change in free energy for the reaction (see Chapter 1.7) according to ½1:5:A1:2 2ΔG RT K 5 e where ΔG* is the free-energy change per mole for the formation of the activated complex from the reactants. This energy change is identified with the activation energy, Ea, as described in Figure 1.5.A1.4. 500 400 rHH'' Internuclear distance (pm) 300 Y X 0 X 100 400 300 200 Y (B) Z 200 (A) Z 100 rHH'' Internuclear distance (pm) 500 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 0 70 0 100 200 300 400 rFH Internuclear distance (pm) 500 0 100 200 300 400 rFH Internuclear distance (pm) 500 FIGURE 1.5.A1.5 Hypothetical trajectories of reactions. (A) An inelastic collision that does not result in a completed reaction. (B) A completed reaction. The oscillations are due to vibration of the molecule in which bond length oscillates around an average value. In 1884, van’t Hoff proposed that the temperature dependence of equilibrium constants could be described by ½1:5:A1:3 d ln Keq ΔG0 5 dT RT 2 It was discovered shortly thereafter that the rates of chemical reactions approximately double for every 10 C increase in temperature. By analogy to the van’t Hoff law, Arrhenius proposed in 1889 that the temperature dependence of the rate constant could be given as ½1:5:A1:4 d ln Kr ΔE0 5 dT RT 2 The integrated form of this equation is similar to what we see from transition state theory: ½1:5:A1:5 Kr 5 Ae2Ea =RT where A is the preexponential factor that incorporates the transmission coefficient. This is close to what one would expect by inserting the definition of K* into Eqn [1.5.A1.1]. The natural log of both sides gives ½1:5:A1:6 Ea ln Kr 5 ln A 2 RT Thus the plots of the logarithm of the rate constant against 1/T should be linear with a slope 5 2Ea/R. Such a plot is an Arrhenius plot. Although transition state theory does not predict a perfectly linear relation between ln Kr and 1/T, most reactions obey Eqn [1.5.A1.6] well. more complicated because of the large number of atomic nuclei involved in typical biochemical reactions. However, the idea of the reaction coordinate and activation energy remains. By absorbing reactants onto their surfaces, enzymes completely alter the potential energy surfaces of a reaction. The net effect is to lower Ea. According to Eqn [1.5.A1.6], lower activation energies mean larger rate constants. Thus enzymes and other catalysts increase the speed of reactions by providing an alternate mechanism such that Ea is reduced. This idea is depicted graphically in Figure 1.5.3 in the text. APPENDIX 1.5.A2 UNIDIRECTIONAL FLUXES OVER A SERIES OF INTERMEDIATES DEPEND ON ALL OF THE INDIVIDUAL UNIDIRECTIONAL FLUXES UNIDIRECTIONAL FLUXES DIFFER FROM NET FLUXES Consider the diagram shown in Figure 1.5.A2.1. Each node, indicated by a dot, represents a state, compartment, or individual chemical species in a series. Movement between nodes indicates either a flux of material between compartments or the transition of one chemical species into another, or the transition between intermediate states of an enzyme, for example. We will treat all of the transitions between states as being an elementary reaction so that the flux obeys the relation THE ACTIVATION ENERGY DEPENDS ON THE PATH ½1:5:A2:1 The potential energy surfaces described in Figure 1.5. A1.3 pertain only to the collisions between the molecules HH0 and F, which is a simple system. Similar potential energy surfaces could be described for more complicated reactions, but the surface becomes much where Jij is the flux from node i to node j, Ni is the population of node i, and αij is a pseudo-first order rate constant. It is called a pseudo-first order rate constant because it may incorporate the concentration of a ligand if the transition between node i and node j requires it. Jij 5 αij Ni Chemical Foundations of Physiology II: Concentration and Kinetics J12 Jnet N1 J21 Jnet J23 N2 J32 N3 FIGURE 1.5.A2.1 Chemical species can be transformed into various intermediates here represented as three nodes, N1, N2, and N3. The fluxes between the nodes are indexed with the source node first and sink node second. Alternatively, the nodes can represent compartments into which the same chemical is transported, or different states of the same object, such as an enzyme. At steady state, there could be a throughput of material such that sum flux, Jnet, of material enters node 1 and the same flux exits the end, at node 3 in this case. At steady state, this requires that the net flux between any two nodes is the same, equal to Jnet. However, there can be a higher unidirectional flux of material from node 1 to node 2 than this net flux. All that steady state requires is that ½1:5:A2:2 Jij 2 Jji 5 Jnet for each ij pair. The unidirectional flux, Jij, can be determined by somehow labeling the population of state i and seeing how fast they are converted to state j. One way this can be done is by adding a pulse of radioactive material of state i, and watching its conversion to state j. We will note the population of radioactive tracer in each state or compartment with an asterisk, which is added into state i alone. We can write the following sets of equations that govern the time change in the population of radioactivity in each state: ½1:5:A2:3 Measuring the unidirectional flux from state 1 to state 3 requires a similar condition as the unidirectional flux from state 1 to 2; that is, there must be negligible back flux during the measurement. In this case, this means that N3*/N3 is negligible. We also require a steady state. In this case, steady state means dN2*/dt 5 0. What we are searching for is an overall unidirectional flux that obeys the relation: dN3 N 5 J13 1 dt N1 ½1:5:A2:5 Because N3*/N3 0 as a criterion for measuring unidirectional flux, the bottom part of Eqn [1.5.A2.3] becomes dN3 N 5 J23 2 dt N2 ½1:5:A2:6 If dN2*/dt 5 0 as a condition of steady state, then by the middle part of Eqn [1.5.A2.3] we can write N1 ðJ21 1 J23 Þ N2 5 J12 N1 N2 ½1:5:A2:7 Substituting in for N2*/N2 from Eqn [1.5.A2.3] into Eqn [1.5.A2.7], we obtain dN3 J12 J23 N1 5 dt ðJ21 1 J23 Þ N1 ½1:5:A2:8 dN1 N N 5 2 J12 1 1 J21 2 dt N1 N2 Comparison of Eqn [1.5.A2.5] with Eqn [1.5.A2.8] allows us to identify J13 as dN2 N N N 5 J12 1 2 ð J21 1 J23 Þ 2 1 J32 3 dt N1 N2 N3 ½1:5:A2:9 dN3 N N 5 J23 2 2 J32 3 dt N2 N3 This result has an intuitive interpretation. It says that the unidirectional flux from node 1 to node 3 is the flux from node 1 to node 2 times the proportion of the flux that goes on to node 3. This proportion is the flux from node 2 to node 3 divided by the total flux away from node 2: the sum of J21 and J23. The idea about unidirectional flux is that there is no back flux. Thus, for the transition between state 1 and state 2, we can measure unidirectional flux J12 only when N2* is negligible. For the top equation in Eqn [1.5.A2.3], we have dN1 N N 5 2 J12 1 1 J21 2 dt N1 N2 N2 ½1:5:A2:4 is called the specific activity of the added radioactivity. The units of J12 are in units of N per unit time. N2 dN1 dt 0 5 2 J12 N1 N1 dN1 N1 J12 5 2 dt N1 The last equation in Eqn [1.5.A2.4] defines what is meant by a unidirectional flux obtained by tracer means. In this case, dN1*/dt is negative—the amount of tracer in state 1 decreases with time—and thus the flux J12 is positive. The factor N1*/N1 in the denominator J13 5 J12 J23 J23 5 J12 ð J21 1 J23 Þ ðJ21 1 J23 Þ It can be seen readily that we could just as easily added radioactivity at node 3 and watch its appearance in node 1. From the symmetry, we can write J31 as ½1:5:A2:10 J31 5 J32 J21 J21 5 J32 ðJ21 1 J23 Þ ðJ21 1 J23 Þ These results are completely general. Suppose that the sequence of nodes, states, or compartments was longer than just three, as shown in Figure 1.5.A2.2. We could use the unidirectional flux to reduce the diagram to one fewer states, and then do so again using the same principle that we discovered here. The results for J14 and J41 are given as J13 J34 5 J14 5 ðJ31 1J34 Þ ½1:5:A2:11 J23 J34 J12 J23 J34 ðJ21 1 J23 Þ 5 J21 J32 J21 1J21 J34 1J23 J34 J32 1J34 J21 1 J23 J12 71 72 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION J12 Jnet N1 Jnet J21 J23 N2 J32 J34 N3 J31 Jnet N4 J34 J13 N1 J43 N3 J43 J41 J54 N4 J54 J54 A P B P Jnet FIGURE 1.5.A3.1 Two-pool sequential model. Pool A of size A is converted to pool B of size B with a steady-state rate of P. Jnet turnover or the rate of renewal of the pool. The fractional turnover rate is the fraction of the total pool which is replaced per unit time. If A is the size of pool (in g or moles) and P g or moles of A are renewed each minute, then the fractional turnover rate is given as N5 J45 N4 P N5 J45 J14 N1 Jnet J45 N5 FIGURE 1.5.A2.2 Unidirectional fluxes over a series of intermediate states. The flux between states 1 and 3 can be replaced by two unidirectional fluxes. The process can be repeated over any number of intermediates. J21 J32 J43 J31 J43 J32 J21 ð J211J23 Þ 5 J41 5 5 J21 ð J31 1J34 Þ J32 J211 J21 J34 1J23 J34 J32 1J34 J211J23 ½1:5:A2:12 J43 In this way, the unidirectional flux over any number of intermediate states can be calculated from the set of individual unidirectional fluxes. APPENDIX 1.5.A3 SIMPLE COMPARTMENTAL ANALYSIS THE AGGREGATE OF A CHEMICAL SPECIES CONSTITUTES ITS BODILY “POOL” Compounds in the body are normally in a dynamic state in which materials are converted from precursors to products. For example, there is a steady state between the phospholipids in the cytoplasm and in the mitochondria, with continuing interconversion between both. The aggregate of the cytoplasmic phospholipids in the its pool, as is the aggregate of the mitochondrial phospholipids. A “pool” or “compartment” is defined as the set of molecules of a specific compound or group of compounds found in a specific part of the organism. For example, we may speak of “liver glycogen” or “muscle glycogen” or “liver phospholipid” or “plasma phospholipid.” In investigating these pools, we are most interested in the identity of the pools, their size, and the rate at which they exchange material with other pools. Radioactive labeling is an effective way of determining these parameters. Generally these pools change only slowly with time. Their maintenance at a fairly steady level is a consequence of homeostasis, the maintenance of a constant internal environment, which is the hallmark of physiological systems. Though blood glucose or blood [Ca21] is maintained fairly constant, there is always material entering and leaving the plasma pool. THE TURNOVER DESCRIBES THE RATE OF EXCHANGE OF A POOL The rate of exchange of material is the rate at which material enters or leaves the pool, and can be given in units of amount per unit time. This is also called the ½1:5:A3:1 k1 5 P A where the units of k are reciprocal time. Here we consider pool A of size A which is a precursor of pool B of size B, as shown in Figure 1.5.A3.1. We will assume that the pools are in a steady state (their sizes do not change with time); there is rapid and uniform mixing within the pools and that the rate of transfer does not change with time. Suppose we inject an amount of radioactive compound A, which we denote as A*, into pool A at time t 5 0. If the radioactivity mixes well and uniformly, we may describe the change in radioactivity in pool A as dA A 52P dt A This equation derives from the idea that the radioactive portion of pool A will be turned over in proportion to its concentration in pool A. Thus when P unit of pool A turns over per unit time, PA*/A unit is radioactive and P (A 2 A*)/A unit is not radioactive. The quantity A*/A is called the specific activity of pool A, denoted as SA. Equation [1.5.A3.2] can be rewritten as ½1:5:A3:2 ½1:5:A3:3 1 dA P 5 2 SA A dt A Since A is constant under steady-state conditions, we may rewrite Eqn [1.5.A3.3] as ½1:5:A3:4 dðA =AÞ P 5 2 SA dt A dSA 5 2 k1 SA dt Here we have used the definition of the fractional turnover rate given in Eqn [1.5.A3.1]. The last equation is of the form of first-order decay and is easily integrated to give ½1:5:A3:5 SA 5 SA0 e2k1 t where SA0 is the specific activity of pool A at time t 5 0. Plots of ln SA against time will allow the calculation of the fractional turnover rate as the negative of the slope. Extrapolation back to time zero gives the pool size: if A*/A is known, and the amount of radioactivity injected, A*, is known, then A can be calculated. It is important to note that the determination of the specific Chemical Foundations of Physiology II: Concentration and Kinetics activity, SA 5 A*/A, does not require knowing A; it does require knowing A* and A in an aliquot of the pool. Thus the specific activity can be determined even if the pool size A cannot be directly measured. The validity of estimating pool sizes in this way depends on the assumptions of uniform and rapid mixing of the radioactive label with the endogenous material. Integration from t 5 t to t 5 t gives dðρ SB Þ dB 5 P SA 2 P SB dt 1 dB P 5 ðSA 2 SB Þ B dt B dB =B 5 k2 ðSA 2 SB Þ dt dSB 5 k2 ðSA 2 SB Þ dt We can rewrite the last part of this equation as ½1:5:A3:8 dSB 1 k2 SB 5 k2 SA dt This equation cannot be solved by integration directly, because SA is a function of time and we don’t know SB as a function of time. We can multiply both sides by an integrating factor, ρ, such that ½1:5:A3:9 dðρSB Þ 5 ρ dSB 1 ρ k2 SB dt We choose ρ such that ρSB is an exact differential, so that ½1:5:A3:10 dðρSB Þ 5 ρ dSB 1 SB dρ Comparison of Eqns [1.5.A3.9] and [1.5.A3.10] indicates that ½1:5:A3:11 ½1:5:A3:16 SB ek2 t 5 k2 SA0 ðk2 2k1 Þt ðe 21Þ k2 2 k1 k2 SA ðe2k1 t 2 e2k2 t Þ k2 2 k1 0 We have now derived equations for SA and SB for the conditions shown in Figure 1.5.A3.1. The graph of ln SA versus time will give ln SA0 as the intercept and 2 k1 as the slope. Eqn [1.5.A3.6] shows that the maximum of the graph of ln SB will occur with SA 5 SB, because at this point dB*/dt 5 0, which means dSB/dt 5 0 and therefore d ln SB/dt 5 0. P may be obtained from A and k1. Figure 1.5.A3.2 shows the results of calculations for Eqns [1.5.A3.5] and [1.5.A3.16] for assumed values of A, A*, k1 and k2. The plot of ln SA gives an intercept of 9.2103. Thus SA0 5 10,000 cpm μmol21. Since the amount of injected radioactivity was 1 3 108 cpm, we can calculate the pool size, A, as A 5 A =SA0 5 1 3 108 cpm=104 cpm μmol21 5 104 μmol 5 0:02 mol; thus A 5 0:02 mol The fractional turnover rate is given as the negative of the slope: k1 5 0.02 min21. The turnover of the pool at steady state can be calculated as P 5 k1 A 5 4 3 1024 mol min21 ; P 5 4 3 104 mol min1 ln SA0 = 9.2103 9.0 ρ5e k2 t dðρ SB Þ 5 ρ k2 SA dt 2k1 t dðρ SB Þ 5 k2 e SA0 e k2 t ln SA vs Time ln SB vs Time 8.5 ln SA or ln SB ρ dSB 1 ρ k2 SB dt 5 ρ k2 SA dt 8.0 7.5 7.0 –k1 = –0.0200 6.5 6.0 Integration of this equation is now possible because the left-hand side is a function of ρSB alone and the right-hand side is a function of t alone. Substitution of ρ from Eqn [1.5.A3.12] and for SA from Eqn [1.5.A3.5] gives ½1:5:A3:15 k2 SA0 ðk2 2k1 Þt t e j k2 2 k1 0 SB 5 9.5 We have chosen ρ so that Eqn [1.5.A3.9] is valid, so we can substitute d(ρSB) for the left-hand side of Eqn [1.5.A3.13]: ½1:5:A3:14 5 dρ 5 ρk2 dt Multiplying both sides of Eqn [1.5.A3.8] by this multiplication factor gives ½1:5:A3:13 ρ SB j 0 Solution of this equation gives ½1:5:A3:12 0 t Dividing both sides by the exponent on the left finally leaves us with We can divide both sides of the equation by B to obtain ½1:5:A3:7 5 k2 ek2 t SA0 e2k1 t dt 0 ½1:5:A3:16 The specific activity of pool B is affected by influx from pool A and efflux from pool B. Here we write ½1:5:A3:6 ðt ðt dt 5.5 0 20 40 60 80 100 120 140 160 Time (min) FIGURE 1.5.A3.2 Plot of ln SA and ln SB against time. The plot of ln SA is linear with a Y-intercept of 9.2103 and a slope of 20.02. The plot of ln SB intersects that of ln SA at the point where SA 5 SB. 73 74 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Note that the plot of ln SB peaks at the intersection of the ln SA and ln SB curve. This occurs at 40.55 minutes and SA 5 SB 5 4444.2 cpm μmol21. Eqn [1.5.A3.16] can be reorganized to attempt to solve for k2. It can be rewritten as ½1:5:A3:17 ðk2 2 k1 Þ SB 5 SA0 e2k1 t 2 SA0 e2k2 t k2 k1 12 SB 5 SA 2 SA0 e2k2 t k2 k1 12 SB 5 SB 2 SB ek1 t e2k2 t k2 2k1 t k1 e 2k2 t 5 k2 e This result predicts that the point of intersection of the two curves depends only on the two fractional turnover rates and not on the amount of radioactivity injected. This makes intuitive sense. We know k1 and the time of intersection t 5 40.55 min. Thus k1 e2k1t 5 0.008888 min21 and we search for k2 such that k2 e2k2 t 5 0.008888 min21. The solution is k2 5 0:03 min1 which can be verified by substitution. This solution can be obtained graphically by the intersection of the line y 5 ln k2 with the line y 5 40.55k2 2 4.723. The pool size B can be calculated from k2 5 P/B; we have k2 5 0.03 min21 and P 5 4 3 1024 mol min21, giving B as B 5 4 3 1024 mol min21 =0:03 min21 5 0:0133 mol; B 5 0:0133 mol Diffusion Learning Objectives G G G G G G G G G G Define flow and flux Describe the meaning of the continuity equation Write Fick’s First Law of Diffusion Recognize Fick’s Second Law of Diffusion Identify the units of the diffusion coefficient Identify the three major assumptions of the onedimensional random walk Describe the diffusion coefficient in terms of the parameters of the one-dimensional random walk Describe what is meant by the time of diffusion and its dependence on distance Describe the diffusion coefficient in the cytoplasm compared to that in water Write the StokesEinstein equation and identify the parameters in it FICK’S FIRST LAW OF DIFFUSION WAS PROPOSED IN ANALOGY TO FOURIER’S LAW OF HEAT TRANSFER Adolph Fick (18291901) was a German physiologist who enunciated what we now call Fick’s First Law of Diffusion in 1855. Fick argued from analogy to two well-known laws of physics involving flows and their driving forces. The first was Fourier’s Law of Heat Transfer: ½1:6:1 @T JH 5 2λ @x where JH is the rate of heat energy transfer per unit area per unit time and @T/@x is the temperature gradient. The second law was Ohm’s law: ½1:6:2 Je 5 2 σ @ψ @x where Je is the electrical flux and @ψ/@x is the voltage or potential gradient. By analogy, Fick wrote: ½1:6:3 Js 5 2D @C @x 1.6 This is Fick’s First Law of Diffusion in one dimension. This law says that the positive J is in the direction of the negative spatial slope of the concentration. In analogy to the other laws, this law says that solutes move from regions of high concentrations to low concentrations, and that the driving force for such a movement is the concentration gradient. The term gradient has a specific meaning that is defined by vector calculus, as described in Chapter 1.3. Fluxes are usually expressed in units of moles per square centimeter per second. Since concentration is in units of moles cm23, and x is in units of cm, the gradient is in units of moles cm24. Dividing the units of J by the units of the gradient, we have the units of the diffusion coefficient as cm2 s21. In free water solutions, most low molecular weight materials have diffusion coefficients on the order of 1 3 1025 cm2 s21. Larger materials diffuse more slowly, as we will see. FICK’S SECOND LAW OF DIFFUSION FOLLOWS FROM THE CONTINUITY EQUATION AND FICK’S FIRST LAW In Chapter 1.2, we derived the continuity equation from the conservation of material flowing along one dimension. It was given as ½1:6:4 @Cðx; tÞ @JðxÞ 52 @t @x If we differentiate Fick’s First Law (see Eqn [1.6.3]) with respect to x, we obtain ½1:6:5 @JS @2 C 5 2D 2 @x @x By substitution into the continuity equation, we obtain ½1:6:6 @C @2 C 5D 2 @t @x This is Fick’s Second Law of Diffusion. It follows from the First Law of Diffusion by application of the continuity equation. It relates changes in concentration with time with the spatial distribution of solute particles. Given initial conditions of C(x,0) and boundary conditions, solutions of this equation, or its three-dimensional analogue, allow one to determine concentration as a function of time and position (C(x,t)). 75 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00007-0 76 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION FICK’S SECOND LAW CAN BE DERIVED FROM THE ONE-DIMENSIONAL RANDOM WALK Fick’s Second Law of Diffusion can be derived from molecular kinetic theory elucidated by Maxwell and Boltzmann in the latter half of the 19th century. This can be accomplished by analyzing the statistics of the random walk. Application of the continuity equation to the Second Law can then result in the derivation of Fick’s First Law. It should become clear, then, that both these laws describing diffusion are a consequence of the random motion of particles. The process of diffusion thus appears to be a statistical result only, there being no literal “driving force” for the diffusive flux. Despite this, we will find that there is energy associated with concentration or dilution of solutes, and that opposition of diffusion, e.g., during active transport, requires real forces. The random walk we will discuss is a one-dimensional model of real events. We choose to simplify the analysis because otherwise it is intractable and because it turns out to be a good model due to the enormously large number of collisions which occur during diffusion of real solute particles. The main assumptions of the model are: G G G Each particle moves in a straight line between collisions. The motion consists of steps of length λ taken either to the left or to the right. The probability of making a step to the right is equal to the probability of taking a step to the left. This surprisingly simple model will allow us to derive a one-dimensional form of Fick’s Second Law. If the probability of taking a step to the right is p and the probability of taking a step to the left is q, the third assumption gives p 5 q 5 1/2. The question we ask first is this: if a particle starts out at x 5 0, what is the probability that, after some elapsed time, t, it will be found some distance x away, where x is an integral multiple of λ? We can take λ to be a measure of the mean free path. If tc is the time between collisions, and we wait for the interval t, then there will be t/tc collisions in this time. These simple assumptions allow us to derive an expression for the probability density that a particle starting at x 5 0 will be found in a distance interval centered at x at time t. First we will convert the elapsed time to the number of steps taken in that interval. The size of the steps to the right and to the left is λ, the mean free path between collisions. If the average speed is v, then the average speed times the time between collisions, tc, will be equal to λ. Thus ½1:6:7 tc 5 λ hvi The number of collisions in the time t is given as the time t divided by the time between collisions: ½1:6:8 N5 t tc Note that we are using N now to count the number of steps a particle of average speed v and mean free path λ will make in the interval t. In this derivation, it does not signify the number of solute molecules. Now the original problem can be reformulated: what is the probability that, after N steps, the particle will be found at a distance x 5 mλ, where m is an integer, away from its starting place? Let R be the number of steps to the right and L be the number of steps to the left. Then ½1:6:9 N5R1L R2L5m where the total number of steps is N. To travel a distance mλ away from the starting point in steps of size λ, R 2 L 5 m. The desired probability of making R steps to the right and L steps to the left out of a total of N steps is given by the binomial probability distribution: N 1 N! ½1:6:10 PN ðR; LÞ 5 2 R!L! This is simply the probability that with N trials, R trials will be to the right and L trials will be to the left. From Eqn [1.6.9], we have N1m 2 N2m L5 2 R5 ½1:6:11 which can be substituted back into Eqn [1.6.10] to give N 1 N! PN ðR; LÞ 5 PN ðmÞ 5 2 ððN 1 mÞ=2Þ!ððN 2 mÞ=2Þ! ½1:6:12 In this form the probability distribution is a discrete variable, with only integral values of m. This random walk model of diffusion is shown in Figure 1.6.1 where the distribution models diffusion from an initially sharp distribution at m 5 0 and P 5 1.0. After 10 steps, the material is distributed between 210 # m # 10; after 20 steps, the material is distributed between 220 # m # 20, but the probability at the extreme ends of the distribution is small. The probability profile loses density in the center and gradually spreads out with the number of steps. Remember here that the number of steps is directly proportional to the time according to Eqn [1.6.8]. This behavior corresponds subjectively to the idea of diffusion: the material gradually diminishes at its source and spreads out with time. If you look carefully at the binomial probability distribution in Figure 1.6.1, you will notice that its profile appears to be a fairly well-behaved function. As written, it is a discrete probability function, being defined for only integral values of N and m. What we would like is a continuous distribution that describes the envelope of the binomial probability distribution. Now normally the value of N, the number of collisions in the time interval t is extraordinarily large. In a gas at room temperature, there are typically 5 3 109 collisions per second, while Di ffusion 0.25 Binomial distribution, N = 10 steps 0.10 Gaussian distribution (N = 50) 0.20 0.08 PN (m) 0.15 0.10 0.06 0.04 0.05 Binomial distribution 0.02 0.00 Binomial distribution, N = 20 steps 0.20 0.00 –40 –20 0 m 0.15 40 FIGURE 1.6.2 Binomial probability distribution for N 5 50, with the envelope of the Gaussian distribution with N 5 50. The binomial probability distribution is discrete, with only integral values of m allowed. The Gaussian distribution is continuous. 0.10 0.05 PN (m) 20 0.005 0.00 Binomial distribution, N = 50 steps 0.20 0.004 PN (m) 0.15 0.10 Gaussian distribution (N = 50) 0.003 Δx 0.002 X 0.05 0.001 0.00 Binomial distribution, N = 100 steps Binomial distribution 0.000 0.20 18 0.15 20 22 24 m 26 28 30 FIGURE 1.6.3 Enlargement from Figure 1.6.2. 0.10 Stirling’s formula for n factorial is pffiffiffiffiffiffiffiffiffi ½1:6:13 n! 5 2πnen lnðn2nÞ 0.05 0.00 –100–80 –60 –40 –20 0 20 40 60 80 100 m FIGURE 1.6.1 Binomial probability density for the probability of finding a particle that initially was placed at m 5 0, as a function of the number of steps, N 5 10, 20, 50, and 100. Note that material initially present in a narrow band centered at the origin spreads out with increasing number of steps, corresponding to increasing time. the mean free path is typically 1025 cm or so. We cannot count the number of steps accurately nor can we measure the distance accurately, so we resort to a simpler question: What is the probability that a particle starting at x 5 0 and t 5 0 will end up in the displacement interval Δx centered at x at time t? We can get this probability from PN(m) first by letting N get very large to justify the use of Stirling’s approximation, and then by counting the number of m’s which land the particle in the displacement interval Δx centered at x. This truly amazing formula is the key to converting the discrete probability distribution into a continuous one. Using Stirling’s formula in Eqn [1.6.13], plus the approximation that ln(1 1 α) α for small values of α, it is possible to convert Eqn [1.6.12], by a lot of algebraic manipulation, to rffiffiffiffiffiffiffi 2 2ðm2 =2NÞ e ½1:6:14 PN ðmÞ 5 πN This is the Gaussian approximation to the probability distribution function PN(m). This function provides a continuous envelope to the discrete probability function given in Eqn [1.6.12], as shown schematically in Figure 1.6.2. The Gaussian probability distribution is discussed in Appendix 1.2.A2. As mentioned earlier, what we desire is to find the probability that a particle starting at x 5 0 and t 5 0 will end up in the interval of Δx centered at x at some time t later. How we can accomplish this can be made clearer if we enlarge part of Figure 1.6.2, as shown in Figure 1.6.3. 77 78 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Since each step is of length λ, a net displacement of R 2 L 5 m steps leads to the displacement x 5 mλ. Thus, m 5 x/λ, and we may replace m in Eqn [1.6.14] with x/λ to obtain rffiffiffiffiffiffiffi x 2 2ðx2 =2Nλ2 Þ e PN ðmÞ 5 PN ½1:6:15 5 λ πN What we desired was the probability that a particle will be in the interval Δx centered at x at some time t after beginning the random walk. Specifying the time t is equivalent to specifying the value of N, as these are related according to Eqn [1.6.8] as N 5 t/tc, where tc is the time between collisions. We may substitute this value of N into Eqn [1.6.15] to obtain rffiffiffiffiffiffi x x 2tc 2ðx2 tc =2λ2 tÞ ;t 5 PN ðmÞ 5 PN e 5 Pðm; tÞ 5 P λ λ πt ½1:6:16 The probability the particle will be in the interval Δx centered at x is the sum of the probabilities that the value of mλ will fall in this interval. For a given N (which is the same as for a given time, t) values of m 5 R 2 L are either all odd or all even. Thus the possible values of displacement are separated by 2λ. The number of values of m consistent with landing in the interval Δx is thus Δx/2λ. If we take P(x/λ,t) given in Eqn [1.6.16] as the average probability in the interval Δx, then the probability of finding the particle in the interval is X Δx x P ;t Pðm; tÞ 5 2λ λ mAΔx 1 x 5 P ; t Δx ½1:6:17 2λ λ sffiffiffiffiffiffiffiffiffiffiffiffi tc 2 2 5 e2tc x =2λ t Δx 2πλ2 t From this result, we define a probability density function: rffiffiffiffiffiffiffiffiffiffiffiffi tc 2ð2tc x2 =2λ2 tÞ ½1:6:18 Pðx; tÞ 5 2 e 2πλ t When multiplied by the length of the interval, Δx, this function gives the probability of finding the particle in the interval Δx at time t. The diffusion coefficient is defined as ½1:6:19 D5 λ2 2tc Note that this definition is consistent with the units of the diffusion coefficient of cm2 s21 that we obtained from Fick’s First Law of Diffusion. Using this definition, Eqn [1.6.18] becomes rffiffiffiffiffiffiffiffiffiffiffi 1 2 e2ð2x =4DtÞ Pðx; tÞ 5 ½1:6:20 4πDt This is still the probability density function for finding a particle in an interval at some time after beginning the random walk, but all of the parameters of the random walk, the mean free path, and the time between collisions are submerged into a single constant, D. Thus the diffusion coefficient derives its values from microscopic characteristics of the diffusing substance. Since we have defined the average velocity as hvi 5 λ/tc (see Eqn [1.6.7]), we see that the diffusion coefficient is related to the square of the average velocity. Thus thermal agitation should increase the average velocity and thereby increase diffusion. We can use Eqn [1.6.20] to derive an expression for the concentration of solute particles at position x at time t, which we denote as C(x,t). Ordinarily the concentration is the number of particles per unit volume. In our onedimensional analogue, it is the number of particles per unit length. At any time, t, the number of particles per unit length, Δx, at position x is the sum of all particles which have random walked into the displacement interval from all other areas. This is the probability density function times the initial concentration summed over all intervals. This can be written as ðN ½1:6:21 C0 ðx0 ÞPðx 2 x0 ; tÞdx0 Cðx; tÞ 5 2N 0 where C0(x ) is the initial concentration at point x0 . Here rffiffiffiffiffiffiffiffiffiffiffi 1 0 2 e2ðx2x Þ =4Dt Pðx 2 x0 ; tÞ 5 ½1:6:22 4πDt Eqn [1.6.21] expresses the concentration at position x and time t in terms of the concentration everywhere at an initial time t 5 0. The initial time t 5 0 is chosen arbitrarily. We can write ðN ½1:6:23 Cðx0 ; tÞPðx 2 x0 ; ΔtÞdx0 Cðx; t 1 ΔtÞ 5 2N 0 Let us substitute in s 5 x 2 x, so that C(x0 ,t) 5 C(x 1 s,t) and P(x 2 x0 ,Δt) 5 P(2s,Δt) 5 P(s,Δt) (because of symmetry in x in the probability distribution function) and ds 5 dx0 . Then Eqn [1.6.23] becomes ðN Cðx; t 1 ΔtÞ 5 ½1:6:24 Cðx 1 s; tÞPðs; ΔtÞds 2N We let Δt be small, so that the concentration is altered only by local events. We approximate C(x 1 s,t) as Cðx 1 s; tÞ 5 Cðx; tÞ 1 s ½1:6:25 @Cðx; tÞ 1 2 @2 Cðx; tÞ 1 s 1? @x 2 @x2 which is a Taylor’s series expansion of C(x 1 s,t). Inserting Eqn [1.6.25] back into Eqn [1.6.24], we obtain ðN Cðx 1 t; ΔtÞ 5 Cðx; tÞ Pðs; ΔtÞds 2N ðN ½1:6:26 @Cðx; tÞ 1 @x 2N 1 @2 Cðx; tÞ 1 2 @x2 sPðs; ΔtÞds ðN 2N s2 Pðs; ΔtÞds Di ffusion The function P(s,Δt) is given from Eqn [1.6.22] and our definition of s 5 x0 2 x as rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 2 e2s =4DΔt Pðs; ΔtÞ 5 ½1:6:27 4πDΔt Evaluation of the integrals gives ÐN 2N Pðs; ΔtÞds 5 1 ÐN ½1:6:28 2N sPðs; ΔtÞds 5 0 ÐN 2 2N s Pðs; ΔtÞds 5 2DΔt We shall not prove these integration results. You should note their meaning, however. The first integration result is the normalization of the probability distribution, which means that the particle must be somewhere with the probability equal to 1. The second result is the average displacement about the initial zero displacement s. The zero value of the average displacement means that the distribution is symmetrical: displacement to the right and to the left are equally likely in the random walk. Recall by the definition of s 5 x0 2 x that s may be both negative and positive, so that the average is zero. The last integration gives the average square displacement. Both positive and negative values of s contribute to s2. This integration gives the variance, or the average squared displacement, for the distribution. For the case of this one-dimensional model of diffusion, the variance is 2DΔt. Inserting these integration results into Eqn [1.6.26] gives ½1:6:29 Cðx; t 1 ΔtÞ 5 Cðx; tÞ 1 1 @2 Cðx; tÞ 2DΔt 2 @x2 This may be rewritten as ½1:6:30 Cðx; t 1 ΔtÞ 2 Cðx; tÞ @2 Cðx; tÞ 5D Δt @x2 If we take the limit as Δt-0, we recognize the lefthand side of Eqn [1.6.30] as the partial derivative of the concentration with respect to time. We then have ½1:6:31 @Cðx; tÞ @2 Cðx; tÞ 5D @t @x2 This is Fick’s Second Law of Diffusion. From the continuity equation, Fick’s First Law can be derived. Although Fick’s First Law was originally derived on phenomenological grounds and the Second Law followed it from the continuity equation, this derivation shows that it can be done the other way using a quite simple model which nevertheless embodies the main ideas giving rise to diffusion: there are an enormous number of collisions which give rise to a random motion of particles from one region of space to another. THE TIME FOR ONE-DIMENSIONAL DIFFUSION INCREASES WITH THE SQUARE OF DISTANCE How long does it take for a given material to diffuse some distance, x? This question is deceptively simple. It is not asking how long it takes the first molecule to get to x, or how long it takes for the concentration at x to reach a particular value. It is asking about the population of molecules that are moving from high concentration to low concentration. From Figure 1.6.1, you can see that an initially sharp distribution gradually broadens with time because of diffusion. What we want is a quantitative measure of the shape of the concentration profile. The time of diffusion is usually calculated from the variance of the Gaussian distribution: ð 1N rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 2 e2x =4DΔt dx 5 2DΔt ½1:6:32 x2 5 x2 4πDΔt 2N where the elapsed time of diffusion is Δt. For twodimensional diffusion, the variance is 4DΔt; for threedimensional diffusion it is 6DΔt. The elapsed time for diffusion, Δt, is the time taken for the inflection point of the distribution to move from x 5 0 to x 5 x, given an initially sharp distribution at x 5 0. The time taken to diffuse a given distance, x, is the mean square displacement divided by 2D. Alternatively, we can calculate the distance of diffusion in time Δt as pffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi ½1:6:33 x 5 x2 5 2DΔt For distances smaller than the cell (010 μm), diffusion takes less than a ms up to a few ms. For distances on the order of the diameter of muscle cells (40100 μm), diffusion takes several seconds. DIFFUSION COEFFICIENTS IN CELLS ARE LESS THAN THE FREE DIFFUSION COEFFICIENT IN WATER If the initial concentration of a substance is very narrow, then the subsequent distribution some time later due to diffusion will be given by Eqn [1.6.20] as rffiffiffiffiffiffiffiffiffiffiffi 1 2x2 =4Dt e Cðx; tÞ 5 C0 ½1:6:34 4πDt This equation can be shown to obey Fick’s Second Law of Diffusion, Eqn [1.6.31], which is left as an exercise for the student (see Problem 15 in Problem Set 1.2). Dividing Eqn [1.6.34] by C0 and taking the logarithm of both sides, we obtain: sffiffiffiffiffiffiffiffiffiffiffi Cðx; tÞ 1 2 5 ln 1 lne2x =4Dt ln C0 4πDt ½1:6:35 pffiffiffiffiffiffiffiffi x2 2 ln ð2 πDt Þ 52 4Dt Kushmerick and Podolsky (Ionic mobility in cells, Science 166:12971298, 1969) microinjected a 36 mm segment of muscle fiber from the semitendinosus muscle of the frog with small amounts of tracer materials, and then allowed the materials to diffuse along the fibers for various periods of time. They immersed the fibers in oil to avoid diffusion through the water phase outside the muscle. After the prescribed period, the muscles were dehydrated in acetone, stained, embedded in paraffin, 79 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Example 1.6.1 Time of One-Dimensional Diffusion The diffusion coefficient for Ca21 in water is 25 2 21 DCa 5 0.8 3 10 cm s ; for calbindin, a 9500-Da protein that binds Ca21 with high affinity, DP 5 0.12 3 1025 cm2 s21. How long does it take for Ca21 or calbindin to diffuse x 5 {0.1, 1, 10, 20, 50, 100 μm}? Here we use Eqn [1.6.32] to calculate Δt 5 x2/2D. As a representative calculation, we find Δt for Ca21 for 0.1 μm: Δt 5 which is amazingly quick. The other times can be calculated by the same method, and we can fill out Table 1.6.1. Variation of this calculation was performed by A.V. Hill in 1949 to show that a diffusable substance could not be the activator of muscle cells because the delay between nervous excitation and contraction of the muscle was too fast, only a few ms. ð0:1 3 1026 m 3 102 cm m21 Þ2 5 6:25 3 1026 s 2 3 0:8 3 1025 cm2 s21 TABLE 1.6.1 The Calculated Time for One-Dimensional Diffusion of Free Ca21 and Calbindin in Water for Various Distances Distance (µm) Calcium DCa 5 0.8 3 1025 cm2 s21 0.1 6.3 3 1026 1.0 0.6 3 1023 10.0 23 62.5 3 10 Calbindin DP 5 0.12 3 1025 cm2 s21 42 3 1026 4.2 3 1023 420 3 1023 20.0 0.25 1.67 50.0 1.56 10.4 100.0 6.25 41.7 and cut into 25 μm sections, which were then counted. The plots of ln (counts at distance x/total counts) were linear with the square of the distance, as predicted by Eqn [1.6.35], and the diffusion coefficient in the muscle was estimated from the slope. An example of their results for 42K is shown in Figure 1.6.4. Kushmerick and Podolsky found that the diffusion coefficient for most substances they injected was about one-half of the free water diffusion coefficient. These materials included K1, Na1, SO422, sorbitol, sucrose, and ATP. For Ca21, however, the apparent diffusion coefficient was about 50 times less than the free water diffusion. From this they concluded that Ca21 was retarded by interaction with fixed components within the muscle cell. It is important to recognize that this effect is a reduction in the apparent diffusion coefficient because the total Ca21 is being partitioned between a freely diffusable form and a bound or fixed form. Thus the apparent diffusion coefficient is reduced because they calculated the diffusion from the total concentration and not just the free concentration. EXTERNAL FORCES CAN MOVE PARTICLES AND ALTER THE DIFFUSIVE FLUX The mathematical relations describing the diffusion of nonelectrolytes have been presented to you in the form of Fick’s First and Second Laws of Diffusion. These expressions were derived from the one- 2 Ln (42K per 0.012 cm length/total 42K) 80 4 6 8 0.03 0.06 0.09 Length2 (cm2) FIGURE 1.6.4 Typical results for the Kushmerick and Podolsky experiment. The longitudinal distribution of 42K1 at 0.02 cm intervals after diffusion for 320 s at 20 C was plotted as a logarithmic transform of the diffusion equation for an infinite slab from an infinitely thin distribution: ln (counts at distance x/ total cts) 5 2x2/4Dt 2 ln 2(πDt)1/2. dimensional random walk, which considered that the probability of making a step to the right and to the left was the same. Under some circumstances, this is not true. For example, diffusive flux is altered when a Di ffusion bulk flux of fluid occurs simultaneously with the diffusive flux. In this case, we write ½1:6:36 Js ðx; tÞ 5 2D @Cðx; tÞ 1 Jv ðtÞCðx; tÞ @x The first term on the right describes the diffusive flux and the second term describes the flux of solute due to solvent drag. Here Jv is the volume flux, equal to the volume of fluid moving across an area per unit area per unit time. This volume times the concentration of solute will give the amount of solute moving across that area per unit area per unit time. Jv is equal to the velocity of fluid flow: the volume flux is V/(AΔt) 5 (AΔx)/ (AΔt) 5 Δx/Δt, the velocity of fluid flow. Eqn [1.6.36] is the convectiondiffusion equation, because the bulk flow is described as convection. There are other circumstances which alter the flux from that described by Fick’s First Law. These circumstances occur when there are external forces applied to the solute particles. Examples of these forces include electrical forces and gravitational forces. Recall in Chapter 1.3 that we considered that electrical forces accelerate ions in solution until they reach a terminal velocity in which the electrical force is balanced by the drag force. The drag force, Fd, is proportional to the terminal velocity. We wrote: Fe 5 zeE ½1:6:37 Fd 52β v Fe 1 Fd 5 0 Fe 5 β v where β is the frictional coefficient or drag coefficient. It is given as ½1:6:38 β5 kT D as originally proposed by Einstein in 1905 (see Appendix 1.6.A1). Here k is Boltzmann’s constant, the ideal gas constant divided by Avogadro’s number: k 5 R/N0, and D is the diffusion coefficient. This result makes sense: a large diffusion coefficient is usually associated with small particles, and these would have a small frictional coefficient, encountering less resistance to movement. Incorporating this definition of β into the last equation in Eqn [1.6.37], we have ½1:6:39 Fe 5 kT v D In Chapter 1.3, we also established that the ratio of J to C defines an average velocity of movement of particles: ½1:6:40 J5v C Substituting for v from Eqn [1.6.40] into Eqn [1.6.39] and rearranging, we obtain D Fe C kT What this equation means is that, in the absence of a concentration gradient, an external force will produce a flux that is linearly related to the magnitude of the ½1:6:41 J5 force per particle and the concentration, with a coefficient related to the diffusion coefficient. We have derived this for an electric force, but the result is completely general for any external force. In the presence of a concentration gradient, we expect the diffusive flux to add to the flux caused by the application of an external force. From Fick’s First Law of Diffusion and Eqn [1.6.41], we obtain, for one dimension, Fick’s First Law of Diffusion for solutes subjected to an external force: ½1:6:42 J 5 2D @C D 1 fC @x kT This can be written in vector notation as ½1:6:43 J 5 2DrC 1 D CF kT THE STOKESEINSTEIN EQUATION RELATES THE DIFFUSION COEFFICIENT TO MOLECULAR SIZE Stokes showed that for a spherical particle, the drag force was related to its size and to the viscosity of the medium: ½1:6:44 Fd 5 2β v 5 2 6πηas v where η is the viscosity of the fluid and as is the radius of the sphere. Clearly, Stokes derived an expression for the frictional coefficient. From Eqns [1.6.38] and [1.6.44], we can write 6πηas 5 ½1:6:45 D5 kT D kT 6πηas The last expression is the StokesEinstein equation. It indicates that the diffusion coefficient for a spherical particle should be a function of its radius, the absolute temperature, and the viscosity of the fluid in which it is diffusing. Recall earlier that Kushmerick and Podolsky found that the apparent diffusion coefficient of readily diffusable substances like K1, Na1, and sucrose inside cells was about one-half of their diffusion coefficients in water. From the StokesEinstein equation, it would seem that the variable most likely responsible for these decreases in diffusion coefficients is the viscosity of the medium. Thus the cytoplasm appears to be a watery environment, where most small molecular weight materials are free to diffuse but with reduced diffusion coefficients owing to the greater viscosity of the cytoplasm. The tortuosity of the path is not included in this analysis. This refers to the blockade of direct diffusion by large structures in the cytoplasm, including organelles and cytoskeleton. Because materials cannot diffuse in a straight line, the diffusion apparently takes longer because the actual path length in the microscopic domain is larger than the apparent path length. In fact, tortuosity and 81 82 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION increased viscosity are both partial explanations for the reduced diffusion coefficient inside cells. expression for the drag force in terms of the diffusion constant and gave us the StokesEinstein equation: D5 SUMMARY Solutes move by diffusion from regions of high concentration to regions of low concentration. Fick’s First Law of Diffusion states that the flux is proportional to the negative of the gradient of C: D @C Js 5 2 @x where Js is the solute flux, D is the diffusion coefficient, and @C/@x is the one-dimensional gradient. The continuity equation states that changes in concentration with time must be due to changes in flux with distance: @C @Js 52 @t @x Fick’s Second Law of Diffusion derives from his First Law and the Continuity Equation: @C D @2 C 5 @t @x2 Fick’s Second Law can be derived from a random walk model of diffusion in which molecules take large numbers of small steps. Using Stirling’s approximation, the discrete binomial probability distribution can be converted to a continuous one, resulting in a Gaussian probability distribution. For a narrow starting distribution at time t 5 0, the distribution of solute at time t is given as pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 Cðx; tÞ 5 C0 1=4πDt e2x =4Dt From the random walk, D is identified as λ2/2tc, where λ is the distance between collisions and tc is the time between collisions. The time of diffusion is typically estimated as the variance of the Gaussian distribution, which gives x2 2D Although diffusion is a statistical result, it is equivalent to a force in that it produces a flow of material. Other forces can also make solutes move. These forces include electrical forces on charged solutes, solvent drag (convection), and gravitational forces. An external force applied to solute particles causes a flux given by t5 J5 D fC kT where f is the force per molecule, k is Boltzmann’s constant (k 5 R/N0), T is the absolute temperature, and C is the concentration. In the presence of a concentration gradient, the total diffusive flux in the presence of an external force is J52 D@C DfC 1 @x kT Stokes derived an equation for the drag force on a spherical object. Einstein combined this with his kT 6πηas where η is the viscosity of the medium in which diffusion occurs and as is the radius of a spherical solute. REVIEW QUESTIONS 1. Why is diffusive flux proportional to the negative of the gradient and not the gradient? 2. What are the assumptions of the random walk derivation? 3. What are the units of the diffusion coefficient? 4. How does the time of diffusion vary with distance? 5. Why are diffusion coefficients slower in the cytoplasm of cells than in water? 6. What is solvent drag? 7. What is meant by convective flow? 8. How does the diffusion coefficient vary with molecular size? With viscosity of the medium? With temperature? What is Boltzmann’s constant? 9. How is Fick’s First Law of Diffusion altered in the presence of additional forces acting on the diffusing particles? APPENDIX 1.6.A1 DERIVATION OF EINSTEIN’S FRICTIONAL COEFFICIENT FROM MOMENTUM TRANSFER IN SOLUTION Here we consider a right cylindrical volume V of crosssectional area A and thickness Δx, so that V 5 AΔx. We imagine that solute particles may move with velocity 1 v in the x-direction (to the right) and velocity 2 v (velocity v to the left), as shown in Figure 1.6.A1.1. Let the number of particles in V with velocity v be N1(t) and the number with velocity 2 v be N2(t). The concentration of particles with velocity 1 v or 2 v in V at any time will be given by ½1:6:A1:1 N1 ðtÞ 5 C1 ðtÞ V N2 ðtÞ 5 C2ðtÞ V The number of particles with a given velocity may change with time. This may happen in four different ways: (1) particles with velocity 1 v may enter the volume element from the left; (2) particles with velocity 1 v may leave the volume element at the right; (3) particles with velocity 1 v within the volume V may convert to particles with velocity 2 v by colliding with solvent particles; (4) particles with velocity 2 v could convert to velocity 1 v by collisions with solvent particles. The entry of particles from the left is given by ½1:6:A1:2 Qi ðxÞ 5 v A C1 ðxÞ Di ffusion Cross-sectional area, A J(x) +v +v J(x +Δx) –v –v x x +Δx x axis FIGURE 1.6.A1.1 Solutes within a hypothetical volume. Solutes have a velocity 1 v in the positive x-direction or 2 v in the opposite direction. All particles have velocity 1 v or 2 v, although only a few are shown. where we recognize that vC1 is the flux of particles with velocity 1 v. The exit of particles at the right is ½1:6:A1:3 Qo ðxÞ 5 v A C1 ðx 1 ΔxÞ The rate of conversion of N1 to N2 is proportional to N1 within the volume V, with the proportionality constant having dimensions of reciprocal time. This proportionality constant is 1/tc, where tc is the time between collisions. This conversion reduces N1 within V, so we may write: @N1 N1 V C1 ½1:6:A1:4 52 5 @t -collision tc tc The rate of conversion of N2 to N1 adds to the value of N1 and is given by a similar expression: @N1 N2 V C2 ½1:6:A1:5 5 5 @t 1collision tc tc The total net change of N1 is given by the sum of Eqns [1.6.A1.2, 1.6.A1.3, 1.6.A1.4, 1.6.A1.5]: @N1 V 5 vA C1 ðxÞ 2 vA C1 ðx 1 ΔxÞ 1 ðC2 2 C1 Þ tc @t ½1:6:A1:6 where v means velocity and V means volume. We can approximate C1(x 1 Δx) by the first two terms of a Taylor’s series expansion: @C1 C1 ðx 1 ΔxÞ 5 C1 ðxÞ 1 Δx ½1:6:A1:7 1? @C Insertion of Eqn [1.6.A1.7] into Eqn [1.6.A1.6] gives " # @N1 @CðxÞ V 1 ðC2 2C1Þ 5vA C1ðxÞ2C1ðxÞ2Δx @x tc @t ! @N1 @C1ðxÞ V 2 ðC1 2C2Þ 52vAΔx @x tc @t ½1:6:A1:8 Since AΔx 5 V, we write ½1:6:A1:9 1 @N1 @C1 1 5 2v 2 ðC1 2 C2 Þ V @t tc @x @C1 @C1 1 5 2v 2 ðC1 2 C2 Þ tc @t @x By completely analogous reasoning, we can determine the rate of change in C2 in V as ½1:6:A1:10 @C2 @C2 1 5v 1 ðC1 2 C2 Þ tc @t @x Since the total concentration, C(x) is the sum of C1 and C2, then ½1:6:A1:11 @CðxÞ @t 5 @C1 ðxÞ @C2 ðxÞ 1 @t @t and substituting into Eqn [1.6.A1.11] from Eqns [1.6.A1.9] and [1.6.A1.10], we obtain @CðxÞ @C1 1 @C1 1 5 2v 2 ðC1 2 C2 Þ 1 v 1 ðC1 2 C2 Þ @t tc tc @x @x ! @C1 @C2 5 2v 2 @x @x 52 @ðvðC1 2 C2 ÞÞ @x ½1:6:A1:12 Now the net flux across any area element within the volume V can be described as the difference between two unidirectional fluxes: ½1:6:A1:13 J 5 J1 2 J2 5 v C1 2 v C2 5 v ðC1 2 C2 Þ Substitution of this relation into Eqn [1.6.A1.12] gives ½1:6:A1:14 @CðxÞ @JðxÞ 52 @t @x 83 84 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION which is the continuity equation we derived earlier (see Eqn [1.2.9]). The purpose of deriving the continuity equation in this way was to familiarize you with this method of accounting for all of the particles and to impress upon you that the collisions with the solvent involving changes in the velocity canceled. These collisions did not affect the total number of particles since a particle contributing to N1 before collision still contributes to N2 after collision. Now let us consider what happens to the total momentum, P, of the particles in volume V. Let m be the mass of each solute particle. The total momentum of the particles is ½1:6:A1:15 P 5 ðmvÞN1 1 ð2mvÞN2 Because N1 and N2 change with time, so does P. We write @P @N1 @N2 5 mv ½1:6:A1:16 2 mv @t @t @t Since N1 5 VC1 and N2 5 VC2, we have ½1:6:A1:17 @P @C1 @C2 5 mvV 2 mvV @t @t @t Inserting our earlier results from Eqns [1.6.A1.9] and [1.6.A1.10], we obtain 1 @P @C1 @C2 2mv 5 2mv2 1 ðC1 2 C2 Þ 2 V @t tc @x @x ½1:6:A1:18 Recalling that C 5 C1 1 C2, and using Eqns [1.6.A1.11] and [1.6.A1.13], we obtain ½1:6:A1:19 1 @P @C 2m 5 2 mv2 2 J V @t @x tc The two terms on the right-hand side of Eqn [1.6.A1.19] have specific interpretations. The first term, 2 mv2@C/@x V, represents the net flow of momentum carried by particles with velocity 1 v or 2 v. The second term, 22m/tc JV, is the average force exerted by the solvent particles on the solute particles. To see these interpretations, we write the flow of momentum across area A at any point x within the volume V as mv ðvC1ðxÞÞA 1 ð2 mvÞð2 vC2ðxÞÞA 5 mv2 ðC1ðxÞ 1 C2ðxÞÞA ½1:6:A1:20 The net flow of momentum into volume V across the boundaries between x and x 1 Δx is mv2 ðC1 ðxÞ 1 C2 ðxÞÞA 2 mv2 ðC1 ðx 1 ΔxÞ 1 C2 ðx 1 ΔxÞÞA ½1:6:A1:21 Since C(x) 5 C1(x) 1 C2(x), the expression in Eqn [1.6. A1.21] becomes ½1:6:A1:22 mv2 A½CðxÞ 2 Cðx 1 ΔxÞ Expanding C(x 1 Δx) as C(x) 1 Δx @C/@x, the expression in Eqn [1.6.A1.22] becomes ½1:6:A1:23 2mv2 @C @C ΔxA 5 2 mv2 V @x @x Thus, in Eqn [1.6.A1.19], 2 mv2 @C/@x is that part of 1/V @P/@t which is due to the net momentum change of the particles in V due to what they carried into or out of the volume. For the second term in Eqn [1.6.A1.19], note that every time a particle moving with velocity 1 v collides with a solvent obstacle, its velocity becomes 2 v, and thus it experiences a net momentum change of 22mv. The total number of such changes per unit time in the volume V is N1/tc. So the time-averaged change for the momentum change per unit time due to this collision is 22mv N1/tc. In a similar way, particles traveling with velocity 2 v experience a momentum change of 12mv upon collision with the solvent and the time-averaged total change in momentum per unit time for this type of collision in the volume V is 12mv N2/V. The time-averaged rate of net change in momentum for these types of collisions is just their sum: @Pc N1 N2 5 Fc 5 2 2mv 1 2mv @t tc tc ½1:6:A1:24 5 Fc 5 2 2mv VðC1 2 C2 Þ tc 1 @Pc Fc 2m 5 52 J V @t tc V where the subscript c denotes that the momentum change and force are due to collisions of the solute molecules with the solvent. Thus the second term in Eqn [1.6.A1.19] is identified as the change in momentum produced by collisions with the solvent. It is equivalent to a force per unit volume exerted by the solvent particles on the solute particles. Equation [1.6.A1.24] can be rewritten as a differential equation in J(x,t). Recall that the total momentum, P, of the solute particles in the volume V is given by Eqn [1.6.A1.15]; substituting in for the definition of the concentration (Eqn [1.6.A1.1] and (J) Eqn [1.6.A1.13]), we have: P 5 ðmvÞN1 1 ð2 mvÞN2 ½1:6:A1:25 5 mV vðC1 2 C2 Þ P 5 mV Jðx; tÞ Insertion of this result into Eqn [1.6.A1.19] gives ½1:6:A1:26 1 @Jðx; tÞ @Cðx; tÞ 2m mV 52mv2 2 Jðx; tÞ V @t @x tc This can be rearranged to ½1:6:A1:27 @Jðx; tÞ 2 tc v2 @Cðx; tÞ 5 2 Jðx; tÞ 2 @t tc @x 2 Eqn [1.6.A1.27] describes the buildup of J(x,t) in time. We will pay particular attention to the situation Di ffusion where the concentration gradient, @C(x,t)/@x, is constant and steady-state flux is achieved. Steady-state flux means that @J(x,t)/@t 5 0. That is, the flux no longer changes with time. Under these circumstances, by Eqn [1.6. A1.27], we obtain the steady-state flux as tc v2 @C ½1:6:A1:28 J52 2 @x Comparing this to Fick’s First Law of Diffusion, ½1:6:A1:29 @C J 5 2D @x we can identify tc v2 2 In the one-dimensional random walk model of diffusion, we defined the diffusion coefficient to be ½1:6:A1:30 D5 D5 ½1:6:A1:31 5 5 λ2 2tc ðvtc Þ2 2tc 2 tc v 2 Thus the derivation performed here is completely consistent with the random walk model of diffusion. Eqn [1.6.A1.24] describes the momentum change of the solute particles that result from the collisions with solvent. It is given per unit volume as ½1:6:A1:32 1 @Pc Fc 2m 5 52 J V @t tc V In this equation, Fc is the force per unit volume on the solute particles. This is the same as the drag force on the solute particles when they move through the solution. If we imagine that the particles are subjected to a uniform external force, then the particles will accelerate until they reach a terminal velocity, v. The force FC will be the sum of all of the drag forces on the particles within the volume, which is the drag force per particle times the number of particles in the volume. The flux, J, will be given as vC. Eqn [1.6.A1.32] can then be rewritten as ½1:6:A1:33 fc N 2m 52 vC V tc where fC is the force on a single particle, N is the number of particles in the volume, V, v is the terminal velocity, and C is the concentration. Since N/V 5 C, the equation is further simplified to ½1:6:A1:34 fc 5 2 2m v tc Here the force fC is equal to the drag force on the particle traveling at velocity v, and the coefficient is the drag or frictional coefficient: ½1:6:A1:35 FD 5 2β v which allows us to identify the drag coefficient as ½1:6:A1:36 2m tc β5 The equipartition theorem of thermodynamics gives 1 2 1 mv 5 kT 2 2 ½1:6:A1:37 v2 5 kT m where k is Boltzmann’s constant (51.38 310223 J K21, which is equal to the gas constant, R, divided by Avogadro’s number). Insertion of this result into the equation for the diffusion coefficient, Eqn [1.6.A1.31] gives D5 tc kT 2m tc 5 2mD kT ½1:6:A1:38 Insertion of this into Eqn [1.6.A1.36] gives ½1:6:A1:39 β5 2m 2m kT 5 5 2mD tc D kT This is Einstein’s frictional coefficient, given as ½1:6:A1:39 β5 kT D Recall Eqn [1.6.A1.19] reproduced here: ½1:6:A1:19 1 @P @C 2m 5 2 mv2 2 J V @t @x tc substituting in v2 5 kT/m from Eqn [1.6.A1.37] and tc 5 2mD/kT from Eqn [1.6.A1.38], we obtain ½1:6:A1:40 1 @P @C kT 5 2 kT 2 J V @t @x D Recall that this equation describes the rate of momentum change of the solute per unit volume V. It has two components: the first is the net momentum carried into the volume by the diffusing solutes, and the second is the momentum change produced upon collisions with the solvent. Note that if C(x) is decreasing with increasing x then the gradient, @C/@x, will be negative and the first term on the right-hand side of Eqn [1.6. A1.40] will be positive. In this case of diffusion of solute towards increasing values of x, J will be positive and the second term, denoting the change in momentum by collisions with solvent, will be negative. This second term denotes the drag of solvent on solute movement. Steady-state diffusion occurs when @P/@t is zero and @C/@x is constant. Under these conditions, Eqn [1.6.A1.40] becomes Fick’s First Law of Diffusion. Suppose now that we add an additional force to the solute particles in the volume V. Let the force acting on each particle be f. In the volume element, there are 85 86 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION N(x) 5 C(x) V particles. The total force acting on these particles is f C(x) V. This force clearly contributes to the rate of change of the momentum of the solute particles in the volume. Eqn [1.6.A1.40] becomes ½1:6:A1:41 1 @P @C kT 52 kT 2 J 1 fC V @t @x D To recapitulate, the first term on the right-hand side of Eqn [1.6.A1.41] represents the rate of momentum change by particles entering or leaving the volume V; the second term is due to collisions with the solvent particles; the third term is due to the action of some external force. At steady state, @J/@t 5 0, and this implies by Eqn [1.6.A1.25] that @P/@t also is zero. Under this constraint, Eqn [1.6.A1.41] gives ½1:6:A1:42 J52D @C D 1 fC @x kT This is Fick’s First Law of Diffusion for solutes subjected to an external force, which is Eqn [1.6.42]. Thus this equation, which forms the basis of the derivation of the electrochemical potential, can be derived using momentum transfer of solutes in a solution and the equipartition theorem of thermodynamics. Electrochemical Potential and Free Energy derivative, we retain the three-dimensional vector that describes diffusive flux in three dimensions. Learning Objectives G G G G G G G G G G Write Fick’s First Law of Diffusion and explain how a concentration gradient makes a flux Describe how an external force such as electrostatic force can make a flux of charged solute Write the formula for the electrochemical potential Explain the formula for electrochemical potential in terms of the component driving forces Explain how the driving forces that produce a flux equal the negative gradient of the electrochemical potential Define the term free energy Describe the relationship between the free energy and the direction of any process Be able to calculate the free energy of ATP hydrolysis under specified conditions of temperature and concentrations of reactants ATP, ADP, and Pi Write the relationship between the standard free energy change and the equilibrium constant Know the approximate value for the free energy of ATP hydrolysis under cellular conditions Third, in the absence of a concentration gradient, any external force acting on solute particles causes them to be accelerated until they reach a terminal velocity. At this terminal velocity, the external force is balanced by the drag force on the particle by the solvent. This drag force is proportional to the velocity. Because of this, there is a relationship between the flux and the external force, given as ½1:7:3 J5 DIFFUSIVE AND ELECTRICAL FORCES CAN BE UNIFIED IN THE ELECTROCHEMICAL POTENTIAL Let us recapitulate what we learned in Chapters 1.3 and 1.6 about the movement of charged particles in solution. First, the relationship between the flux and the concentration defines an average velocity, given as ½1:7:4 J 5 2D @C @x where D is the diffusion coefficient. Here we drop the vector notation because the equation used here is one dimensional and the direction is assumed to be along the x-axis. If we use the gradient of C, instead of the ½1:7:2 J 5 2D © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00008-2 @C D 1 Cf @x kT If the force is an electrical force, its magnitude is given by F 5 FE 5 zeE where z is the valence of the particle ( 6 integer), e is the charge on the electron, and E is the electric field. Assuming one dimension, we drop the vector notation for force and insert Eqn [1.7.5] into Eqn [1.7.4] to give ½1:7:6 J 5 2D @C D 1 CzeE @x kT Thus we have these two equations: J 5 vC where J is the one-dimensional flux, v is the velocity, and C is the concentration. Second, the presence of a concentration gradient, in the absence of any other forces, produces a flux given by Fick’s First Law of Diffusion: D Cf kT where f is the force acting on the solute particles, per particle. The concentration gradient has some of the appearances of a force in that it causes a flux. So does an externally applied force. What we seek to do is to combine diffusion and other forces into a single equivalent force. When both are operating, the flux is given as ½1:7:5 ½1:7:1 1.7 D Cf kT @C D 1 CzeE J 5 2D @x kT J5 ½1:7:7 The top equation says that there is a flux produced by some unknown force. The bottom equation says that now we know what these forces are and we have parceled them out. Part of the flux is caused by diffusion and part of the flux is caused by an electrical force. What we desire is an expression for the total force in the top equation that will produce the flux in the bottom equation. To unite the diffusive force and 87 88 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION the electrical force, we solve these two equations for the unknown force, f, and find that ½1:7:8 f 52 kT @C 1 zeE C @x The electric field, E, is the electric force per unit charge. This electric force is the negative gradient of its potential (see Chapter 1.3 for a definition of potential, gradients, and conservative forces). In a three-dimensional model, the electric field is a vector and the potential is a scalar. We will ignore these realities in this one-dimensional model because in one dimension the force and gradient have a single direction. Using Ψ as the symbol for electrical potential energy, we can rewrite Eqn [1.7.8] as ½1:7:9 f 52 kT @C @Ψ 2 ze C @x @x N0 kT @C @Ψ 2 z N0 e C @x @x RT @C @Ψ F5 2 2 z` C @x @x F 5 N0 f 5 2 21 where R is the gas constant (58.314 J mol K ) and ` is the Faraday (9.649 3 104 C mol21 5 6.02 3 1023 electrons mol2131.6310219 C electron21). We define an electrochemical potential so that the overall force (produced by both diffusion and electrical forces, in this case) is the negative derivative of the electrochemical potential: F52 @µ @x Combining Eqns [1.7.10] and [1.7.11], we get ½1:7:12 2 @µ RT @C @Ψ 52 2 z` @x C @x @x Integrating, we obtain ½1:7:13 µi 5 µ0i 1 RT ln Ci 1 zi `Ψ THE OVERALL FORCE THAT DRIVES FLUX IS THE NEGATIVE GRADIENT OF THE ELECTROCHEMICAL POTENTIAL The flux produced by a concentration difference and electric potential difference is given by 21 ½1:7:11 In a typical physiological solution, several kinds of particles are dissolved. Since their concentrations are nearly independent of one another, they each have an identifiable electrochemical potential. Each solute within a local region, on the other hand, experiences the same electrical potential. Therefore, we can write an electrochemical potential for each material in the solution: ½1:7:14 The units of Ψ are volts. The force in this equation is the force per particle. It is customary to convert this to the force per mole of particles by multiplying by Avogadro’s number: ½1:7:10 by Eqn [1.7.13]. The second term, RT ln C, refers to the work necessary to concentrate the solute, per mole; the third term, z`Ψ, is an electrical work term. It is the work necessary to bring one mole of particles from zero potential to Ψ. If other kinds of work are involved, we would need to expand Eqn [1.7.13] to include the other work terms. µ 5 µ0 1 RT ln C 1 z`Ψ This is the electrochemical potential. We will use it to make calculations about membrane potential and the energetics involved in physiological processes. It is a potential in the same sense as the electrical potential. It has the units of energy per mole. The electrical potential, in volts, is equivalent to joules per coulomb, which can be converted to units of energy per mole. Similarly, the electrochemical potential can be converted into volts. As can be seen from Eqn [1.7.13], the electrochemical potential has three components. The first component, µ0, has the sense of a constant of integration. It sets the zero of the electrochemical potential and its reference is a standard state. For solutions, µ0 refers to the chemical potential of a hypothetical solution of unit molarity and no potential. That is, when C 5 1 and Ψ 5 0, µ 5 µ0 ½1:7:3 J5 D Cf kT where J and f are both vectors. f is the force per molecule. We convert to molar dimensions by multiplying and dividing by Avogadro’s number: D CN0 f N0 kT D CF J5 RT J5 ½1:7:15 This emphasizes that J is a vector in the same direction as F, the force on the solute particles per mole. The force per mole is, in turn, given by the negative gradient of the electrochemical potential: @µ @µ @µ 1j 1k ½1:7:16 F 5 2rµ 5 i @x @y @z Inserting in µ from Eqn [1.7.13] and F into Eqn [1.7.15], we get D RT @C @C @C C 2 i 1j 1k J5 RT C @x @y @z @Ψ @Ψ @Ψ 1j 1k ½1:7:17 2 z` i @x @y @z which is simplified to ½1:7:18 z` CrΨ J 5 2D rC 1 RT The one-dimensional version of this equation is ½1:7:19 J 5 2D @C D @Ψ 2 z`C @x RT @x Electr ochemical P otentia l and Fre e E nergy THE ELECTROCHEMICAL POTENTIAL IS THE GIBBS FREE ENERGY PER MOLE Throughout the derivation of Fick’s Laws of Diffusion, we made an unstated assumption that the pressure and temperature of the diffusing particles were constant. Under these conditions, we could derive the electrochemical potential. The constraints of constant temperature and pressure are useful when applied to problems in mammalian physiology, where these conditions are usually met. Under these conditions, it is useful to define a thermodynamic variable called the Gibbs free energy: ½1:7:20 G 5 E 1 PV 2 TS where E is the internal energy, P is the pressure, V is the volume, T is the temperature, and S is the entropy. The internal energy consists of all of the myriads of movements of the particles, including their internal motions. This is the part of the free energy where chemical bonding energy is stored. When compounds undergo chemical transformations, energy is either released or stored, depending on whether the reaction is exothermic or endothermic, respectively. The entropy has a specific thermodynamic definition which has been shown to be related to the number of ways that the particles can be arranged in the system which are consistent with the state of the system (its temperature, pressure, volume, and number of particles). This is related to the probability of finding the state for randomly arranged particles. Thus states with high probabilities have high entropy. Highly organized states, which can be accomplished in only a few ways, have a low probability and a low entropy. For example, a concentrated solution has only a few ways to crowd all the solute particles together compared to a dilute solution, so a concentrated solution has less entropy than a dilute solution. As mentioned above, the Gibbs free energy is used to describe transformations that occur at constant temperature and under constant pressure. It is a state variable, meaning that it depends only on the state of the system and not on the path required to get there. Thus there is a defined difference in the free energy between an initial state and a final state: ½1:7:21 ΔG 5 Gfinal 2 Ginitial The energy involved in a chemical transformation can be harnessed to do useful work. For example, a reaction that gives off heat can be used to expand a gas to move a piston, producing work. The relationship between the free energy and the work that can be accomplished by a chemical transformation is ½1:7:22 2ΔG $ W where W is the work energy. That is the decrease in free energy in any transformation at constant T and P is equal to the maximum amount of work that can be performed by that transformation. It is for this reason that the function G is called the “free energy”; it is the energy which is available to perform work. The work that can be performed could be mechanical or electrical or chemical or concentration work. Another variable determining the state of a system is its composition. Suppose that the system is composed of a number of substances, i, j, k, . . . of mole amounts ni, nj, nk, . . . If a small amount of substance i is added, then the system will experience a change in its free energy, G. The chemical potential is @G ½1:7:23 µi 5 @ni T;P;nj1i Thus the chemical potential is the free energy per mole. The Gibbs free energy is an extensive variable that increases with the volume of the system or the number of moles of materials in the system. The chemical potential is an intensive variable, meaning that it depends on the state of the system and not its extent. However, usually ΔG values are calculated per mole so that G and µ are often used interchangeably. Equation [1.7.23] can be integrated to give X ½1:7:24 G5 µi ni i THE SIGN OF ΔG DETERMINES THE DIRECTION OF A REACTION One of the conclusions of thermodynamics is that natural processes always occur in such a way that the free energy of the universe decreases. That is, in all natural processes, ΔG in Eqn [1.7.21] is negative. A corollary of this conclusion is the condition of equilibrium. Equilibrium occurs when no further change can occur. In this case, the free energy has reached a minimum, and the free energy change is zero with respect to this process. On the other hand, if ΔG is positive, then the reverse reaction will occur spontaneously: ΔG , 0.spontaneous reaction ΔG 5 0.reaction is at equilibrium ΔG . 0.opposite reaction occurs spontaneously ½1:7:25 PROCESSES WITH ΔG . 0 CAN PROCEED ONLY BY LINKING THEM WITH ANOTHER PROCESS WITH ΔG , 0 As described earlier, a natural process can proceed only if ΔG , 0 for that process. Many processes that occur in biological systems require energy, meaning that ΔG . 0. These processes cannot proceed on their own. They can be made to proceed, however, by linking the process for which ΔG . 0 with another process for which ΔG , 0. The combined processes will proceed spontaneously only if the global ΔG for both of them is less than zero. We will consider these kinds of processes in detail when we consider active transport in Chapter 2.6. “Proceed spontaneously” here means that the linked processes will occur with no further input of energy outside the combined processes. It does not mean that the processes will occur rapidly or slowly. Thermodynamics tells us whether or not a process will occur, and with what energy changes, but it does not speak of the kinetics of the processes. In this sense, the term “thermodynamics” 89 90 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Wcounter currency. Energy derived from the oxidation of foodstuffs is stored in this terminal phosphate bond, and it is used in myriad reactions involving synthesis of materials, transport of materials, and the performance of mechanical work. The hydrolysis reaction of ATP to ADP and Pi is shown in Figure 1.7.3. The free energy change for the ATP hydrolysis reaction is given by a combination of Eqns [1.7.21] and [1.7.24]: ΔG 5 Gfinal 2 Ginitial nfinal ATP µATP 1 nfinal ADP µADP 1 nfinal Pi µPi 5 2ninitial ATP µATP 2ninitial ADP µADP 2ninitial Pi µPi Welevator 5 ΔnATP µATP 1 ΔnADP µADP 1 ΔnPi µPi 5 Δnrxn µADP 1 Δnrxn µPi 2 Δnrxn µATP ½1:7:26 FIGURE 1.7.1 Coupling of a spontaneous process to a nonspontaneous process. An elevator sits on the ground floor. It will not rise spontaneously because the free energy for that process is positive: ΔGelevator . 0. A large counterweight at the top floor will fall spontaneously; the free energy change for this process is negative: ΔGcounter , 0. The elevator can be raised if it is coupled to the larger negative free energy of the counterweight’s fall: ΔGelevator 1 ΔGcounter , 0. In this case, the elevator can be raised only once. Repetition of the process would require coupling to yet another process that provides the free energy to raise the counterweight back up again. In a real process, some of the free energy for the counterweight’s fall would be dissipated. One can never recover 100% of the energy as useful work. is something of a misnomer, as really it is about the statics of energy transformation. An example of coupling is lifting a weight against gravity by using another weight, as shown in Figure 1.7.1. The coupling of biological process can be visualized using the energy diagrams for chemical reactions such as those shown in Chapter 1.5. Reactions that involve a decrease in free energy (Gfinal 2 Ginitial , 0 and thus Gfinal , Ginitial) are called exergonic reactions, and they proceed spontaneously. A schematic example of an exergonic reaction is shown in Figure 1.7.2A. Reactions that involve an increase in free energy (Gfinal 2 Ginitial . 0 and thus Gfinal . Ginitial) are called endergonic reactions, and they do not proceed spontaneously (see Figure 1.7.2B). If an exergonic reaction can be coupled to an endergonic reaction, and the sum of the ΔG for the two reactions is less than zero, then the combined processes will both proceed spontaneously (see Figure 1.7.2C). THE LARGE NEGATIVE FREE ENERGY OF ATP HYDROLYSIS POWERS MANY BIOLOGICAL PROCESSES ATP is adenosine triphosphate and its structure is shown in Figure 1.7.3. It occupies a special position in the cellular flow of energy because of the energy stored in its terminal phosphate bond. It takes a lot of energy to add a phosphate to ADP to form ATP, and that chemical energy becomes available to do work when the bond is split. The cell uses ATP as its energy where the subscripts denote the chemical species, µ is the chemical potential, and Δn is the number of moles of each material that participates in the reaction. The last line in Eqn [1.7.26] relates the change in the number of moles of participating reactants to the number of completed reactions. This is just the stoichiometry of the reaction. We may write it as follows: ½1:7:27 1 ATP 5 1 ADP 1 1 Pi 2 1 ATP 1 1 ADP 1 1 Pi 5 0 The coefficients here are the stoichiometry. Here they indicate that, for every completed reaction, the number of molecules of ATP decreases by one and the number of molecules of ADP and Pi increases by one. Thus the last line in Eqn [1.7.26] relates the change in free energy upon completion of a Δn number of reactions. This number can be specified in moles because, as described in Chapter 1.5, using moles is just another way of counting a large number of things. We divide the last line of Eqn [1.7.26] by the number of completed reactions to obtain ½1:7:28 ΔG 5 µADP 1 µPi 2 µATP Δnrxn Substituting in for the chemical potentials, we get ΔG ½1:7:29a Δn ½1:7:29b 5 µ0ADP 1 RT ln½ADP 1 µ0Pi 1 RT ln½Pi 2 µ0ATP 2 RT ln½ATP ΔG ½ADP½Pi 5 Δµ0 1 RT ln Δn ½ATP If we will let Δn 5 1.0 to indicate that we are speaking of the free energy change for Avogadro’s number of completed reactions, then we have ½ADP½Pi 0 ½1:7:30 ΔG 5 ΔG 1 RT ln ½ATP where ΔG0 is the standard free energy change. It refers to the free energy change per mole under standard Gibbs free energy (kJ mol–1) (A) (B) Exergonic process ΔGA < 0 process is spontaneous Endergonic process ΔGB > 0 process is not spontaneous Activated complex, A* Final state, D Initial state, B Initial state, A Final state, C ΔGB > 0 ΔGA < 0 Progress of the reaction Sum of A and B (C) Coupling of an exergonic process with an endergonic process allows both to occur. The free energy change of the exergonic process drives the endergonic process. A+B Gibbs free energy (kJ mol–1) Progress of the reaction C+D ΔGA+B < 0 Progress of the reaction FIGURE 1.7.2 Coupling of an endergonic reaction with an exergonic reaction. Exergonic reactions are those for which ΔGA 5 Gfinal 2 Ginitial , 0, and so these reactions proceed spontaneously (panel A). Endergonic reactions are those for which ΔGB 5 Gfinal 2 Ginitial . 0, and these reactions do not proceed spontaneously (panel B). However, an endergonic reaction can be made to proceed if it can obtain a decrease in free energy by linking it to the exergonic reaction. In essence, a coupled reaction involves a completion of both the exergonic and endergonic reactions. If the combined ΔGA1B 5 ΔGA 1 ΔGB , 0, then the combined reaction will occur spontaneously (panel C). NH2 O– NH2 O– O– –O P O P O O O P O O N N O– –O P O N N O– CH2 O O P O O N N H2O OH OH CH2 O N N Adenosine diphosphate, ADP OH OH Adenosine triphosphate, ATP O– + –O P OH O Phosphate FIGURE 1.7.3 ATP and its hydrolysis to ADP and Pi. Chemical energy is stored in each of the phosphate bonds of ATP. The last one, the γ-phosphate, is typically used to power mechanical, electrical, and chemical energy needs of the cell. 92 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION EXAMPLE 1.7.1 Free Energy of ATP Hydrolysis Under “Typical” Cell Conditions “Typical” values for the concentrations of ATP, ADP, and Pi are about 5 mM for ATP, 5 mM for Pi, and 40 μM for ADP. Calculate the free energy of ATP hydrolysis under these conditions. 0 Here we use Eqn [1.7.33]. The value of ΔG0 is given in the text as 27.4 kcal mol21 or 230.9 kJ mol21. The value of R we use is 8.314 J mol21 K21 and T 5 37 C 5 310 K. In the calculation, all concentrations must be expressed in M. We insert these values and calculate: ΔG 5 231:0 kJ mol21 1 8:314 J mol21 K21 3 310 K 5 231:0 kJ mol21 1 2:58 kJ mol21 3 ð210:12Þ 5 257:1 kJ mol21 5 213:6 kcal mol21 This free energy of ATP hydrolysis under cell conditions is the energy available for the various kinds of work undertaken by the cell, including chemical work (synthesis), mechanical work (movement and transport), and electrical work. In the final analysis, nearly all of the work produced by the cells is eventually degraded and appears as heat. 3 ln½40 3 1026 M 3 5 3 1023 M=5 3 1023 M conditions of 25 C, 1 atmosphere pressure and unit concentration. Note that we have said all along that ΔG is an extensive variable, and now it seems that we have transformed it into an intensive variable. Tabulated values of ΔG0 necessarily report it in units of energy per mole, so these values are actually values of Δµ0. You can see that if all species were at unit concentration, the second term on the right-hand side of Eqn [1.7.25] would be zero. Under these conditions, ΔG5ΔG0. Experimental determination of ΔG0 takes advantage of the fact that at equilibrium ΔG 5 0 (see Eqn [1.7.25]) and the fact that the second term in Eqn [1.7.30] incorporates the equilibrium constant for the reaction: ½ADP½Pi Keq 5 ½1:7:31 ½ATP So that at equilibrium (ΔG 5 0), Eqn [1.7.30] becomes ½1:7:32 ΔG0 5 2 RT ln Keq MEASUREMENT OF THE EQUILIBRIUM CONCENTRATIONS OF ADP, ATP, AND PI ALLOWS US TO CALCULATE ΔG0 It turns out that the free energy of ATP hydrolysis is a bit more complicated than we have let on here. The chemical species produced have different ionizations at different pH values, and ATP and ADP both bind Mg21 ions. The energy of binding of H1 and Mg21 ions should be incorporated into the reaction. In addition, we are interested in the hydrolysis of ATP under physiological conditions of 37 C. The free energy of ATP hydrolysis has been determined for these different conditions. Since the details of these conditions differ from cell to cell, no one value can be used. However, a 0 “typical” value is given a special symbol, ΔG0 , and signifies the free energy of ATP hydrolysis under the “typical” cell conditions. Its value is about 27.4 kcal mol21 or 231.0 kJ mol21. The units of ΔG are those of RT. Values of R usually used here are 1.987 cal mol21 K21 5 8.314 J mol21 K21. The conversion 1 J 50.239 cal. between joule and calorie is Under cellular conditions, [ATP], [ADP], and [Pi] are not at their equilibrium values, nor are they unit concentrations. The free energy under these conditions is given by ½ADP½Pi 0 ½1:7:33 ΔG 5 ΔG0 1 RT ln ½ATP It is important to realize that the argument of the logarithm ([ADP][Pi]/[ATP]) is generally not equal to the equilibrium constant, but only when the reaction is at equilibrium. SUMMARY A difference of concentration produces a flow of material in solution. Application of an electric force to charged particles also produces a flow of material in solution. These two forces can be united in the definition of a single force that is proportional to the negative gradient of a potential. This potential is the electrochemical potential written as µ 5 µ0 1 RT ln C 1 z`Ψ This electrochemical potential is the Gibbs free energy per mole. The Gibbs free energy, G, is an extensive variable, whereas µ is an intensive variable. G is also a state variable, depending only on the state of a system and not on the path it took to reach that state. The free energy is the maximum energy that can be extracted to do useful work. For all spontaneous reactions, the change in free energy, ΔG 5 Gfinal 2 Ginitial, is negative. Processes that require work occur spontaneously only if they are linked to other processes that lose free energy, so that for the overall process ΔG , 0. For any process at equilibrium, ΔG 5 0. Many cellular processes require energy for the synthesis of materials, transport, or mechanical or electrical work. These occur because energy is supplied by the hydrolysis of ATP. ATP hydrolysis to ADP and Pi has a large Electr ochemical P otentia l and Fre e E nergy negative ΔG under cellular conditions. The free energy of ATP hydrolysis per mole is given as ½ADP½Pi 0 ΔGATP hydrolysis 5 ΔG0 1 RT ln ½ATP Under “typical” cellular conditions, ΔGATP hydrolysis 5213.6 kcal mol21 5 257.1 kJ mol21. REVIEW QUESTIONS 1. In the formula for the electrochemical potential, what is R? What is T? What are the units of RT ln C? What units must C be in? What is z? What is the Faraday? What is Ψ? 2. What is the Gibbs free energy? 3. What is the relationship between the electrochemical potential and the Gibbs free energy? 4. How does the sign of ΔG determine the direction of a process? 5. Describe thermodynamic coupling. 6. What is the relationship between the free energy change and the equilibrium constant? 7. How would you determine the free energy change for ATP hydrolysis under cellular conditions? 8. What is the approximate free energy change of ATP hydrolysis in cells? 93 1.2 Problem Set Kinetics and Diffusion 1. A. The empirical formula of glucose is C6H12O6. What is its molecular weight? B. Isotonic glucose is 5% (w/v) glucose. How much glucose would we need to make 100 mL of isotonic glucose? 2. A. You need to make 250 mL of a stock solution of 0.1 M Na2 ATP. Its formula weight is 605.2 g mol21. How much Na2 ATP should you weigh out? B. Your advisor is skeptical of your abilities. He wants you to check out the 0.1 M ATP solution and tells you to do it spectrophotometrically. Spectrophotometry relies on the different abilities of chemicals to absorb light of specific wavelengths. A diagram of a spectrophotometer is shown in Figure 1.PS2.1. At particular wavelengths, chemicals absorb light according to their chemical structure and their concentration. The law governing the absorption of light is the BeerLambert Law: A 5 εCd where A is the absorbance; ε is a constant that depends on the chemical and typically varies with the wavelength of light—it is the molar extinction coefficient and is in units of M21; C is the concentration of the chemical (in M); and d is the path length. The molar extinction coefficient is defined for a path length of 1 cm. The absorbance is defined as A 5 logðI0 =IÞ where I0 is the incident light intensity and I is the transmitted light intensity. Your advisor tells you that ε259 5 15.4 3 103 M21; this is the molar extinction coefficient of ATP at a wavelength of incident light of 259 nm. He tells you to make a dilution of the stock by taking 25 μL of the stock solution and diluting it to 100 mL. What absorbance do you expect of the final diluted solution, if you made it up correctly, at λ 5 259 nm? 3. A. The molecular weight of ryanodine is 493.54 g mol21. You want to make 10 mL of a 10-mM stock solution. How much ryanodine should you weigh out? B. You make a dilution of the 10-mM ryanodine stock by pipetting 10 μL of the stock solution into a 10-mL volumetric flask and adding Light source Mirror Motor to set wavelength Mirror Sample cuvette PMT I0 I Diffraction grating FIGURE 1.PS2.1 Light path in a single beam spectrophotometer. The view is from above. Light from a source is collimated (making a narrow beam) and passed through a monochromator that selects a narrow band of wavelength of light to be passed through the sample. A photomultiplier tube (PMT) detects the light and measures its intensity. Comparison of this intensity, I, to the intensity when the sample is missing, I0, allows the calculation of the absorbance. Absorbance is recorded with time or as a function of wavelength. water to the mark. You measure the absorbance as a function of wavelength (against a water blank, using a standard 1 cm path length optical cell) and find a peak at 271 nm with an absorbance of 0.179. What is ε271 for ryanodine? (See Problem #2 for a discussion of the BeerLambert Law and a definition of the molar extinction coefficient.) 4. A. Magnesium chloride has a formula of MgCl2 6H2O. What is its formula weight? B. You desire to make 1 L of 0.1 MgCl2 solution. How much MgCl2 6H2O should you weigh? C. You need to make 25 mL of a 25-mM solution of MgCl2. How much of the 0.1 M stock solution do you add to the 25 mL volumetric flask? 5. The extracellular fluid volume varies with the size of the person. Suppose in an individual we determine that the ECF is 14 L. The average [Na1] in the ECF is about 143 mM. A. What is the total amount of Na1 in the ECF, in moles? in grams? B. Suppose this person works out and sweats 1.5 L with an average [Na1] of 50 mM. During this time the urine output is 30 mL 94 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00009-4 Problem Set with an average [Na1] of 600 mM. How much Na1 is lost during the workout? C. If the person does not drink fluids at all during the workout, what will be the [Na1] in the plasma at the end of the workout? Assume that all of the fluid in the sweat and urine originated from the ECF. 6. The body normally produces about 2 g of creatinine per day. The amount varies with individuals and is approximately proportional to the muscle mass. It is excreted through the kidneys according to urinary excretion of creatinine 5 GFR 3 plasma concentration of creatinine, where GFR is an abbreviation for “glomerular filtration rate.” If the GFR is 120 mL min21, what is the plasma concentration of creatinine at steady state? Hint: Assume the body is at steady state with respect to creatinine. 7. Just before noon, your plasma glucose concentration was 100 mg dL21. This plasma glucose is approximately evenly distributed among 3.5 L of plasma and 10.5 L of interstitial fluid that comprises your 14 L of ECF. Glucose is readily distributed in both compartments. You drink a can of soda that contains 35 g of glucose. A. How much would your blood glucose rise if all the glucose in the soda was absorbed and none of it was metabolized? B. Given that postprandial (after eating) increases in blood glucose amount to maybe 40 mg dL21, depending on the meal, over a period of an hour, give a crude estimate of the rate of glucose uptake by the peripheral tissues. Assume that the meal contains 100 g of carbohydrates and all of it is absorbed in 1 hour. 8. The association reaction for Ca21 and EGTA (a chemical that binds Ca21) is written as Ca21 1 EGTA$CaUEGTA Under defined and particular conditions of temperature and ionic mixture, the association constant was determined to be KA 5 2.52 3 106 M21. In a chemical mixture, 400 μM total EGTA was included and the free [Ca21] determined by a Ca21-selective electrode was found to be 4 3 1027 M. Assuming that there are no other binding agents for Ca21, what is the total [Ca21] in the mixture? 9. 2,4-Dinitrophenyl acetate decomposes in alkaline solution with a pseudo-first-order rate constant of 11.7 s21 at 25 C. It is a “pseudo”-first-order rate constant because it depends on the pH. A. If the initial concentration of DNPA is 1 mM, what is its concentration after 15 seconds? B. At what time is the concentration reduced to 0.5 mM (i.e., what is the half-life of the reaction)? C. After 5 minutes of reaction, what is the concentration of DNPA? 10. The following data were obtained for the rate of the Mg, Ca-ATPase activity of vesicles of cardiac sarcoplasmic reticulum as a function of temperature. What can you tell about the activation energy? Temperature ( C) ATPase Rate (µmol min21 mg21) 6.9 11.5 15.8 19.8 20.2 25.6 26.1 31.0 34.8 39.2 0.068 0.138 0.300 0.568 0.585 1.236 1.154 2.238 3.030 4.220 11. Superoxide reduces cytochrome C in the reaction 21 Cyt CUFe31 1 O2 1 O2 2 .Cyt CUFe where Cyt C Fe31 is the oxidized form and Cyt C Fe21 is the reduced form of cytochrome C. The reaction can be followed spectrophotometrically at 550 nm. The extinction coefficient for the reduced form of cytochrome C is εRED 5 2.99 3 104 M21 and the extinction coefficient for the oxidized form εOX 5 0.89 3 104 M21 (V. Massey, The microestimation of succinate and the extinction coefficient of cytochrome C. Biochimica et Biophysica Acta, 34:255256, 1959). See Problem #2 for a discussion of extinction coefficients and spectrophotometry. When xanthine oxidase converts xanthine to uric acid, it produces superoxide that can be measured using cytochrome C reduction. The following data were obtained for A550: Time (min) A550 0 1 2 3 4 5 6 0.1326 0.1478 0.1637 0.1791 0.1941 0.2073 0.2202 A. Calculate the rate of cytochrome C reduction. B. The xanthine oxidase was added in 75 μL of 6.5 mg XO per mL into a 3-mL reaction mixture. Calculate the specific activity of cytochrome C reduction (moles of cytochrome C reduced per min per mg of XO protein). 12. You suspect you are anemic and your physician orders some tests. He finds that your hemoglobin is 13 g%. The molecular weight of hemoglobin is 66,500 g mol21. A. What is the concentration of hemoglobin in molar in your blood? B. Each hemoglobin binds four oxygen molecules. If the hemoglobin is saturated with oxygen, what is the concentration of O2 bound to Hb, in molar? 95 96 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION ATPase ATP CH3 CH2 C O Pyruvate kinase HO Pyruvate NADH + H+ O P OH O– C O Protein O C C HO TABLE 1.PS2.1 Diffusion Coefficients and Mr for a Variety of Proteins ADP + Pi Lactate dehydrogenase (LDH) CH3 H C D 3 107 (cm2 s21) Milk lipase 6600 14.5 Metallothionein 9700 12.4 Cytochrome C 12,000 12.9 Ribonuclease 12,600 13.1 Myoglobin 16,890 11.3 Chymotrypsinogen 23,200 9.5 Carbonic anhydrase 30,600 10.0 O Phosphoenol pyruvate NAD+ Molecular Weight OH C HO O Lactic acid FIGURE 1.PS2.2 ATP hydrolysis by pyruvate kinase converts phosphoenolpyruvate to pyruvic acid. This is coupled by lactate dehydrogenase to the conversion of pyruvic acid to lactic acid and conversion of NADH to NAD1. The progress of the reaction can be followed spectrophotometrically by the change in absorbance of NADH. C. Convert the answer in B to volume using the ideal gas equation, PV 5 nRT, where T is the absolute temperature, R 5 0.082 L atm mol21 K21, V is the volume that we seek, and P 5 1 atm. The conditions for volume of gas are usually STPD—standard temperature and pressure, dry. The standard temperature is 0 C and pressure is 1 atm. 13. The rate of ATP hydrolysis by ATPases can be followed by the coupled enzyme assay shown in Figure 1.PS2.2. The progress of the reaction can be followed by A340. The extinction coefficient of NAD1 at 340 nm is negligible. The extinction coefficient of NADH at 340 nm is 6.2 3 103 M21. See Problem #2 for a discussion of extinction coefficients and spectrophotometry. In one reaction, the concentration of Ca-ATPase was 0.22 mg mL21 and A340 was 0.65 at t 5 0 min and 0.455 at t 5 2.0 min. What is the activity of the Ca-ATPase in units of μmol min21 mg21? 14. Show by representative calculations that Stirling’s formula pffiffiffiffiffiffiffiffiffi 0 n! 5 2πn en ln ðn 2nÞ is a good approximation for n! Use n 5 1, 2, 3, 4, 5. 15. Show that the equation rffiffiffiffiffiffiffiffiffiffiffi 1 2X2 e 4Dt Cðx; tÞ 5 C0 4πDt obeys Fick’s Second Law of Diffusion. 16. The intestinal enterocytes form a covering over the intestinal lining which, to the first approximation, can be considered to be a plane. Assuming no binding or sequestration within the cell, what is the estimated time of diffusion Peroxidase II 44,050 6.8 Albumin 68,500 6.1 Lactoperoxidase 92,620 6.0 149,100 4.6 Aldolase 17. 18. 19. 20. of Ca21 across the intestinal enterocyte? The length of the enterocyte is 20 μm and assume that the effective diffusion coefficient of Ca21 is about 0.4 3 1025 cm2 s21. Table 1.PS2.1 lists the diffusion coefficients and the molecular weight of a variety of proteins. What relationship can you deduce between the size and the diffusion coefficients of these soluble proteins? (Hint: regress ln D against ln Mr). Is the relationship you found consistent with the StokesEinstein equation? The free diffusion coefficient of oxygen in aqueous solutions is about 1.5 3 1025 cm2 s21. If the diffusion distance between air and blood is 0.5 μm, about how long is the diffusion time? Suppose a soluble protein has a molecular weight of 45 kDa and a density of 1.06 g cm23. Suppose further that the viscosity of the cytoplasm has a viscosity of 0.005 Pa s (about five times that of water—there is debate about the viscosity of cytoplasm with numbers varying from 0.001 to over 0.1 Pa s). A. Estimate the diffusion coefficient for the protein in the cytoplasm at 37 C. B. If the proteins were synthesized in the cell body, or soma, of a neuron in the spinal cord, about how long would it take to diffuse to the axon terminal 75 cm away? Diffusion coefficients in cytoplasm have been estimated by a technique of photobleaching recovery. In this technique, an area of the cytoplasm is irradiated with light to photobleach a fluorescent probe. Recovery of fluorescence in the region is achieved by diffusion of unbleached probes from adjacent areas of the cytoplasm. The translational diffusion coefficient can be estimated from the half-time of fluorescent recovery. (D. Axelrod et al., Mobility measurements by analysis of fluorescence photobleaching recovery kinetics. Problem Set Biophysical Journal 16:10551069, 1976.) This technique was applied to estimate the relative viscosity of cytoplasm and nucleoplasm by microinjecting fluorescein isothiocyanate-labeled dextrans of varying molecular sizes and measuring the fluorescence photobleaching recovery (I. Lang et al., Molecular mobility and nucleoplasmic flux in hepatoma cells. Journal of Cell Biology 102:11831190, 1986). These authors obtained the following data: Probe Molecular Weight (kD) Equivalent Radius (nm) D in Dilute Solution D in Cytoplasm D in Nucleoplasm D is in units of 1026 cm2 s21 FD20 FD40 FD70 FD150 17.5 41.0 62.0 156.9 3.30 4.64 5.51 9.07 0.651 0.463 0.390 0.237 0.080 0.044 0.029 0.015 0.069 0.056 0.036 A. Plot D against 1/a, where a is the molecular radius, for each of the solutions. From the StokesEinstein relation, you would expect the resulting curves to pass through the origin of zero diffusion coefficient with infinite radius. Do the curves extrapolate back in this way? Why or why not? B. Regardless of the intercept, the slope of the plot from part A ought to be related to the viscosity of the medium. Use the slopes to estimate the relative viscosity of the dilute solution, cytoplasm, and nucleoplasm. 97 Cell Structure Learning Objectives organisms and cells in multicellular organisms must solve a number of problems in order to survive. These include: G G G G G G G G G G G G G List the main categories of cellular function Describe the general structure, location, and function of the plasma membrane Describe the composition, location, and function of the cytosol Compare the general structure and function of microtubules, microfilaments, and intermediate filaments Describe the general structure, location, and function of the nucleus and its envelope Describe the general structure, location, and function of the mitochondria Describe the general structure, location, and function of the endoplasmic reticulum Describe the general structure, location, and function of the Golgi apparatus Describe the general structure, location, and function of lysosomes and peroxisomes Describe the general structure, location, and function of proteasomes Explain the general purpose of ubiquitinylation of proteins List the different types of cell-to-cell junctions G G G G G G FOR CELLS, FORM FOLLOWS FUNCTION In Chapter 1.1, we discussed the following points: G G G G 2.1 Cells are the organization unit of life. Cells come in a multitude of forms, specialized for their function. The vast majority of cells are somatic cells, and all of these contain the same set of genetic information, present in DNA and organized into genes. The multitude of forms comes from using only specific sets of the genes to make proteins. Multicellular organisms such as ourselves evolved because multicellular structures can provide an internal environment that is more stable than the natural environment, and thereby enhance the survival of the component cells and of the organism. Free-floating, single-celled organisms live at the mercy of environmental conditions, whereas protected, multicellular organisms can better withstand changes in the environment. Single-celled © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00010-0 Catalysis: Cells must be able to change one metabolite into another in order to synthesize cellular constituents, degrade them, or provide energy. Transport: Cells must be able to move things from outside the cell to inside or from one compartment to another within the cell. This includes bulk secretion or uptake of materials from the extracellular fluid. Signal transduction: Cells must have mechanisms for responding to signals from other cells or from within the cell. These may be chemical signals or electrical signals. Recognition: Cells attach to other cells and to extracellular structures. They must be able to recognize where they should form attachments. Movement: At some stage of their development, all cells must be able to move so as to position themselves properly within the cellular matrix that makes us up. Control: All of the activities of the cell must be coordinated. Control here also means that cells must select the parts of the genome that they will use. “Control” thus implies differentiation—the formation of specialized cells uniquely suited to their task. Proliferation: At appropriate times of development, cells must make new cells. This involves cell division and its control. ORGANELLES MAKE UP THE CELL LIKE THE ORGANS MAKE UP THE BODY The cells of the body typically are composed of a relatively small number of organelles that carry out specific functions of the cell, much like our organs do for us. These are called organelles because their relation to the cell is like the organs’ relation to the body. Table 2.1.1 lists these organelles along with their major function. The disposition of these organelles in a “typical” cell is shown in Figure 2.1.1. THE CELL MEMBRANE MARKS THE LIMITS OF THE CELL The cell membrane, also known as the plasma membrane, defines the inside and outside of the cell. Like all biological membranes, it consists of two lipid 101 102 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Microvilli TABLE 2.1.1 Major Organelles of the Cell and Their Function Organelles Function Plasma membrane Customs officer of the cell: determines what gets into or out of the cell, also signal transduction and cell recognition Cytosol Cell sap: the fluid medium in which soluble biochemicals diffuse and in which the other organelles are suspended Cytoskeleton Support, movement, and cell attachment Cytoskeleton Endocytotic vesicle Golgi Transport vesicles Secondary lysosome Polysome CIS face Rough ER Nucleus Command center: contains the hereditary material and organizes and controls differentiation Free ribosomes Factory of the cell for soluble proteins Rough ER Factory of the cell for membrane proteins and secreted proteins Nucleolus Smooth ER Synthesis of lipids and steroids Nucleus Golgi apparatus Shipping department: finishing and targeting of proteins to specific locations Plasma membrane Free ribosomes Smooth ER Peroxisome Centrioles Mitochondria Powerhouse of the cell: site of oxidation and energy transfer Lysosomes Exocytotic vesicle Primary lysosome Trans face Nuclear membrane Outer membrane Inner membrane Garbage disposal: destruction of worn-out organelles Mitochondria Proteasomes Destruction of tagged proteins Peroxisomes Oxidation of fatty acids and detoxification of xenobiotics layers, or leaflets, in which are embedded a variety of proteins that serve specific functions (see Chapter 2.4 for a discussion of the structure of lipids in biological membranes). Both the inside and outside layers may be coated with carbohydrate units that partly define the function of the lipids. An important function of this membrane is to determine what goes into and what comes out of the cell. The “customs officer” of the cell has several mechanisms that can transport materials into or out of the cell. These include: G G G G Passive transport Active transport Exocytosis Endocytosis. We will discuss all of these in more detail in Chapters 2.5 and 2.6. Passive transport requires no metabolic energy and may involve diffusion through the lipid layer of the membrane or through water-like channels that are established by proteins that span the membrane. Active transport requires the input of cellular metabolic energy and may be primary or secondary. Primary active transport directly couples transport to metabolic energy. Secondary active transport indirectly links transport of a material to metabolic energy. Exocytosis and endocytosis refer to the movement of materials that are enclosed in vesicles. These vesicles are tiny hollow spheres of membrane. Secretory vesicles 1×10–6 m FIGURE 2.1.1 A “typical” human cell showing various subcellular organelles including the plasma or cell membrane, nucleus, nuclear membrane, rough ER, smooth ER, transport vesicles, Golgi apparatus, mitochondria, cytoskeletal elements, lysosomes, peroxisomes, and endocytotic and exocytotic vesicles. Although there is no “typical” cell, most cells contain this set of organelles. Exceptions to this rule include the adult erythrocytes. are filled with some material by the cell. These vesicles can fuse with the plasma membrane, and, in so doing, they dump their contents into the extracellular space. This process is called exocytosis. Endocytosis is similar but occurs in the opposite direction. In this case, parts of the plasma membrane invaginate and pinch off to form endocytotic vesicles inside the cell. Endocytosis of fluid is called pinocytosis or “cell drinking”; endocytosis of particulate stuff is called phagocytosis or “cell eating.” In addition to these functions, the cell membrane forms the hub of signal transduction and surface recognition. It must transfer extracellular signals originating from other cells to an intracellular signal inside the cell. This is what is meant by “signal transduction.” The cell membrane receives the first messenger in the form of a chemical or electrical signal, and receipt of the first messenger causes the formation of a second messenger inside the cell. Surface recognition occurs through surface proteins that are members of the major histocompatibility complex or MHC. These proteins are responsible for beginning transplant rejections by recognizing the transplants as foreign matter. Cell St ructur e THE CYTOSOL PROVIDES A MEDIUM FOR DISSOLUTION AND TRANSPORT OF MATERIALS The cell membrane surrounds all the constituents of the cell, which themselves exist in a fluid medium that allows transfer of materials among them. This fluid medium is the cytoplasm, which literally means “cell fluid.” It includes the cytosol and all of the organelles suspended in it. The cytosol itself is the fluid that contains dissolved ions and organic compounds of a bewildering variety. These include amino acids for building proteins, glucose for energy, a tremendous variety of metabolic intermediates, and cytoplasmic enzymes for glycolysis, the first stage in burning carbohydrates for energy. The ionic composition of the cytoplasm varies with different cell types, but Table 2.1.2 gives some reasonable approximate numbers for the “typical” cell. The cytoplasm also serves as a medium for the transmission of control signals from the outer surface of the cell to the interior, and from the nucleus to the rest of the TABLE 2.1.2 Selected Components of the Cytosol Component 1 Concentration Na 14 3 1023 M K1 120 3 1023 M Cl2 10 3 1023 M HCO32 10 3 1023 M ATP 5 3 1023 M ADP 50 3 1026 M Mg 0.5 3 1023 M Ca21 0.1 3 1026 M 21 pH 7.17.2 Osmolarity 295 mOsm L21 α-Tubulin cell. Although we describe it here as a fluid, the cytosol is not like water: ions diffuse through the cytosol slower than they do through water (see Chapters 1.6 and 1.PS2, Problem #20). Cutting a muscle fiber, for example, does not cause its fluid to leak out like water. The cytoplasm is more akin to a gel. Most of this behavior is due to the small volume of fluid and the abundance of cell surfaces. Although bulk water flows, a thin film adheres to any wettable surface. This thin film generally exceeds the thickness of a cell. Thus, at the cell level, surface forces govern much of bulk fluid flow. THE CYTOSKELETON SUPPORTS THE CELL AND FORMS A NETWORK FOR VESICULAR TRANSPORT Arrays of protein filaments form a network within the cytoplasm. These filaments determine the shape of the cell and provide for the movement of the cell as a whole or for the movement of organelles from one part of the cell to another. There are four major types of filaments comprising the cytoskeleton: 1. 2. 3. 4. Microtubules Intermediate filaments Actin filaments or microfilaments Myosin filaments. MICROTUBULES ARE THE LARGEST CYTOSKELETAL FILAMENTS The microtubules are about 25 nm in diameter and are constructed of heterodimers of tubulin, a globular polypeptide of 50,000 Da. These dimers are assembled into protofilaments of tubulin dimers with the β-tubulin of one joined to the α-tubulin of the next. The microtubules are assembled from 13 such protofilaments arranged in a cylinder with a hollow core (see Figure 2.1.2). A number of microtubule-associated proteins (MAPs) bind to microtubules. Some of these MAPs are “motor β-Tubulin GTP GDP 8 nm Singlet Doublet Protofilament – End 25 nm + End GTP FIGURE 2.1.2 Schematic diagram of the structure of a microtubule. Microtubules consist of protofilaments composed of tubulin dimers, one α-tubulin and one β-tubulin. Both bind GTP, guanosine triphosphate. GTP, like ATP, stores chemical energy in its phosphate bonds. Hydrolysis of GTP to GDP can be used to stabilize protein shapes. The α-tubulin does not hydrolyze its GTP, whereas the β-tubulin hydrolyzes its GTP to GDP; 13 of the protofilaments assemble to form a singlet microtubule. Because of the asymmetrical arrangement of the tubulin monomers, the microtubule has an asymmetry, with a plus (1) end and a minus (2) end. The plus (1) end is the end pointing away from the origin of the microtubule and is the end to which monomers add to the growing microtubule. These ends differ in the rates of tubulin association and dissociation. Because these rates differ, the microtubule can treadmill—dissolve at one end while lengthening at the other. Tubulin can also form doublet and triplet structures. Cross-sections of a singlet microtubule and a doublet microtubule are shown. 103 104 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION proteins” that can “walk” along the microtubule. Two of these motor proteins are named kinesin and dynein. Kinesin forms a family of motor proteins with about 40 members. Most of these are “ 1 directed” motors, moving along the microtubules toward the 1 end. Dyneins comprise a family of 2 directed motors. These motor proteins can attach vesicles and then move along the microtubules, carrying their vesicles along. In this way, the microtubules can provide a track along which intracellular transport occurs. This is especially important in neurons in which transport must occur down a long narrow process of the cell. Figure 2.1.3 shows a highly schematic cartoon illustrating kinesin and dynein movement along a microtubule. the cilia to bend. The waving cilia move the extracellular fluid past the fixed cell. In this way, the movement of the cilia causes the movement of extracellular fluid. In the lungs, the cilia move mucus, trapped dust, and foreign material toward the pharynx where they can be expelled. In the Fallopian tubes, the cilia help move the ovum toward the uterus. Figure 2.1.4 shows a schematic diagram of the structure of a cilium. Activation of the dynein arms linking two isolated microtubules will ordinarily cause the two microtubules to slide past each other. In the cilia, linking proteins turn this motion into bending of the cilia. Microtubules also form the mitotic spindle during cell division. Nearly all cells possess two centrioles oriented at right angles. These form a microtubule organizing center (MTOC) that provides a scaffold for the assembly of microtubules. Another kind of tubulin, γ-tubulin, binds to accessory proteins to form a γ-tubulin ring complex (γ-TuRC) that acts as a nucleation site for microtubules. The centrosome consists Microtubules can also form the interior of larger structures called cilia that extend out from the cell into the extracellular fluid. These cilia have a special arrangement of nine doublets arrayed circumferentially around two central microtubules. In this case, the microtubules are cross-linked such that the action of dynein causes Cargo Ankyrin Dynactin complex Spectrin Light chains Dynein Kinesin – End + End FIGURE 2.1.3 Schematic model of microtubule motor proteins. Kinesins typically have two globular heads and an elongated coiled tail. The tail regions of most kinesins bind cargo, either membrane-enclosed vesicles or microtubules. Dyneins may contain two or three globular heads and a large number of accessory proteins that bind vesicle cargo. Dynein itself is a complex assembly that requires a second complex assembly, dynactin, to transport cargo. A possible arrangement of some of these structures is shown. + + – – Plasma membrane Radial spoke Doublet microtubule Outer dynein arm Inner dynein arm Linking protein Cell membrane FIGURE 2.1.4 Structure of a cilium. A cilium contains a “9 1 2” arrangement of microtubules. Nine doublet microtubules containing A and B microtubules with 13 and 11 protomers, respectively, surround a central pair of microtubules. Multiple structures link these microtubules. The dynein arms move toward the minus end of the microtubule. In isolated tubules, this would cause sliding of one microtubule past another. The linking proteins turn this motion into a bending of the cilium. Cell St ructur e of the two centrioles surrounded matrix containing many copies of microtubules that grow out of this plex provide tracks for chromosome cell division. by a centrosome the γ-TuRC. The centrosome commovement during ACTIN FILAMENTS ARISE FROM NUCLEATION SITES USUALLY IN THE CELL CORTEX Actin filaments are present in most cells but are especially abundant in muscle cells. The monomer is a globular protein called G-actin, with a molecular weight of 41,800 Da. G-actin polymerizes noncovalently into actin filaments, called F-actin. Actin filaments consist of two strands of globular molecules twisted into a helix with a repeat distance of about 36 nm. The filament is asymmetric having distinguishable ends that are detectable by the way in which it interacts with myosin, another protein that is present in many cell types but is especially abundant in muscle. Thus the actin filament also has a plus end (the growing end) and a minus end (the nucleation or beginning end). Each individual actin filament is about 3.5 nm across, so that F-actin has a diameter of about 7 nm. Assembly and stabilization of filamentous, or F-actin, is described in Figure 2.1.5. Actin filaments determine the shape and movement of the cell’s surface, including structures such as microvilli, which are fingerlike extensions of epithelial cells that line internal structures like the intestinal villi and kidney tubules. The membrane of these cells anchors the actin filaments and extends them into a web of cytoskeletal elements in the main body of the cell. Other proteins can cross-link actin microfilaments together to form bundles of filaments or gel-like networks. These cross-linking proteins include α-actinin, fimbrin, and villin, which bundle actin filaments together. Spectrin and filamin both have two actinbinding sites. They join two actin filaments together to form a web of supporting filaments. INTERMEDIATE FILAMENTS ARE DIVERSE Intermediate filaments were originally named because with diameters between 8 and 10 nm, they are intermediate in size between the microtubules (at 25 nm) and the microfilaments at 7 nm. These intermediate filaments are composed of a number of different proteins. They play some structural or tension-bearing role. They differ from microtubules and microfilaments in that: G Both microtubules and microfilaments are made by the polymerization of globular monomeric Thymosin binding inhibits actin association Thymosin ARP-2, ARP-3 complex nucleates actin polymerization G-actin is activated by binding ATP G-actin ATP Profilin binding activates actin binding to the plus end ARP complex After polymerization, actin hydrolyzes its ATP but keeps ADP tightly bound Profilin Capping protein (Cap Z ) + – – + Tropomodulin caps some F-actin at the minus end Cap Z caps some F-actin at the plus end FIGURE 2.1.5 Assembly and stabilization of microfilaments (actin filaments). Actin binds ATP and begins assembly by binding to actin-related proteins (ARPs) that serve as a nucleation site, usually just under the cell membrane in the cortex of the cell. The ARP complex can also bind F-actin on the side of the filament, so it can build a tree-like web from individual actin filaments. After assembly, actin hydrolyzes its bound ATP, but the ADP remains tightly bound. Formation and stabilization of F-actin is regulated by proteins that bind the free monomer. Thymosin binds to the free monomer and inhibits its association with either the minus or plus end of the F-actin. Profilin binds to the free monomer and inhibits its association with the minus end but markedly enhances its association with the plus end. Cap Z binds to the plus end of the F-actin and stabilizes it. The minus end can be stabilized by remaining bound to the ARP complex. In muscle cells, tropomodulin binds to the minus end and stabilizes it. 105 106 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Head Rod Tail 45 nm dimer N terminal C terminal 60 nm tetramer 2 dimers 60 nm unit-length filament 8 tetramers Full length filament > 240 nm FIGURE 2.1.6 Highly schematic representation of the structure of intermediate filaments. The elementary subunit of the intermediate filaments consists of an elongated rod with an N-terminal “head” and a C-terminal “tail.” A variety of these elementary subunits are made by the body. These associated laterally to form a homo- or heterodimer, approximately 45 nm long. These dimers further associate laterally, with an offset, to form a 60nm-long tetramer. Eight of these tetramers further associate laterally to form a unit length filament. These unit length filaments can associate end to end to form longer filaments. The width of the mature filaments is not eight times the width of the tetramer, as these associate in three dimensions to form a mature filament about 11 nm in diameter. After the initial association into full length filaments, the filaments are radially compacted to form the mature filaments. G G G proteins, but the intermediate filaments are made of elongated (45 nm) and thin (23 nm) rod-like dimers. The intermediate filament units align with their long axis parallel to the filament axis, and filament width is determined by lateral association of the dimers (see Figure 2.1.6) Both microtubules and microfilaments are polar, which allows the active movement of motor proteins with their associated cargo along the filaments. Assembled intermediate filaments have no polarity because individual monomers are oriented in both directions along the axis of the filament. Intermediate filaments differ from both microtubules and microfilaments in that reversible association and dissociation of intermediate filament dimers can occur all along the length of the filament, whereas association and dissociation of microtubules and microfilaments occur only at their ends. This process is called dynamic subunit exchange. However, the exchange occurs much slower than the exchange of subunits in microtubules and microfilaments. Unlike tubulin and actin, the subunits of the intermediate filaments do not bind a nucleotide. The intermediate filaments are diverse; some 73 separate proteins in humans have been identified encoded by over 70 genes. They all consist of three parts: a “head,” a long rod-like central part, and a “tail.” The members of the IF family have been subdivided into five distinct groups based on their structure, mode of assembly, and developmental expression in different tissues. These groups and their types are summarized in Table 2.1.3. There is considerable variation within types. For example, there are over 50 different varieties of keratin. CYTOSKELETAL UNITS FORM FREEFLOATING STRUCTURES BASED ON TENSEGRITY Buckminster Fuller in the 1960s invented the word “tensegrity” as a blend of tension and integrity. He used it to describe architectural structures of remarkable rigidity that were composed of compressive rods and elastic cables. These two elements can be combined to form stable structures. Cells cannot rely on their membranes for structural stability because the membranes themselves are weak. But if you drape the membranes over cytoskeletal elements, structural strength can be achieved. In this view, the microtubules are the rigid rods and intermediate filaments are the tension-bearing elements. The actin and myosin elements allow for the movement of the cytoskeleton and the consequent movement of the attached membrane. In this way, the cell can extend processes or move from one place to another. Cell St ructur e TABLE 2.1.3 Classification of the Intermediate Filaments Types of IFs Protein Tissue Distribution Proposed Function Associated Diseases Type I Type II Acidic keratins Basic keratins Epithelial tissues Epithelial tissues Tissues strength and integrity Epidermolytic hyperkeratosis Type III Desmin GFAP Peripherin Vimentin Syncoilin Muscle Glial cells Sarcomere organization Dilated cardiomyopathy Alexander disease Type IV Neurofilaments NF-L, NF-M, and NF-H Nestin Synemin α, β Neurons Axon organization Amyotrophic lateral sclerosis; Parkinson disease Type V Nuclear lamins type A, B1, B2, C1,C2 Nucleus Nuclear organization and signaling HutchinsonGilford progeria; limbgirdle muscle dystrophy; EmeryDreifuss muscular dystrophy Type VI Filensin Phakinin Lens Cataracts Cytoplasmic filaments Outer nuclear membrane Transporter subunit Ring subunit Column subunit Lumenal subunit Inner nuclear membrane 50 nm Nuclear lamina proteins Basket FIGURE 2.1.7 Cartoon of the structure of the nuclear pore. The nuclear pores consist of a complex of more than 50 different proteins that form a complicated structure with octagonal symmetry. Small molecular weight materials (, 20 kD) can pass through these pores in both directions, but the movement of larger materials, such as RNA and ribosomes, is regulated. The outer membrane faces the cytosolic compartment and the inner membrane faces the nuclear compartment. Two rings made of eight subunits each are connected by columnar scaffold subunits. Lumenal subunits anchor the scaffold to the membrane. The ring subunits connect to fibrils that form a basket structure on the nuclear side. Transport subunits in the interior of the pore actively transport materials into or out of the nucleus. MYOSIN INTERACTS WITH ACTIN TO PRODUCE FORCE OR SHORTENING THE NUCLEUS IS THE COMMAND CENTER OF THE CELL Myosin also exists in multiple isoforms. A major form is a protein of about 200 kDa that forms a homodimer with two long tails forming a coil, a hinge region, and a head region that binds actin filaments, hydrolyzes ATP, and “walks” along the actin filament. The interaction of the actin filament relative to the myosin filament causes either shortening of the acto-myosin thread or production of force. This mechanism is responsible for muscle force and also produces movement in nonmuscle cells. Most cells have linear dimensions on the order of 2050 μm. The nucleus is the largest organelle, with a diameter of about 35 μm. The nucleus is bounded by a double membrane, the nuclear envelope, that has pores for materials to move between nucleus and cytoplasm, as shown in Figure 2.1.7. The nucleus contains nearly all of the DNA of the cell. As described in Chapter 2.2, this DNA carries the information that allows the synthesis of specific proteins. The nucleus 107 108 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION also contains a specialized region called the nucleolus. This is a diffuse region that is not delimited by a membrane. The nucleolus is involved in the synthesis of ribosomes. RIBOSOMES ARE THE SITE OF PROTEIN SYNTHESIS Ribosomes exist either free in the cytoplasm or bound to membranes of the rough endoplasmic reticulum (ER). Both types consist of two main subunits, designated 60S and 40S, where the S refers to Svedbergs and describes how fast the particles sediment during centrifugation (see Appendix 2.1.A1). The larger subunit consists of three different strands of ribonucleic acid (RNA) and about 49 different proteins. The smaller subunit has a single RNA strand and about 33 other proteins. Both are assembled in the nucleolus, a specialized region of the nucleus that is not membrane bound. There, large loops of DNA containing ribosomal RNA genes are transcribed by the enzyme RNA polymerase I to ribosomal RNA or rRNA. Ribosomal proteins made in the cytoplasm are imported back into the nucleus and assembled, along with the rRNA, into the two subunits. The two subunits join to form a functional ribosome in the cytoplasm that makes a platform on which proteins are synthesized, as described more fully in Chapter 2.2. THE ER IS THE SITE OF PROTEIN AND LIPID SYNTHESIS AND PROCESSING The ER is a membranous network within the cell fluid, or cytoplasm, that is continuous with both the outer nuclear membrane and the plasma membrane. The membrane forms flattened disks with an enclosed space called cisternae. Electron micrographs and density gradient centrifugation reveal two functionally distinct regions of the ER; these are rough ER and smooth ER. The “rough” ER was given that name because the many ribosomes attached to the membrane give it a granular appearance in electron micrographs. The smooth ER lacks these attached ribosomes. Special proteins, called translocons, span the ER membrane and bind ribosomes to the cytoplasmic face of the rough ER. Protein synthesis occurs on the ribosomes. Some of these are free within the cytoplasm, and the proteins made there are generally soluble proteins that remain in the cytoplasm. The ribosomes on the ER synthesize proteins that pass through the ER membrane as they are being synthesized. These proteins may become embedded in membranes or they may be destined for secretion from the cell. After synthesis, many proteins are further processed within the ER cisternae, preparing them for secretion or targeting them for some location within the cell. THE GOLGI APPARATUS PACKAGES SECRETORY MATERIALS The Golgi apparatus consists of sets of membranedelimited smooth-surfaced cisternae. Each set of flattened, disk-shaped cisternae resembles a stack of pancakes. This structure is called a Golgi stack or dictyosome. It is about 1 μm in diameter and is usually located near the nucleus and near the centrioles that define the cell center. The number of cisternae in a stack varies from 6 to 30, and the number of Golgi stacks in the cell varies enormously with the biochemical activity of the cell. Golgi stacks are polarized with two distinct faces. The cis or forming face is nearest a smooth transitional portion of the rough ER. The trans or maturing face typically faces the plasma membrane. Swarms of small vesicles (about 50 nm in diameter) cluster on the cis face of the Golgi stack. A large number of vesicles associate with the sides of the stack near the dilated rims of each cisternae. In electron micrographs, these vesicles sometimes appear to bud off the Golgi cisternae. In secretory cells, larger vesicles containing high concentrations of secreted proteins appear to originate from the trans face of a Golgi stack. The Golgi stacks are a processing station for proteins manufactured in the rough ER. Proteins made in the rough ER travel to the Golgi through the small transport vesicles. In the Golgi, the proteins are processed and packaged for delivery to various locations throughout the cell, including packaging for eventual fusion with the plasma membrane and secretion into the extracellular space (see Figure 2.1.8). THE MITOCHONDRION IS THE POWERHOUSE OF THE CELL Mitochondria produce much of the cell’s ATP by coupling the chemical energy of oxidation of metabolites to the synthesis of ATP. Their main structural features are shown schematically in Figure 2.1.9. The matrix contains many different enzymes required for the oxidation of pyruvic acid and fatty acids, including those involved in the tricarboxylic acid cycle. The mitochondrial matrix also includes mitochondrial DNA, special mitochondrial ribosomes, tRNAs, and enzymes that are required for the expression of mitochondrial genes. The inner mitochondrial membrane contains a number of important proteins that collectively comprise the electron transport chain and another enzyme complex called the ATP synthetase that makes ATP from ADP and Pi, the reverse of the ATP hydrolysis reaction discussed in Chapter 1.7. These complexes are discussed further in Chapter 2.10. Briefly, these complexes couple the chemical energy derived from the oxidation of fuels obtained from food to the synthesis of ATP. This is the site of oxygen consumption by aerobic cells. Lynn Margulis originally postulated that the mitochondria in aerobic cells that contain a nucleus (eukaryotic cells) Cell St ructur e Apical cell membrane Zymogen granules Condensing vacuole trans face originated from the engulfment of aerobic bacteria by anaerobic single-celled organisms. This hypothesis, called the endosymbiotic hypothesis, derives from the similarity of mitochondria to bacteria. Both have circular DNA; both are approximately the same size; both reproduce by dividing into two, asexually; mitochondrial ribosomes resemble bacterial ribosomes rather than eukaryotic ribosomes, and both bacteria and mitochondria share a slightly different genetic code from that in the nucleus. LYSOSOMES AND PEROXISOMES ARE BAGS OF ENZYMES Retrograde transport vesicles Golgi stack cis face Rough ER Ribosomes + mRNA FIGURE 2.1.8 Packaging of secreted proteins in secretory cells of the pancreas. Proteins are synthesized on the membrane of the rough ER and translocated into the lumen of the ER as they are being made. The ER forms transport vesicles that fuse to form the cis face of the Golgi stack. These membranes progress through the stack as new layers are added on the cis face and taken away on the trans face. On the trans, or maturing, face of the Golgi stack, the enclosed proteins are collected in secretory vesicles. These then fuse into condensing vacuoles that concentrate the proteins to form zymogen granules. These granules lie in the apical aspect of the secretory cells, adjacent to the plasma membrane. Upon stimulation, these granules fuse with the apical membrane and release their contents of secretory enzymes into the lumen of a duct, or channel, that takes the enzymes into the intestine. Lysosomes are membranous bags of hydrolytic enzymes including proteases, nucleases, glycosidases, lipases, phospholipases, and phosphatases. These hydrolytic enzymes are acid hydrolases, being optimally active in an acid environment. Lysosomes are typically 0.20.3 μm in diameter. They originate from the trans face of the Golgi stack and are formed first as primary lysosomes. The primary lysosome fuses repeatedly with a variety of membrane-bound substrates including endocytotic vesicles, phagocytotic vesicles, and worn-out intracellular organelles. After fusion, the combined vesicle forms a secondary lysosome. Because of its diverse substrate contents, the secondary lysosomes have a diverse morphology. Lysosomes degrade phagocytosed material and worn-out parts of the cell. The peroxisome is another membrane-bounded vesicle, with a diameter of about 0.5 μm. It contains oxidative enzymes such as catalase, d-amino acid oxidase, and urate oxidase. Like the mitochondria, the peroxisomes are a major site of O2 utilization. The peroxisome detoxifies foreign chemicals and metabolizes fatty acids. Beta-oxidation, a process in which fatty acids are shortened by two carbons to form acetyl-coenzyme A, occurs in both mitochondria and peroxisomes (see Chapter 2.11). Outer mitochondrial membrane Matrix Cristae Intermembrane space Inner mitochondrial membrane FIGURE 2.1.9 Typical features of a mitochondrion. The mitochondria, shown cut longitudinally, are membrane-delimited structures about 0.51.0 μm wide and 14 μm long. The outer mitochondrial membrane is permeable to many materials with molecular weights below 10 kDa. The inner mitochondrial membrane is impermeable to most materials but can transport specific materials. It is folded to form the shelf-like cristae and encloses the matrix. The intermembrane space lies between the inner and outer membranes. 109 110 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION PROTEASOMES DEGRADE MARKED PROTEINS Proteasomes are large, multisubunit protein complexes that are scattered throughout the cytoplasm and that degrade cell proteins. Proteins are tagged for degradation by the attachment of a ubiquitin molecule to the proteins. Ubiquitin is a protein consisting of 76 amino acids. Proteins can be marked by more than one ubiquitin chain through a complex series of reactions involving several different enzymes. This process is called ubiquitinylation, in which a series of enzymes recognizes the proteins to be degraded and adds activated ubiquitin onto them. Different sets of enzymes in this pathway recognize different degradation signals on proteins. In this way, each set of enzymes targets distinct subsets of proteins that bear particular degradation signals. The process of ubiquitinylation is illustrated in Figure 2.1.10. 1. a barrier function that disallows free movement of materials between cells in an epithelial sheet. This is the so-called “tight” part of tight junctions, but in reality the junctions vary considerably in their permeabilities. 2. a fence function that prevents the free migration of membrane components from the apical surface Ub ATP AMP + PPi Ub E1 E1 E2 E3 Ub E2 The proteasome consists of a central hollow cylinder capped at both ends. The central cylinder consists of a stack of four 7-membered rings. The caps recognize the ubiquitin coats of proteins that have been tagged for degradation. Each cap contains ATPase activity, releasing chemical energy in the process. These caps use the energy to unfold the ubiquitinylated proteins and move them into the central cylinder where proteases cleave the protein into its constituent amino acids. The structure of the proteasome is shown in Figure 2.1.11. E3 Ub Ub Ub Substrate E2 E3 Proteasome CELLS ATTACH TO EACH OTHER THROUGH A VARIETY OF JUNCTIONS Cells form a variety of attachments to each other or to the extracellular matrix. These include: G Zonula occludens: This “tight” junction joins epithelial cells in an occluding zone at one pole of the epithelium. It has three functions: FIGURE 2.1.10 The ubiquitinylation reaction. Ubiquitin, noted as Ub in the figure, is a 76 amino acid protein that is used to “tag” proteins as being ready for degradation. First, Ub is attached to E1 through an ATPrequiring reaction. It is then passed from E1 to E2, and from there to the protein substrate being tagged for demolition. The ubiquitinylated protein is then degraded by a specialized cell structure called the proteasome. 19S regulatory particle Lid Base α ring β ring β ring α ring 20S core particle Chambered protease Complete proteasome FIGURE 2.1.11 Structure of the proteasome. The overall structure consists of a 20S core particle of 28 subunits capped on one or both ends by a 19S regulatory particle containing at least 19 subunits. The core particle consists of four 7-membered rings that is symmetrical about a plane perpendicular to the long axis of the particle. Proteolytic activity resides in the two middle layers of the core particle. Cell St ructur e G G G of the cell to the lateral surface. This effectively partitions the cell membrane into components. 3. a signal function that help regulate cell proliferation, differentiation, and polarity. Zonula adherens: This is a belt of attachment that typically surrounds epithelial cells just below the zonula occludens—meaning toward the basolateral pole of the cell. Desmosomes: These are “spot welds” between cells and are constructed of different proteins than those that make up zonula adherens. Gap junctions: These junctions serve to electrically connect cells because they allow small ions to pass from one cell to another, and these ions carry electrical current. Each of these junctions is made up of complexes of many proteins and a variety of proteins can associate with these. Figure 2.1.12 illustrates the structure of the desmosome, zonula adherens, zonula occludens, and gap junction, and Figure 2.1.13 shows their use in an epithelial sheet in the small intestine, consisting of a lining of cells that separate the ingested food and gastrointestinal secretions from the blood. Many of the proteins that make up these structures are present in the body in various isoforms—variants of the protein that serve the same basic function but in different tissues. For example, there are some 20 different types of connexins that associate as hexamers to form the connexons in gap junctions. Adherens junction Gap junction Microtubule Plekha7 Nezha β-catenin p-120 α-catenin Vinculin Actin filament Cadherins α-actinin Connexon Plakoglobin Connexin Nectin Afadin/AF6 Desmosome Desmocolin Desmoglein Tight junction (zonula occludens) Cell membranes Plakophilin Plakoglobin Intermediate filaments Jam ZO1 Claudin ZO2 Inner dense Outer dense plaque plaque Desmoplakin Dense midline Occludin Jam FIGURE 2.1.12 Schematic diagram of the proposed structures of gap junctions, adherens junctions, and desmosomes. Gap junctions form by linking of connexons on opposing membranes. Each connexon consists of a hexamer of connexin units that form a central pore. When two connexons link, the pore of one lines up with the pore of another, forming a watery path between the cells. This allows diffusion of small, soluble materials from one cell to another without crossing the cells’ membranes. Adherens junctions form by interaction of extracellular parts of cadherin molecules. These proteins are embedded in the cell membrane and have a short cytoplasmic tail. This cytoplasmic tail binds to a variety of proteins including p-120, plakoglobin, and β-catenin. These in turn bind other proteins that eventually form initiation sites for actin polymerization. Cadherin also directly binds microtubules, although other proteins may stabilize this interaction. Desmosomes adhere two cells together because of the interaction of extracellular domains of desmoglein and desmocolin. These penetrate the membrane and their cytoplasmic domains bind other proteins, plakophilin, plakoglobin, and desmoplakin, that eventually connect to intermediate filaments. Tight junctions consist of binding of claudins and occludins, along with junctional adhesion molecule (JAM). There are multiple isoforms of each of these that produce junctions of varying permeability. These are stabilized by multiple cytoplasmic accessory proteins, including zonula occludens proteins (ZO-1, ZO-2, and ZO-3). These can connect to the cytoskeleton through other proteins such as afadin. In the junction between three cells, a special protein called tricellulin is required to seal the gap between the cells. 111 112 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Tight junction: seals adjacent cells in an epithelial sheet Adherens junction: joins actin bundles in adjacent cells Actin Desmosome: joins intermediate filaments in adjacent cells Junctional complex Intermediate filaments Gap junction: joins cells with a watery channel Connexons Hemidesmosome: anchors the cell to extracellular matrix through intermediate filaments Basal lamina FIGURE 2.1.13 Cell attachments found in an epithelial sheet. The intestinal cells form a layer, referred to as an epithelium, that lines the intestine. The intestinal cells are bound at their apical pole, the side facing the gastrointestinal lumen, by a junctional complex consisting of tight junction, adherens junction, and desmosomes. The tight junction and adherens junction form a belt completely surrounding the cell. The desmosomes are located in spots. The cells are also joined by gap junctions that allow free passage of small molecular weight materials between the cells. The cells attach to the extracellular basal lamina through hemidesmosomes that connect the fibers in the extracellular matrix with the cytoskeletal intermediate filaments. SUMMARY The cell is the fundamental organizational unit of the body. Although cells come in many different forms, they share many features. Each cell consists of a number of organelles, so named because they contribute to overall cell function in much the same way as our organs contribute to bodily function. The cell membrane determines the inside and outside of a cell. This is the “customs officer” of the cell, determining what enters or exits the cell. Materials move through the cytoplasm, the watery cell fluid that transports materials from one place to another, largely by diffusion. A cytoskeleton maintains cell shape and provides movement. Cytoskeletal elements include microtubules, which are hollow rods formed from 13 protomer filaments of a heterodimer of tubulin, actin filaments, composed of a double helix of actin monomers strung together, intermediate filaments of various descriptions, and myosin filaments. These cytoskeletal elements are dynamic and allow the cell to change shape and to transport materials along cytoskeletal tracks. The nucleus is the single largest organelle, and it is the “command post” of the cell. It is enclosed by a double membrane that is pierced by numerous nuclear pores. In response to signals, the nucleus “expresses” select regions of the genome. This means that the nucleus specifically converts some DNA into mRNA, but not all. The mRNA then makes proteins in the “factory” of the cell located on ribosomes either free in the cytoplasm or bound to the surface of the rough ER, another membranous network in the cytoplasm. The smooth ER makes lipids; the rough ER makes membrane-bound proteins and proteins destined for export from the cell. The rough ER transfers its protein content to the cis side of the Golgi apparatus, which is the “shipping and packaging” department of the cell. The materials move from one part of the Golgi apparatus to another in tiny membrane spheres called vesicles. At each stage, the proteins are processed further. The final vesicles leaving the trans face of the Golgi stack are ready for export from the cell by exocytosis. All of these activities of the cell require chemical energy supplied as ATP. The mitochondria make ATP by coupling the chemical energy liberated by oxidation of foodstuffs to the synthesis of ATP. Thus the mitochondrion is the “powerhouse” of the cell. It consists of a double membrane structure. The electron transport chain is on the highly folded inner membrane. This membrane synthesizes ATP from chemical precursors and is the main site of oxygen consumption by the cell. The cell also contains a variety of other organelles including the lysosome, peroxisome, and proteasome. These structures degrade material engulfed by the cell Cell St ructur e by endocytosis and also degrade worn-out organelles and cell proteins. Cells form a variety of attachments to other cells and to the extracellular matrix. These include zona occludens and adherens junctions, gap junctions, and desmosomes. These form by the complex association of a variety of different proteins: connexons on opposite membranes link up to form the gap junction; cadherins link up to form adherens junctions; desmoglein and desmocolin form the desmosomes; claudin and occludin link up in the tight junction. These junctions form by the association of extracellular parts of these transmembrane proteins. The intracellular parts join up with still other proteins that eventually connect the junctions with the cell’s cytoskeleton. REVIEW QUESTIONS 1. What is the plasma membrane? What compartments does it separate? What is endocytosis? Exocytosis? Pinocytosis? Phagocytosis? 2. What is the major cation (positively charged ion that migrates toward the cathode, the negative electrode) of the cytosol? Is the cytosol comparable to water? 3. What makes up the cytoskeleton? What is a microtubule? What is a microfilament? What is an intermediate filament? How do intermediate filaments differ from microtubules and microfilaments? 4. What does the plus (1) end of a microtubule or microfilament mean? What does the minus (2) end mean? What are “motor proteins”? What does it mean to be “ 1 directed”? Name the family of 1 directed motor proteins. Name the family of 2 directed motor proteins. How do motor proteins carry cargo? 5. Describe the structure of a nuclear membrane. Why does the nuclear membrane have an elaborate pore structure? How big is the nucleus? What is the nucleolus? 6. What do ribosomes do? What are they made of? Where are they made? What do “40S” and “60S” descriptions of the major ribosomal subunits mean? 7. What distinguishes “rough ER” from “smooth ER”? What does each make? 8. What is the Golgi apparatus? What is the cis face? Trans face? What are all those vesicles doing hanging around the Golgi rims and cis and trans faces? What goes on inside the Golgi cisternae? 9. How big is a mitochondrion? Why do you suppose it has two membranes? Where is the electron transport chain? What do the mitochondria do? 10. What do lysosomes contain? What is the difference between a primary and a secondary lysosome? What goes on inside peroxisomes? 11. What do proteasomes do? How do they know which proteins to degrade? 12. Name four junctions between cells. Which ones involve connexons? Which ones join actin filaments from one cell to another? Which ones join intermediate filaments from one cell to another? APPENDIX 2.1.A1 SOME METHODS FOR STUDYING CELL STRUCTURE AND FUNCTION THE MICROSCOPE HAS REVOLUTIONIZED OUR UNDERSTANDING OF BIOLOGY The invention of the microscope literally opened up a new world view in biology. For the first time, we could look upon the microscopic world and we discovered that it was full of animated objects. Robert Hooke (16351703) first used the microscope to study biological material in 1665 and coined the term “cell” from the likeness of the empty cells in cork to the monks’ cells within a monastery. Antoine van Leeuwenhoek (16321723) was the first to observe live cells with hand-made microscopes that could magnify up to 500 times. He observed an astonishing array of cells from single-celled protists to red blood cells, bacteria, sperm, and muscle fibers. These observations, along with many others, prompted Theodor Schwann to enunciate the Cell Theory in 1839, based in part on conversations with Matthias Schleiden, whom Schwann did not credit. It contained three important elements: 1. The cell is the unit of structure, physiology, and organization in living things. 2. The cell is a distinct entity but is the building block of complicated organisms. 3. Cells form from cell-free material, as in crystallization. We now know that this last point is incorrect. Its first correction was famously uttered by the German pathologist, Rudolf Virchow, when he said “Omnis cellula e cellula”—all cells come from cells. The entire development of the cell theory was supported by a single advance—the light microscope. MICROSCOPIC RESOLUTION IS THE ABILITY TO DISTINGUISH BETWEEN TWO SEPARATED OBJECTS What do we mean when we say that red blood cells, erythrocytes, are invisible to the naked eye? Surely we can see blood. What we mean here is that we cannot make out the individual cells. The ability to distinguish two objects separated by a distance is called the resolution. Light produces a diffraction pattern around all objects. This pattern consists of areas of maximum and minimum intensity of light. The resolution of optical devices is determined by considering the diffraction pattern produced by a circular aperture of radius a. This circular aperture is called an Airy disk. The convention is that two apertures can be distinguished if the first maximum 113 114 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION structure of cells only after artificial preparation. Without functional studies, the activity going on in the structures remains elusive. Objective α Specimen Condensor Light FIGURE 2.1.A1.1 Angle of the cone of light passing through the specimen and incident to the objective. of the diffraction pattern from one pattern falls on the first minimum of the second. This is called the Rayleigh criterion. The diffraction pattern can be described analytically using Bessel functions, which we will not do here. The result is given by ½2:1:A1:1 Resolution 5 0:61 λ η sin α where λ is the wavelength of light, η is the refractive index of the medium between the specimen and the lens, and α is the angle of the cone of light that passes between the specimen and the lens (see Figure 2.1. A1.1). The expression η sin α is called the numerical aperture of the lens being used. This description of resolution is the inverse of what is ordinarily meant. Eqn [2.1.A1.1] defines resolution in terms of a distance. However, when we can resolve two objects that are close together, we ordinarily speak of a high resolution. According to Eqn [2.1.A1.1], the resolved distance is smallest (and the resolution is highest) when α is 90 and the refractive index is increased by placing oil between the specimen and lens. The refractive index of air is close to 1.0, whereas that of oil is 1.41.5. THE ELECTRON MICROSCOPE HAS ADVANCED OUR UNDERSTANDING OF CELL STRUCTURE Until the 1950s, the workhorse of cellular structure was the optical microscope. Our ideas of the structure and composition of cells were further enhanced by the electron microscope. This device allows tremendous magnification of cell components but suffers from the disadvantage that the cells must be fixed and stained before viewing their structures because the electron microscope views specimens in a vacuum. Otherwise, the incident electron beam would be scattered by air molecules. The electron microscope illuminates the The electron microscope can achieve much higher resolution (it can resolve objects that are separated by a smaller distance) than the optical microscope because the wavelength of electrons is so much shorter than the wavelength of visible light. The fact that electrons have a wavelength at all was an astounding discovery. The first clue to the wavelength of electrons was the Compton effect, reported by Arthur H. Compton in 1922. Compton found that incident X-rays were scattered by a carbon target, subsequently shifting the X-rays to a lower wavelength and dislodging an electron from the crystal. The dislodged electron has a momentum. To describe this collision of light with electrons, it was necessary to postulate that the incident light possessed momentum, in order to preserve the conservation of momentum theorem of physics. Compton produced a quantum-mechanical analysis of the scattering that differed from classical explanations, but agreed with the experimental observations. The photoelectric effect and Compton effects showed that light had distinctive particle-like aspects. It was already known, from interference and diffraction experiments, that light also had distinctive wave-like aspects. Louis de Broglie in 1924 postulated that the wave particle duality of photons might also apply to particles; if photons have momentum, particles such as the electrons ought to have a wavelength. From the Compton experiments, the wavelength would be given by ½2:1:A1:2 λ5 h mv where h is Planck’s constant 5 6.625 3 10234 J s. The wave nature of electrons was confirmed in 1927 by Davisson and Germer, who exposed a single crystal of nickel to electrons having 54 electron volts of kinetic energy. They observed an electron diffraction pattern that confirmed de Broglie’s relation. The short wavelength of electrons opened the possibility of a microscope with astounding resolution. The electron microscope, invented in the 1930s, was first applied to living tissues by Albert Claude, Keith Porter, and George Palade in the late 1940s and 1950s. The resolution of the electron microscope is determined not only by the wavelength of the incident radiation, which in this case is a beam of electrons, but also by the numerical aperture. Although theoretically the resolution of the electron microscope should be close to 0.002 nm, in principle it is much larger than this because the inherent properties of magnetic lenses limit the aperture angle to about 0.5 . Because biological specimens lack inherent contrast, the practical resolution is further reduced to about 12 nm. Nevertheless, this is a marked improvement over the optical microscope. The optical microscope was useful to a magnification of about 1000 3 ; the electron microscope could attain more than a 100-fold better magnification. Cell St ructur e SUBCELLULAR FRACTIONATION ALLOWS STUDIES OF ISOLATED ORGANELLE BUT REQUIRES DISRUPTION OF CELL FUNCTION AND STRUCTURE Although the optical microscope and electron microscope provided keen insights into the structure of living things, the function of the structures could not be directly investigated. In the late 1940s and 1950s, Albert Claude and Christian de Duve developed methods for subcellular fractionation for separating cells into their component parts and elucidating the function of the cellular constituents. This method disrupted the cells and separated the parts by differential centrifugation. DIFFERENTIAL CENTRIFUGATION PRODUCES ENRICHED FRACTIONS OF SUBCELLULAR ORGANELLES The first step in subcellular fractionation is the disruption of the cell into its component subcellular organelles. This process usually uses homogenization, and its aim is to break the plasma membrane that delimits the cell, thereby releasing the cellular contents, without damaging those contents. Cells can be homogenized by sonication (exposure to high-frequency sound waves), shearing the cells between two surfaces such as a Teflon mortar and glass pestle, or placing the cells in a high-speed blender. These treatments break the cell membrane and leave the remaining parts of the cells relatively intact. These relatively harsh treatments obviously disrupt some of the normal relationships between parts of the cell, and may scramble normal constituents of cells. The resulting mixture, the homogenate, can be separated into its component parts on the basis of their size and density (see Figures 2.1.A1.2 and 2.1.A1.3). Following lysis of the cells, the homogenate is placed in a tube in a centrifuge and spun at relatively low speed (1000 3 gravity) for a short time (1020 minutes). The centrifugation causes materials to move away from the axis of centrifugation. When the particles reach the bottom of the tube, they form a pellet. This process is called sedimentation. Particles that sediment quickly reach a terminal velocity in the centrifuge tube. At this terminal velocity, the frictional drag on the particle provides the acceleration necessary to keep the particle in approximate uniform circular motion (see below for an explanation of the forces during sedimentation). The frictional drag depends on the density of the particle, viscosity of the medium, and speed of centrifugation. The first centrifugation, at low speed and short times, sediments unbroken cells, and the nuclei because these Separate supernatant and pellet Homogenization 1 4 2 Supernatant 5 Pellet Centrifugation at higher speed Centrifugation at low speed 6 3 Rotor FIGURE 2.1.A1.2 Separation of subcellular organelles by differential centrifugation. Whole tissue is first homogenized, which disrupts cell membranes and releases subcellular organelles. The homogenate is then centrifuged to separate out particles on the basis of their sedimentation. In general, large particles sediment with small centrifugal forces and smaller particles require larger forces. Successive centrifugation at progressively faster revolutions per minute (RPM) separates organelles on the basis of their sedimentation characteristics. 115 116 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 2 Separate supernatant and pellet 1 Low-speed centrifugation 3 Medium-speed centrifugation 4 Separate supernatant and pellet 5 6 Separate supernatant and pellet High-speed centrifugation Homogenate Pellet contains whole cells, nuclei Pellet contains mitochondria, lysosomes Pellet contains microsomes FIGURE 2.1.A1.3 Differential centrifugation. The first slow speed spin causes large and heavy particles to sediment. These are separated from the particles that do not sediment. Successively faster centrifugations cause progressively smaller and lighter particles to sediment. are the largest and heaviest of subcellular structures. Therefore, a fraction of the homogenate is produced that is enriched in nuclei. The supernatant fraction, the suspension that lies above the first pellet, is cleared of nuclei but still contains other subcellular particles. Successive centrifugations at higher speeds and longer times produce fractions enriched in the mitochondria, lysosomes, and peroxisomes. Further centrifugation at still higher speeds sediments the broken fragments of the plasma membrane and endoplasmic reticulum. De Duve identified these subcellular fractions through the use of marker enzymes. The basic idea here is that each of the subcellular structures has a unique biochemical composition that enables their unique biochemical function within the cell. Part of their biochemical composition is their component enzymes. By measuring the distribution of an enzyme, the scientist can track the distribution of the particles that contain it. Markers for the mitochondria, for example, include cytochrome C oxidase and succinate dehydrogenase, among others. By measuring cytochrome C oxidase in the various fractions, one can estimate the amount of mitochondria in those fractions. De Duve deduced the existence of the lysosome on the basis of an enzyme that distributed itself differently from all other known markers. DENSITY GRADIENT CENTRIFUGATION ENHANCES PURITY OF THE FRACTIONS The centrifuge is a crude instrument for the separation of subcellular fractions because of the way in which it separates the different subcellular organelles. Differential centrifugation can produce fractions enriched in one particle or another, but the fractions are not pure. This is due to the fact that sedimentation occurs over the considerable length of the centrifuge tube. When the tube is spun, particles throughout the tube are subject to centrifugal forces that cause them to sediment. Although heavier particles sediment more quickly, the heavier particles at the top of the tube have much further to travel than those near the bottom. By the time all of the heavy particles sediment, some of the lighter particles near the bottom of the tube, or at the middle, also sediment. Therefore, pure fractions of subcellular particles cannot be achieved easily by simple differential centrifugation. Further purification can be achieved by using density gradient centrifugation. In this method, subcellular organelles are separated by centrifugation through a gradient of a dense substance, such as sucrose. In velocity centrifugation, the material to be separated is layered on top of a sucrose gradient, and then centrifuged. Particles of different sizes and density sediment through the gradient at different rates moving as discrete bands. At the end of the centrifugation, the different layers consist of purified organelles, and they can be collected for further experiments. In equilibrium centrifugation, the density gradient is used to separate particles based on their buoyant density. Instead of being separated by their sedimentation velocity, particles will sediment until they reach a layer with the same density as the particles. At this point, sedimentation stops and the purified organelles can be collected at the equilibrium position. These methods of separating subcellular organelles are illustrated in Figure 2.1.A1.4. ANALYSIS OF CENTRIFUGATION SEPARATION CIRCULAR MOTION REQUIRES AN INWARD CENTRIPETAL FORCE Centrifugation typically involves spinning tubes of material at a constant angular velocity, except for the angular acceleration to that velocity, and the deceleration when the spin stops. The position of any particle in the tube at any time may be represented by the vector r, as shown in Figure 2.1.A1.5. Cell St ructur e Velocity sedimentation Centrifugation Equilibrium sedimentation Slow-sedimenting component Fast-sedimenting component Shallow sucrose gradient (5–20%) Centrifugation Low buoyant density component High buoyant density component Steep sucrose gradient (20–60%) FIGURE 2.1.A1.4 Comparison of velocity sedimentation and equilibrium sedimentation for the separation of subcellular particles. In velocity sedimentation, separation relies on different velocity of sedimentation through a shallow sucrose (or other material) gradient. The sucrose is added to prevent mixing by convection. In equilibrium sedimentation, subcellular particles sediment until they reach a zone of solution that matches their density. At equilibrium, the organelles distribute themselves over a narrow region of the tube corresponding to the density of the organelle. The organelles can then be harvested from a narrow band. v = dr dt r = i cos ωt + j sin ωt θ = ωt FIGURE 2.1.A1.5 A particle in uniform circular motion around a central pivot point. The angular displacement is a linear function of time. This vector makes the angle θ from an arbitrary zero reference, the x-axis, and this angle increases with time. The angular velocity is defined as dθ dt By integrating this we see easily that: ½2:1:A1:3 ω5 ½2:1:A1:4 The position vector indicating the location of the point relative to the center of rotation is thus given as - - - r 5 i cos ωt 1 j sin ωt where r is the position vector and i and j are unit vectors along the x-axis and y-axis, respectively. The velocity vector at any time is the derivative of this position vector: ~ r 5 2 i ω sin ωt 1 j ω cos ωt dt This velocity vector is orthogonal to the position vector as seen by the dot product: it is given as ½2:1:A1:6 - - ½2:1:A1:8 - V 5d r U V 5 2 ω sin ωt cos ωt 1 ω sin ωt cos ωt 5 0 ½2:1:A1:7 ~ a 5d V 5 2 i ω2 cos ωt 2 j ω2 sin ωt dt ~ r a 5 2 ω2~ We see here that the magnitude of the acceleration is ω2r and its direction is directly opposite that of the position vector (the negative sign in Eqn [2.1.A1.8]). (Compare Eqn [2.1.A1.8] to [2.1.A1.5] for the identity of r in a.) This is the centripetal acceleration. In order for the particle to remain in uniform circular motion, the velocity must be continuously bent toward the center. The acceleration vector is orthogonal to the velocity vector, so that all of the acceleration is used to change the direction of the velocity, and not its magnitude. CENTRIPETAL FORCE IN A SPINNING TUBE IS PROVIDED BY THE SOLVENT As described above, particles that are spinning in a rotor at a constant angular velocity, ω, are subjected to a centripetal acceleration given by ½2:1:A1:9 θ 5 ωt ½2:1:A1:5 This means that the velocity vector is oriented at 90 to the position vector, as shown in Figure 2.1.A1.5. The acceleration at any time is the derivative of the velocity: a 5 ω2 r where a is the acceleration, r is the radius, and ω is the angular velocity, in radians per second (52π 3 revolutions per second). This centripetal acceleration is the acceleration necessary to keep a particle rotating about the axis, at a distance r. If the actual force is less than this, the particle will move away from the axis of rotation. In the centrifuge, the centripetal acceleration is provided by collisions with solvent particles. The net force of these collisions under ordinary conditions (i.e., not in the centrifuge) adds to the force of gravity on the particle. In the centrifuge, the centripetal force necessary to keep the particle rotating at angular velocity ω at distance r from the axis is ½2:1:A1:10 Fc 5 ðmparticle 2 msolution Þω2 r Here the mass of solution is the mass of the volume of solution which is displaced by the particle and is 117 118 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION the origin of the buoyant force. Eqn [2.1.A1.10] can be rewritten as Fc 5 ðm 2 Vp ρÞω2 r ½2:1:A1:11 5 ðm 2 mvρÞω2 r The relative centrifugal field (RCF) is given as ω2 r ½2:1:A1:15 RCF 5 g 22 where g 5 981 cm s or 9.81 m s22. 5 mð1 2 vρÞω2 r where m is the mass of the particle, V is the partial specific volume, equal to 1/ρ for the particle, and ρ is the density of the solution. As pointed out above, this centripetal force is the force required to maintain an orbit at angular velocity ω at a distance r from the axis of revolution. The source of this force is the collisions with solvent molecules, which have a net direction toward the axis of revolution only when there is a net velocity of the particle in the opposite direction. That is, a particle more dense than the solution will move toward the bottom of the tube (away from the axis of revolution) with some velocity relative to the tube. Because it is moving through the solution, it experiences a drag force that is proportional to the velocity. This drag force on a particle moving outward from the axis of revolution is given by ½2:1:A1:12 FD 5 2βV where β is a frictional coefficient. This equation says that the drag force is proportional to the velocity but opposite in direction. A terminal velocity, v, is reached when there is a balance between Fc and FD. From Eqns [2.1.A1.11] and [2.1.A1.12], this is ½2:1:A1:13 mð1 2 vρÞω r 5 βv 2 THE MAGNITUDE OF THE CENTRIPETAL FORCE CAN BE EXPRESSED AS RELATIVE CENTRIFUGAL FORCE The frame of reference for the analysis presented so far is the nonrotating frame of the laboratory. One can also view the situation from the accelerated, rotating frame of reference of the solution within the rotor. In this case, there is an apparent force on every particle, the centrifugal force, which is equal but opposite to the centripetal force, which appears to drive particles heavier than the solution to the bottom of the tube. This motion within the tube is opposed by the frictional force, given above, which is opposite to the direction of motion. The terminal velocity is reached when the centrifugal force is equal to the frictional force, as given by Eqn [2.1.A1.13]. Note that the centrifugal force is a fictional force which must be invented in order to apply Newton’s laws in a uniformly accelerated frame of reference. The centrifugal force is just that force given by Eqns [2.1.A1.10] and [2.1.A1.11]. Usually the centrifugal force is given in multiples of g, the acceleration due to gravity: ½2:1:A1:14 a5 ω2 r g g THE VELOCITY OF SEDIMENTATION IS MEASURED IN SVEDBERGS OR S UNITS The terminal velocity within the tube can be written as dr ½2:1:A1:16 V5 dt where r is the distance from a sedimenting particle and the center of rotation. Insertion of this definition into Eqn [2.1.A1.13] and rearranging, we obtain: 2 3 dr 6 7 mð1 2 vρÞ 5 β 4 dt ½2:1:A1:17 5 ω2 r The term in the brackets on the right-hand side of the equation defines the sedimentation coefficient, s. The sedimentation coefficient is the rate of sedimentation per unit of centrifugal force. Sedimentation coefficients are usually of the order of 10213 s. They are reported in Svedberg units, where 1 Svedberg 5 10213 s. This unit is named after T. Svedberg, an early pioneer in the design of centrifuges and their use in investigations of biological material. The sedimentation coefficient is obtained experimentally by plotting the logarithm of the radius of the maximum of the concentration profile against the time. To see this, consider again the definition of the sedimentation coefficient: dt ½2:1:A1:18 s 5 dr ω2 r multiplying through by ω2. dr ω s 5 dt r 2 ½2:1:A1:19 5 d ln r dt Thus the slope of a plot of ln r against t will give ω2s. The sedimentation coefficient varies with the concentration of solute, usually decreasing as the total protein concentration increases. The sedimentation coefficient at infinite dilution, s0, is usually obtained by extrapolation of plots of 1/s against protein concentration to zero protein concentration. Sedimentation information can sometimes inform us about molecular dimensions. Eqn [2.1.A1.17] can be rearranged to give ½2:1:A1:20 mð1 2 vρÞ 5β s Cell St ructur e where β is the frictional coefficient described in Eqn [1.2.A1.12]. The Stokes equation gives the frictional coefficient as fluorescence microscopy and confocal microscopy. The main advantage in these techniques is that intact and living cells can be investigated. ½2:1:A1:21 In fluorescence microscopy, specific components of the cells can be labeled by attaching a fluorophore to them. The fluorophore is a fluorescent molecule. It absorbs light at one wavelength and emits it at a second, lower wavelength. The location of the fluorescent molecule is achieved by illuminating the specimen at the excitation wavelength and using filters to collect the emitted light. For example, incubating cells with a fluorophore-tagged antibody directed against a specific protein allows the study of the distribution of the protein. This technique can be used with either living or fixed cells. In other experiments, native proteins expressed by cells can be "tagged" with a fluorescent protein, green fluorescent protein (GFP by incorporating the gene for GFP into the gene for the protein. Thus the location of the tagged protein can be followed by fluorescent microscopy. β 5 6πηas where η is the viscosity of the solution and as is the radius of the molecule, assuming spherical geometry. Thus the measurement of s and the knowledge of the molecular mass, m, and its partial molar volume allow estimate of its size. DENSITY GRADIENT CENTRIFUGATION Macromolecules or subcellular organelles can be sedimented through gradients of increasing density on a preparative scale to purify them. There are a variety of materials which can be used to prepare such density gradients, including CsCl, D2O, Ficoll, glycerol, sorbitol, sucrose, and percoll. The gradients usually are formed by mixing two limiting solutions in varying proportions in order to produce the desired gradient. Gradients can be discontinuous, in which solutions of varying densities are layered on top of each other manually, or continuous. Continuous gradients can have a variety of shapes, although linear gradients are most common. Materials will sediment until they reach a solution with the same density as the particles, and then they stop sedimenting. OTHER OPTICAL METHODS A number of other optical approaches have proven useful for the modern investigation of cell function. These include phase-contrast microscopy and differential interference microscopy. Both of these optical techniques use variation in the refractive index of cell structures, rather than variation in light absorption, to produce contrast between the structures. These images can be clarified by using video cameras and computerized image analysis and processing. Other techniques include Confocal microscopy allows for optical sectioning of live cells. Confocal microscopy refers to the idea that both incident and emitted light are in focus. In a bright field, light illuminates the specimen and the objective lens is moved to focus the light. However, out-of-focus light still reaches the detector. In confocal microscopy, an aperture is placed in front of the detector so that out-of-focus light is eliminated. The result is that the image is formed from a narrow plane of in-focus light. By moving the plane up and down, one can obtain a series of optical sections. By computer techniques, three-dimensional reconstructions of the object can be obtained. Because the image is obtained and stored digitally, the resulting image can be viewed from any perspective. Confocal microscopy can be combined with fluorescent probes to achieve outstanding detail of localization. 119 2.2 DNA and Protein Synthesis Learning Objectives G G G G G G G G G G G G G G G Compare genotype and phenotype Describe the components of a nucleotide Name the different nucleotides that comprise DNA and RNA Describe how the nucleotides are connected to form single-stranded DNA or RNA Describe how hydrogen bonding is useful in combining antiparallel nucleotide strands to form the double helix Explain why antiparallel strands require special arrangement for DNA replication List the ways RNA differs from DNA List the various kinds of RNA and their function in the cell Know the origin of the RNA polymerases responsible for formation of the various RNA classes Describe what is meant by “the genetic code” Define and contrast “transcription” and “translation” Distinguish among “response elements,” “intron,” and “exon” Describe histones and their postulated function in the structure of chromosomes Describe what is meant by the “histone code” Describe how DNA is methylated and how this methylation can be passed from parent to daughter cells DNA MAKES UP THE GENOME As described in Chapter 1.1, almost all cells in the body have the same amount and kind of DNA. The diversity of human cell forms derives from their expression of different parts of the DNA. The total DNA with its division into units, called genes, constitutes the genome. Expression of a gene means that the DNA that makes up the gene is used to direct the synthesis of a specific protein. As we will see in this chapter, genes associate with a host of proteins that regulate the expression of the genes. The human genome refers to the set of genes that are normally present in humans. 120 The DNA in human cells is organized into compact units called chromosomes, meaning “colored body,” which refers to their appearance in fixed and stained preparations. Each chromosome carries a defined set of genes that carries the instructions for making a set of proteins. Because each chromosome is paired, nearly every gene comes in pairs, but the two pairs are usually not identical. Paired genes carry the instructions for the synthesis of analogous materials, but they differ in the details. These alternate forms of the genes in a single person are called alleles. The set of alleles of a particular person is called the genotype. The set of proteins and other materials that the person actually makes, and which determine their outward appearance and behavior, is called the phenotype. Humans have 23 pairs of chromosomes. Two chromosomes determine the sex of the individual. These are the X and Y chromosomes. Persons with two X chromosomes are genotypic females; having one X and one Y makes a genotypic male. Because the Y chromosome is smaller than the X, some of the genes carried on the X chromosome are not paired. This is the one exception to the rule that all genes are paired. DNA CONSISTS OF TWO INTERTWINED SEQUENCES OF NUCLEOTIDES DNA IS BUILT FROM NUCLEOTIDES DNA stands for deoxyribonucleic acid. It is located primarily in the nucleus of cells but important parts are also present in the mitochondria. It is composed of a sequence of building blocks called nucleotides. These nucleotides come in two different types and four varieties. The types are the purines and pyrimidines. The purines in DNA are adenine and guanine, and the pyrimidines are thymine and cytosine. Each of these nucleotides consists of the base (adenine, guanine, thymine, and cytosine) linked to a sugar, deoxyribose, and phosphate. The chemical structures of the nucleotides are shown in Figure 2.2.1. NUCLEOTIDES ARE LINKED TOGETHER TO FORM A CHAIN These four bases are linked together to form a long sequence of nucleotides. The DNA is elongated by reacting a nucleotide triphosphate (with two more phosphates linked to the phosphate shown in Figure 2.2.1) on the 30 end of an existing chain. This reaction is catalyzed by an enzyme called DNA polymerase. This enzyme is involved in the replication of DNA, where two complete DNA strands are made from © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00011-2 DNA and P rot ein Sy nt hes is NH2 O H N 5′ End –O P O O Phosphate CH2 5 N O N N OH OH N –O NH2 3 2 Deoxyribose P 5 O 1 4 5′ End O CH2 3 Base 2 3′ End OH Deoxy guanosine monophosphate Deoxy cytidine monophosphate O NH2 H3C N OH –O P O 5′ End O CH2 5 N O 3 NH N OH –O N P 5′ End O CH2 5 O 1 4 O 1 4 3′ End OH N O 2 O 1 4 3 3′ End OH Deoxy adenosine monophosphate N O 2 3′ End OH Deoxy thymidine monophosphate FIGURE 2.2.1 Structures of the nucleoside monophosphates. one, using the original DNA as a template. The structure of single-stranded DNA is shown in Figure 2.2.2. The single-stranded DNA described in Figure 2.2.2 is just half of the story. In humans, DNA is normally present as double strands that are held together by hydrogen bonds, as shown in Figure 2.2.3. In doublestranded DNA, adenine on one strand pairs with thymine on the opposite strand and guanine on one strand pairs with cytosine on the opposite strand. The two strands have opposite polarity: the 50 end of one strand is opposite to the 30 end of the other. O –O P THE DNA TEMPLATE SETS THE SEQUENCE OF NUCLEOTIDES 0 DNA polymerase adds nucleotides to the 3 end, using a nucleotide triphosphate as a substrate. The base on the opposite strand determines which nucleotide is incorporated. Thus DNA polymerase replicates DNA on the basis of the DNA already present. The DNA strand unwinds to form two single strands. The DNA polymerase adds nucleotides on both strands to form two complete DNA double strands. The hydrogen bonding between nucleotides is crucial to the ability of DNA to N Guanine 5′ End O CH2 5 O N O N NH2 1 4 3 NH2 2 3′ End N N O Adenine 5′ End –O P O CH2 5 O N O N 1 4 3 HYDROGEN BONDING ALLOWS FOR DNA STABILITY WITH RAPID DISSOCIATION As discussed in Chapter 1.4, hydrogen bonds involve sharing of the positive H atom between two electronegative centers. It requires the right spatial separation and orientation of these centers, and has low dissociation energy. This allows the H-bond to form or break rapidly. However, many hydrogen bonds can stabilize large structures like DNA and proteins. H N OH 2 O H2C 3′ End Thymine NH O 5′ End P –O O CH2 5 O N O O 1 4 3 2 OH 3′ End NH2 DNA polymerase OH –O P O N OH O P O Cytosine OH O P O 5′ End O CH2 5 N O 4 O 1 3 2 3′ End OH FIGURE 2.2.2 Arrangement of bases in single-stranded DNA. The phosphatedeoxyribose part of the nucleotide triphosphates forms a backbone of alternating phosphate and deoxyribose molecules. Attached to this backbone are the four bases: guanine, adenine, thymine, and cytosine. The 30 and 50 ends of the strand derive from the numbering of the ribose carbons. 121 122 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION H 3′ End OH Hydrogen bonds N H C O –O OH 5′ End P O CH2 N N O O H N H2C O N N H N H3C N N H O O N O N O CH2 T O O N O– O H N H P P G O O –O O H A O H2C O N 5′ End P O– HO N Hydrogen bonds O OH 3′ End FIGURE 2.2.3 Pairing of bases in double-stranded DNA. serve as a template for its own replication and for the synthesis of RNA. THE DOUBLE HELIX IS SUPERCOILED IN CHROMOSOMES The two DNA strands intertwine around each other to form a double helix. This helix can be further wrapped around proteins called histones. The complex of DNA and its associated proteins is called chromatin. The complex resembles “beads on a string” that are visible in electron micrographs (see Figure 2.2.9). The entire structure can be further coiled and coiled again to form the chromosomes. All of the DNA becomes condensed like this when the cell divides. Between divisions, the chromosomes partially “unravel” to form a less dense form of chromatin that is the working state of DNA. THE DOUBLE HELIX POSES SPECIAL CHALLENGES FOR DNA REPLICATION The replication of DNA by adding nucleotides only to the 30 end of the growing DNA strand poses a problem for the duplication of double-stranded DNA because the strands are of opposite polarity. This leads to the replication of DNA in spurts, as shown in Figure 2.2.4 and described in the legend. RNA IS CLOSELY RELATED TO DNA RNA is ribonucleic acid. Structurally, it is very similar to DNA but it differs in several ways. First, the sugar in the backbone in DNA is deoxyribose, whereas in RNA it is ribose. Second, the nucleotide base thymine in DNA is replaced by uracil in RNA. Uracil hydrogen bonds with adenine, taking the place of thymine. Third, RNA in eukaryotic cells is single stranded. Fourth, RNA is not replicated. All of the RNA is produced from DNA using DNA as a template. There are different kinds of RNA: G G G G G mRNA: “messenger” RNA tRNA: “transfer” RNA rRNA: “ribosomal” RNA snRNA, scRNA: “small nuclear” and “small cytoplasmic” RNA Mitochondrial RNA. MESSENGER RNA CARRIES THE INSTRUCTIONS FOR MAKING PROTEINS mRNA is “messenger” RNA. mRNA is synthesized in the nucleus using the nucleotide sequence of DNA as a template. This process requires nucleotide triphosphates as substrates and is catalyzed by the enzyme RNA polymerase II. The process of making mRNA from DNA is called transcription, and it occurs in the nucleus. The mRNA directs the synthesis of proteins, which occurs in the cytoplasm. mRNA formed in the nucleus is transported out of the nucleus and into the cytoplasm where it attaches to the ribosomes. Proteins are assembled on the ribosomes using the mRNA nucleotide sequence as a guide. Thus mRNA carries a “message” from the nucleus to the cytoplasm. The message is encoded in the nucleotide sequence of the mRNA, which is complementary to the nucleotide sequence of the DNA that served as a template for synthesizing the mRNA. Making proteins from mRNA is called translation. RIBOSOMAL RNA IS ASSEMBLED IN THE NUCLEOLUS FROM A DNA TEMPLATE As discussed in Chapter 2.1, ribosomes are complex structures comprised of ribosomal RNA (rRNA) and a number of proteins. RNA polymerase I makes rRNA form a large loop of DNA called the nucleolar organizer region. The rRNAs then combine with proteins that migrate into the nucleolus from the cytoplasm to form the small and large ribosomal subunits. These ribosomal DNA and P rot ein Sy nt hes is DNA polymerase adds nucleotides to the 3′ end of the leading strand The leading strand is 3′ to 5′ 3′ DNA helicase unwinds the two strands 5′ DNA polymerase cannot fill in the entire complement—makes Okazaki fragments Complementary DNA double helix; the two strands have opposite polarity 5′ 3′ 5′ 3′ 3′ 5′ 3′ Okazaki fragments DNA primase adds short RNA strand to prime DNA polymerase 5′ The lagging strand is 5′ to 3′ DNA ligase joins Okazaki fragments FIGURE 2.2.4 Replication fork of DNA. The double-stranded DNA consists of two strands of opposite polarity. DNA helicase unwinds the strands, forming two single-stranded DNAs. The leading strand is 30 to 50 so its complementary strand being newly synthesized is 50 to 30 and DNA polymerase adds nucleotides to the 30 end as they become available from the helicase. The lagging strand is 50 to 30 and its complementary strand is 30 to 50 . DNA primase adds a short RNA strand that primes the DNA polymerase. DNA polymerase then makes the complement by progressing away from the replication fork. While it is making DNA, the helicase unwinds more DNA so another DNA polymerase starts replication nearer the replication fork. In this way, the lagging strand is filled in with Okazaki fragments that bind to the lagging strand but are not connected. DNA ligase connects the Okazaki fragments to complete replication of the lagging strand. subunits are then transferred to the cytoplasm where they are fully assembled to form an 80S functional ribosome and become protein factories. TRANSFER RNA COVALENTLY BINDS AMINO ACIDS AND RECOGNIZES SPECIFIC REGIONS OF MRNA How does mRNA specify the sequence of amino acids in a protein? Which amino acid is to be incorporated into the protein is specified by a sequence of three nucleotides called a codon. The mRNA triplets do not directly recognize and specify the amino acids; they do so through the use of another kind of RNA called transfer RNA or tRNA. These remarkable molecules are adapters that can link with an amino acid and recognize the triplets of nucleotides on the mRNA, the codons. They do this by containing a sequence complementary to the codon: the anticodon. The function that maps triplets of nucleotides on the mRNA to specific amino acids is called the genetic code. Figure 2.2.5 shows the genetic code in look-up table format. The tRNA consists of a single strand of RNA from 70 to 90 nucleotides long that is held together by hydrogen bonding within nucleotides on the same chain. One end of the tRNA allows for covalent attachment of an amino acid. Another section of the tRNA contains a sequence of three nucleotides that forms the anticodon. Precursors to the tRNA are transcribed from DNA by RNA polymerase III. Another key in the formation of proteins is the attachment of amino acids to the specific tRNA. Specific enzymes called aminoacyl-tRNA synthetases couple the amino acid to the appropriate tRNA. There is a different synthetase for each amino acid. One attaches glycine to tRNAGly, another attaches alanine to tRNAAla, and so on. These synthetases must recognize both the amino acid and the tRNA that contain the right anticodon. The overall processing of RNA and protein synthesis is shown in Figure 2.2.6. Translation is shown in Figure 2.2.7. THE GENETIC CODE IS A SYSTEM PROPERTY The genetic code shown in Figure 2.2.5 lists an amino acid or other signal (such as STOP) for every triplet nucleotide in mRNA. The code is the function that maps the nucleotide sequence onto instructions for protein synthesis. That is, ½2:2:1 f fNi : fA; U; C; Ggg 5 fAj : fcontrol steps; amino acidsgg This describes the function that turns a set of Ni nucleotides on the mRNA (selected from adenosine, uracil, cytosine, and guanosine nucleotides) into a set of Aj amino acids. Where does this code reside? The code should not be confused with the message or the means of writing the message. Therefore, the code itself is not a 123 124 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Second position First position 5′ end Amino acid Third position 3′ end U C A G U Phe Phe Leu Leu Ser Ser Ser Ser Tyr Tyr STOP STOP Cys Cys STOP Trp U C A G C Leu Leu Leu Leu Pro Pro Pro Pro His His Gln Gln Arg Arg Arg Arg U C A G A Ile Ile Ile Met Thr Thr Thr Thr Asn Asn Lys Lys Ser Ser Arg Arg U C A G G Val Val Val Val Ala Ala Ala Ala Asp Asp Glu Glu Gly Gly Gly Gly U C A G ThreeOneletter code letter code Alanine Arginine Asparagine Aspartic acid Cysteine Glutamic acid Glutamine Glycine Histidine Isoleucine Leucine Lysine Methionine Phenylalanine Proline Serine Threonine Tryptophan Tyrosine Valine Ala Arg Asn Asp Cys Glu Gln Gly His Ile Leu Lys Met Phe Pro Ser Thr Trp Tyr Val A R N D C E Q G H I L K M F P S T W Y V Codons GCA GCC GCG GCU AGA AGG CGA CGC CGG CGU AAC AAU GAC GAG UGC UGU GAA GAG CAA CAG GGA GGC GGG GGU CAC CAU AUA AUC AUU UUA UUG CUA CUC CUG CUU AAA AAG AUG UUC UUU CCA CCC CCG CCU AGU AGC UCU UCC UCA UCG ACA ACC ACG ACU UGG UAC UAU GUA GUC GUG GUU FIGURE 2.2.5 The genetic code. Each amino acid that is incorporated into a protein is specified by a triplet sequence of nucleotides on the mRNA. These triplets are called codons. Which codons specify which amino acids is shown here in two formats. The left format shows which amino acids correspond to which codons given as a first, second, and third position. Here we use the Biochemists’ shorthand for RNA bases and amino acids, where U 5 uracil, C 5 cytosine, A 5 adenine, and G 5 guanosine, and each amino acid is given by its three-letter shorthand designation, where Phe 5 phenylalanine, Ser 5 serine, Tyr 5 tyrosine, Cys 5 cysteine, Leu 5 leucine, Pro 5 proline, His 5 histidine, Arg 5 arginine, Gln 5 glutamine, Ile 5 isoleucine, Thr 5 threonine, Asn 5 asparagine, Lys 5 lysine, Met 5 methionine, Val 5 valine, Ala 5 alanine, Asp 5 aspartic acid, Gly 5 glycine, and Glu 5 glutamic acid. The right format lists the amino acids together with their three-letter designation and single-letter designation, with a list of the codons that specify them. Replication: DNA polymerase DNA RNA polymerase III in the nucleus RNA polymerase I in the nucleolus Pre-tRNA Transcription: RNA polymerase II in the nucleus Pre-mRNA (nucleus) 45S pre-rRNA snRNA 5S rRNA Processing and splicing rRNA 28S Ribosomal 18S proteins 5.8S mRNA: sequence of codons Cytoplasm Ribosomes 40S with 18S rRNA 60S with 5S, 5.8S, and 28S rRNA tRNA: anticodon Aminoacyl-tRNA synthetases Amino acids Translation: free in the cytoplasm or on the ER Proteins FIGURE 2.2.6 Processing of RNA and DNA. Replication of DNA is accomplished by DNA polymerase using the original DNA as a template. Messenger RNA is synthesized in the nucleus as a precursor that is processed to form the final mRNA. The synthesis of mRNA is called transcription and is accomplished by RNA polymerase II. The final mRNA travels to the cytoplasm where it binds to ribosomes. The ribosomal subunits are formed in the nucleolus from proteins and ribosomal RNA that is made as a precursor and cut into the final rRNA strands. rRNA is made from DNA by RNA polymerase I. The mRNA directs the sequential addition of amino acids to form proteins in a process called translation. This requires tRNA, made from DNA by RNA polymerase III. DNA and P rot ein Sy nt hes is Growing peptide chain is attached at the P site Next charged tRNA binds to the A site Peptide bond forms and peptide chain moves to new carboxy terminus AA Trp COOH Aminoacyl tRNA NH2 Ala Trp Phe Met Met Phe Phe P site His Met NH2 Ala Trp His NH2 Ala Trp His Trp COOH Ribosome moves down one codon Trp ACC E site A site UAC AAG GCACAUGUUCUGGCAGAGGCACAUGU 5′ End mRNA 3′ End UAC AAG ACC GCACAUGUUCUGGCAGA GGCACAUGU 5′ End mRNA A AG ACC GCACAUGUUCUGGCAGAGGCACAUGU 5′ End 3′ End mRNA 3′ End tRNA on E site leaves UAC FIGURE 2.2.7 Elongation of a growing polypeptide chain. The mRNA binds to a ribosome consisting of some 82 proteins and 4 separate rRNA strands. There are 3 tRNA binding sites: an aminoacyl or A site, a peptidyl or P site, and an exit or E site. The aminoacyl tRNA is escorted to the ribosome by an elongation factor that hydrolyzes GTP. The scheme begins with a short polypeptide bound to a tRNA at the peptidyl site along with the tRNAMet that remains at the exit site. The next aminoacyl tRNA binds to the aminoacyl site; in this case it is tRNATrp that is charged with its amino acid. The peptide bond is formed between the peptide and its next amino acid on the carboxy terminus. The ribosome shifts over one codon; the tRNA at the exit site leaves and the former occupant of the peptidyl site now occupies the exit site. The peptide now occupies the peptidyl site one codon further along the mRNA. property of the mRNA. The DNA itself also does not contain the code, as its message has no meaning without the tRNA. The tRNA is synthesized from other parts of the DNA that are not directly transcribed for protein synthesis. Is the code in the tRNA that links the triplets of mRNA to a specific amino acid? Or is the code in the aminoacyl-tRNA synthetases, the enzymes that couple the amino acid to the specific tRNA? If the code is in the aminoacyl-tRNA synthetases, then the code is in some sense also in the DNA because the DNA directs the synthesis of the aminoacyl-tRNA synthetases! But the DNA does not “make” proteins; mRNA does not “make” proteins. The proper proteins are synthesized with DNA as the store of information, mRNA carrying the message, tRNA converting, in a single step, the nucleotide information into the protein information, and preexisting proteins catalyzing the entire series of events. The code itself is an emergent property. The genetic code does not exist in any single component of the mechanism for making proteins. It is not “in” the DNA, or the mRNA or the tRNA or the tRNA synthetases. Rather, it is a system property that emerges from the interactions of all of these parts. REGULATION OF DNA TRANSCRIPTION DEFINES THE CELL TYPE The differentiation of cell types produces the wide spectrum of cell types in the body, but our understanding of the process is still rudimentary: we do not know how to change one cell type into another. We do know, however, that the hallmark of differentiation is selective expression of the cell’s DNA. Selective expression of DNA as proteins requires selective transcription. Initiation of transcription by RNA polymerase II requires a number of specific proteins called transcription factors. These come in two flavors: some are required for activity at all genes and are therefore called general transcription factors. Other transcription factors bind to DNA sequences that control the expression of specific genes. The overall process is shown in Figure 2.2.8. Most genes contain both transcribed and untranscribed regions. On the 30 end of the DNA gene is a specialized sequence called the promoter that helps regulate gene expression. This region contains a TATAA sequence (called a TATA box) some 2530 nucleotides upstream from the initiation site. Initiation begins when a transcription factor TFIID (transcription factor polymerase II) binds to the DNA TATAA sequence. Some promoters do not contain a TATA box, but instead contain an initiator sequence. Nevertheless, TFIID is involved in initiation of transcription even on promoters that lack the TATA box. In addition to these general transcription factors, there are a number of transcription factors that may act as enhancers or repressors of gene expression. These factors interact with specific sites on the DNA that are generally further away from the gene than the promoter region and act as regulatory elements for gene transcription. The signals that turn on or turn off the production of transcription factors ultimately determine the phenotypic fate of cells. How these transcription factors work is still being investigated. An example of this is the 125 126 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Regulatory sites Promoter Introns DNA strand 5′ End 3′ End Exons Primary RNA transcript 5′ AAAAAA3′ Cap Poly A tail Splicing 5′ AAAAAA mRNA Translation N C Protein product Posttranslational modification S S Processed protein FIGURE 2.2.8 Overall processes of transcription and translation. The gene has a promoter that often contains a TATAA sequence for initiation of transcription by RNA polymerase II, several exons (expressed or coding regions of the DNA sequence) and introns (intervening or noncoding regions), and regions that bind transcription factors to either enhance or repress expression. The primary gene transcript contains the sequence of bases complementary to the introns and exons. This primary RNA transcript is then further processed on large complexes called spliceosomes, composed of protein and small nuclear RNAs or snRNA. The spliceosomes remove the introns by clipping them out and then splice together the remaining exons to form a sequential mRNA. This mRNA is then translated in the cytosol to a primary protein product which may then require further processing. steroid hormone receptors. Steroid hormones and similar materials like vitamin D bind to receptor proteins that in turn bind to specific sequences on the DNA called response elements. Binding of the receptor proteins then activates gene expression. A number of accessory proteins are required for this process. THE HISTONE CODE PROVIDES ANOTHER LEVEL OF REGULATION OF GENE TRANSCRIPTION As described earlier, the double helix of DNA winds around specific proteins called histones. These histones form the basic unit of chromatin called the nucleosome. A schematic of this structure is shown in Figure 2.2.9. When wrapped up in this way, DNA is inaccessible to RNA polymerase II and so cannot be transcribed to form mRNA. Most cells sequester away large portions of their DNA in this way. In order to express DNA, cells must unwrap it from the chromatin. Determining which sections of DNA should be unwrapped is the first step in regulating gene transcription. This is accomplished through covalent modifications of the histones. Covalent modification of histones after synthesis is called posttranslational modification. Histones undergo G G acetylation of lysine and arginine amino acids in the histones; methylation of lysine and arginine amino acids; G G phosphorylation of serine and threonine amino acids in the histones; ubiquitinylation of lysines. All of these modifications are accomplished by enzymes that must themselves be regulated. These enzymes possess histone acetyltransferase (HAT) activity, histone deacetylation (HDAC) activity, histone methyltransferase (HMT) activity, and histone kinase activity. The function of these enzymes is evident from their names: HAT adds acetyl groups to the histones; HDAC removes them; HMT adds methyl groups; and histone kinase phosphorylates the histones. Figure 2.2.10 summarizes the various known modifications of the core histones. These posttranslational modifications of the histones have functional consequences. For example, the combination of acetylation at lysine at position 8 on H4 and lysine at position 14 on H3 with phosphorylation of serine at position 10 on H3 is associated with transcription. Trimethylation of lysine at position 9 on H3 with the lack of acetylation of H3 and H4 correlates with transcriptional repression. These observations have led to the hypothesis that gene expression is regulated in part by a “histone code.” This hypothesis requires two components: 1. Specific enzymes write the code by adding or removing modifications at specific sites in the histones. 2. Other proteins recognize the histone markers and interact with histones and other factors to mediate functional effects. DNA and P rot ein Sy nt hes is Sequence-specific DNA-binding protein H2A H4 H3 H1 Nucleosome DNA strand H2B 30 nm DNA fiber H3 FIGURE 2.2.9 Schematic diagram of nucleosome structure and 30-nm DNA fiber. DNA is wrapped around a complex of eight histone proteins: a tetramer of H3 and H4 and two dimers of H2A and H2B, to form the nucleosome. Strands of DNA bound to nucleosomes resemble beads on a string. The interaction of DNA with the nucleosomes is altered by H1, a linker histone, that enables the nucleosomes to condense to form a 30-nm DNA strand. H1 interacts with both the histones in the core and DNA. The nucleosome contains 147 base pairs of DNA wrapped nearly twice around the core histones. A short “linker” of 1060 base pairs separates each nucleosome from its neighbor. The linear sequence of nucleosomes forms “beads on a string” in electron micrographs. This becomes more highly condensed to form 30-nm-thick fibers that are stabilized by the H1 histones. Ac P H2A Ub Ac-S G R G K Q G G K A R A ... A V L L P K K T E S H H K A K G K-COOH 1 5 119 Ac P Ac Ac H2B 5 12 MeP Ac AcP P 14 15 20 Me Ac Ac 120 Ac MeAc P Me Me NH2 -A R T K Q T A R K S T G G K A P R K Q L A S K A A R K S A ... G V K K ... E F K T D... 2 3 4 P H4 Ub NH2 - P E P V K S A P V P K K G S K K A I N K ... V K Y T S S K-COOH 1 H3 Ac Me 9 10 11 Ac Ac 14 17 18 Ac Ac 23 26 27 28 36 79 Me Ac -S G R G K G G K G L G K G G A K R H R K V L R D N I Q G I T... 1 3 Ac P 5 8 12 16 Acetyl group added Phosphate group added 20 Me Ub Methyl group added Ubiquitin protein attached FIGURE 2.2.10 Posttranslational modifications of the core histones. The one-letter amino acid abbreviation follows that in Figure 1.3.5. The numbers below the amino acids refer to the number of amino acids in the sequence, starting from the amino terminus. Source: From C.L. Peterson and M.-A. Lanie, Histones and histone modifications, Current Biology 14:R546R551, 2004. The histone code may be a misnomer if it is viewed as being like the genetic code. In the genetic code, given triplets of nucleotides on mRNA always produce the same result, independent of which cell or tissue is being analyzed. The term “histone code” implies that a particular combination of histone modifications will always produce the same biological result. Evidence suggests that the same pattern of histone modifications can be interpreted differently by different cells depending on the gene and its cellular context. 127 128 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION DNA METHYLATION REPRESSES TRANSCRIPTION A methyl group can be added to the 5 position of cytosine to form 5 methyl cytosine by the action of DNMT, DNA methyltransferase, as shown in Figure 2.2.11. These enzymes come in two classes. DNMT1 is responsible for maintenance methylation, which methylates the new strand of recently replicated DNA, so that the methylation pattern is passed down from stem cells to daughter cells (see Figure 2.2.12). This explains the unidirectionality of most developmental processes. Stem cells become differentiated cells and the differentiated cells maintain their differentiation, partly through a pattern of DNA methylation. The second class of DNA methylation transferases, represented by DNMT3a and DNMT3b, is responsible for de novo methylation. Both the maintenance methylation enzymes and de novo methylation enzymes methylate cytosine in a CpG sequence in the DNA (see Figure 2.2.11). The consequence of DNA methylation is typically repression of transcription for the genes that are methylated. The methyl group does two things: it interferes with the binding of transcription factors that eventually recruit RNA polymerase II, and it also allows the binding of a set of proteins that specifically recognize the methylated DNA, and these proteins recruit histone deacetylases that modify the histones associated with the DNA. These actions result in a repressed transcription of the methylated parts of the DNA. Cytosine NH2 DNA methyltransferase (DNMT) Cytosine NH2 CH3 6 O -O P 5 1 4 3 2 5′ end O CH2 5 O O -O N O 3 O 2 P O CH2 5 CH2 5 O 1 4 3 O CpG -O N O 1 3 O 2 H N N O N N O N O 4 3′ end H N 6 N 5 1 4 3 2 5′ end O O O -O P O 1 4 3′ end N NH2 P O CH2 5 O Guanosine 3 3′ end N 1 4 2 N O NH2 Guanosine 2 3′ end O Methylated CpG FIGURE 2.2.11 DNA methylation. DNA is methylated at sites with the sequence CpG by DNA methyltransferase. It can also be demethylated, but the process is not simply a reversal of the methylation reaction. Demethylases form hydroxymethyl cytosine, which is then cut out and replaced with cytosine by DNA repair mechanisms. CH3 Parent strand T A C G T T G T A G A C G T A C CH3 Parent strand T A C G T T G T A G A C G T A C 3′ 5′ Nonmethylated CpG Methylation 3′ Methylated CpG 3′ A T G C A A G A T C T G C A T G Daughter strand CH3 T A C G T T G T A G A C G T A C 3′ 5′ 3′ 3′ 5′ DNA replication Not recognized by maintenance DNA methyl transferase A T G C A A G A T C T G C A T G Daughter strand CH3 Recognized by DNA methyl transferase Parent pattern of methylation is preserved in both daughters 5′ A T G C A A G A T C T G C A T G CH3 CH3 Parent strand T A C G T T G T A G A C G T A C Daughter strand T A C G T T G T A G A C G T A C 3′ 5′ 3′ 5′ Methylation 3′ 3′ A T G C A A G A T C T G C A T G Parent strand CH3 A T G C A A G A T C T G C A T G Daughter strand CH3 FIGURE 2.2.12 Maintenance methylation. The parent DNA has some CpG sequences that are methylated, and some that are not. These are determined by de novo methylation reactions catalyzed by DNA methyl transferase 3A and 3B (DNMT3a and DNMT3b). When the DNA is replicated, the daughter strands are not methylated. The maintenance DNA methyl transferase (DNMT1) recognizes the CpG sequence on the daughter strand that corresponds to the methylated CpG sequence on the parent strand, and then methylates the daughter strand cytosine. DNMT1 does not bind to the unmethylated CpG sequences and therefore does not methylated these. The result is that the methylated pattern of the parent strand is replicated in the methylated pattern of the two daughter strands. In this way, the pattern of gene repression in a differentiated cell line is passed onto its descendants. DNA and P rot ein Sy nt hes is SUMMARY Genetic information is stored in DNA in the nucleus and mitochondria of cells. DNA consists of two strands of nucleotides on a phosphodeoxyribose backbone. The two strands form a double helix that is stabilized by the formation of hydrogen bonds between nucleotide bases on the two strands. Replication of DNA is based on making strands complementary to the two strands produced when the double-stranded DNA is unwound. The four bases include two purines (adenine and guanine) and two pyrimidines (cytosine and thymine). They pair up as A:T and G:C, with two hydrogen bonds between A and T and three between G and C. DNA serves as a template for making a variety of RNA types: RNA polymerase I makes ribosomal RNA (rRNA) from nuclear DNA; RNA polymerase II makes messenger RNA (mRNA); RNA polymerase III makes transfer RNA (tRNA). All of these contribute to the synthesis of proteins and control of gene expression. Protein synthesis occurs on the ribosomes, using mRNA as a template for tRNA. The ribosomes themselves are large and complex structures consisting of rRNA and a number of proteins. The ribosomes consist of 60S and 40S subunits. The 60S subunit has three rRNAs (28S, 5S, and 5.8S) and 49 other proteins; the 40S subunit has an 18S rRNA and another 33 proteins. The ribosome binds to mRNA and provides a reaction site for the peptide bound to tRNA (P site), a second reaction site for the next amino acid covalently attached to its tRNA (A site), and a third site for the tRNA about to leave the ribosome (E site). The ribosome brings the peptide carboxyl terminal close to the amino terminal of the next amino acid, and then forms the peptide bond, simultaneously shifting the peptide from the P to the A site. The mRNA provides a sequence of nucleotides. Groups of three nucleotides form codons that are recognized by complementary nucleotide sequences on the tRNA—the anticodons. The specific binding of anticodon to codon allows the mRNA to determine the sequence of amino acids in the proteins being made. The attachment of specific amino acids to specific tRNAs, however, is assured by the tRNA synthetases that hook the amino acids onto tRNA. mRNA directs the synthesis of specific proteins by virtue of its sequence of codons. The set of proteins that are made determines the type of cell, because cell structure and activity derives from the kinds and amounts of proteins expressed by the cell. Cells have developed elaborate methods for determining what parts of the DNA are transcribed into mRNA. These methods involve transcription factors, enhancers and repressors, and the histone code. The DNA in cells is wrapped around a complex of histone proteins, forming a nucleosome. Modification of the histones allows specific sections of the DNA to be either used to make proteins or silenced. A second method for repression of gene transcription is methylation of cytosines in the DNA in sequences of CpG. These are originally methylated by de novo methylation, but the pattern of methylation can survive DNA replication through a maintenance methyltransferase. The result is a stable pattern of gene repression that survives proliferation. DNA methylation interacts with histone modification to determine which DNA sequences will be silenced. REVIEW QUESTIONS 1. What is the genotype? What is the phenotype? What is an allele? What is the usefulness in having two copies of each gene? 2. What are the parts of a nucleotide? Name the purine bases. Name the pyrimidine bases. 3. How are nucleotides linked together to form single-stranded DNA or RNA? What distinguishes the 50 end from the 30 end? What enzyme makes DNA from nucleotides? 4. What holds double-stranded DNA together? Why are hydrogen bonds useful? How many hydrogen bonds link guanine to cytosine? How many such bonds link adenine to thymine? 5. During DNA replication, what determines the sequence of DNA in the new strands? 6. Does DNA polymerase add nucleotides at the 50 end or the 30 end of the strand? What problem does this make for DNA replication? What is an Okazaki fragment? 7. How does RNA differ from DNA? What is mRNA? What is tRNA? What is rRNA? What RNA polymerase makes mRNA? Which makes tRNA? rRNA? 8. What is a codon? Is it on mRNA, DNA, or tRNA? What is an anticodon? 9. What couples tRNA with amino acids? Are these enzymes specific for the anticodon? 10. What is a ribosome? What is the A site? What is the P site? What is the E site? 11. What is the genetic code? Where is it located in the cell? 12. What is meant by “transcription”? What is meant by “translation”? 13. How are inactive portions of DNA locked up by the cell? How do they get unlocked? What is a “response element”? What is an “intron”? What is an “exon”? 14. What are histones? What promotes DNA binding? What promotes DNA unraveling from the histones? What is meant by “the histone code”? 15. What is DNA methylation? Where does it occur? What is the consequence of DNA methylation? How does DNA methylation pattern get passed on to daughter cells? 129 2.3 Protein Structure Learning Objectives G G G G G G G G G G G G G Draw the basic structure of an amino acid, specifying the amino group, carboxyl group, alpha carbon, and R group For each amino acid, tell whether it is polar, nonpolar, acidic, or basic Draw the reaction describing the formation of a peptide bond Describe what is meant by the primary structure of a protein Describe what is meant by the secondary, tertiary, and quaternary structure of a protein Describe the four kinds of noncovalent interactions that stabilize protein structure Define posttranslational modification List four major classes of posttranslational modification List the major kinds of chemical modification of proteins List the amino acids involved in glycosylation and its overall function in proteins Describe the consequence of gamma carboxylation of proteins and name the vitamin involved Describe two distinct ways of varying the activity of proteins in cells Describe three distinct ways membrane proteins can be anchored in the membrane by covalent modification AMINO ACIDS MAKE UP PROTEINS In Chapter 2.2, we described how proteins are made on ribosomes by linking amino acids together in long chains. Which amino acids make up a specific protein, and in which order, is determined by the sequence of triplet nucleotide codons residing on the mRNA, which in turn is produced on a DNA template in the nucleus. Because these constituent amino acids determine the detailed shape of the protein surface, we should find the key to protein activity in the three-dimensional arrangement of these amino acids. The general chemical structure of the amino acids is shown in Figure 2.3.1. The amino acids are named for the two functional groups each of them possesses: an amino group (NH) and a carboxylic acid group (COOH). These two groups have widely different reactivities. The amino group is basic and will be positively charged at neutral pH, forming an ammonium ion (NH31). The carboxyl group is acidic and will form a negatively charged carboxyl ion (COO2) at neutral pH. Thus at neutral pH, many amino acids will have a positive charge on the amino end and a negative charge on the carboxyl end. Compounds possessing both positive and negative charges simultaneously are called zwitterions, as shown in Figure 2.3.2. Although the amino and carboxyl groups are important for the structure of proteins, the tremendous variety of structure and function of proteins is produced by the R groups (R stands for “residue” and refers to the part of the amino acid other than the amino or carboxyl group covalently bonded to the α carbon). There are 20 different amino acids differing only in their R groups. The R groups confer properties on the amino acids that classify them as nonpolar, polar, basic, or acidic. Each of these groups is shown in Figures 2.3.32.3.5. Biochemists often use a three-letter abbreviation for the amino acids or a oneletter code for a further shorthand. These abbreviations are shown under each structural formula (see Figure 2.3.6). As their name implies, the nonpolar amino acids contain side groups that are nonpolar and therefore not attracted to water. Because water is so attracted to itself, these groups naturally are repulsed by the water and are said to be hydrophobic, or water hating. For glycine and alanine, the nonpolar side groups are so small that the amino acid has little preference for either a watery (hydrophilic) or nonwatery (hydrophobic) environment. Phenylalanine and tryptophan are highly hydrophobic and seek environments away from water. HYDROPHOBIC INTERACTIONS CAN BE ASSESSED FROM THE PARTITION COEFFICIENT We can estimate the strength of hydrophobic interactions by measuring the partitioning of a material between the water phase and an immiscible solvent such as N-octanol or ethyl acetate, whose properties may be regarded as being similar to the interior of protein away from the water phase. The partition coefficient is ½2:3:1 ks 5 ½xorganic phase ½xwater phase 130 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00012-4 Prot ein St ructur e L-amino D-amino acid where the concentrations are the equilibrium concentrations, as shown in Figure 2.3.7. This equilibrium condition obeys Eqn [1.7.32]: acid Amino group NH2 NH2 ½2:3:2 R H C H R C Alpha carbon C O C OH O HO Carboxyl group FIGURE 2.3.1 Chemical structure of the amino acids. Amino acids consist of a central carbon, the alpha carbon, to which are bonded a carboxyl group, an amino group, a hydrogen atom, and a variable group, called R. Because the alpha carbon has four different groups bonded to it, it is an asymmetric carbon that is capable of stereoisomerism. All of the amino acids can be made in L- and D-form (L and D originally described the ability of compounds to rotate the plane of polarized light; L for levo, meaning “left,” and D for dextro meaning “right.” The symbols L and D for the amino acids do not actually refer to the direction of rotation of the plane of polarized light). All naturally occurring amino acids in higher organisms are the L-form. The L and D forms are mirror images of each other and are not superimposable. + NH3 R H C C O – O FIGURE 2.3.2 Zwitterion form of an amino acid. At neutral pH (pH 5 7), the carboxyl group is dissociated to form the anionic COO2 group, and the amino group binds a H1 ion to form NH31. Thus both ends of the amino acid are charged but with opposite polarity. O OH C O OH CH2 C CH2 H2N C H CH2 O H2N C OH Aspartic acid (Asp) D C H O C OH Glutamic acid (Glu) E FIGURE 2.3.3 The acidic amino acids aspartic acid and glutamic acid. The R group in both cases contains another carboxylic acid group that dissociates to form an H1 ion and a negatively charged carboxyl group (COO2). The presence of this group classifies these amino acids as acidic amino acids. The negatively charged carboxyl group imparts a negative charge to the region of any protein that contains these groups. ΔG0 5 2RT ln Keq ½xorganic phase ΔG0T 5 2RT ln ½xwater phase where ΔG0T is the standard Gibbs free energy change for transfer of substance X from water to the organic phase. Although this can represent the strength of hydrophobic interaction of a small molecule, it cannot reliably predict the behavior of a polymer of the material or of the material when it participates in a heteropolymer such as a protein. This is true for amino acids in particular because some of their functional groups that determine the octanol/water partition are altered when the amino acids are linked up to make proteins. Nevertheless, parts of molecules can be described as being hydrophobic or hydrophilic, depending on whether or not they would partition themselves into the organic phase. Hydrophobicity has two components: the “squeezing out” of water-insoluble components due to the attraction of water for itself and the self-association of nonpolar materials due to dipoledipole interactions, dipoleinduced-dipole, and induced-dipoleinduceddipole interactions (London dispersion forces) discussed in Chapter 1.4. Water repels nonpolar materials because the surface of the nonpolar material cannot form hydrogen bonds with the water, and its shape therefore reduces the number of hydrogen bonds that the water can form at the surface—the water molecules at the surface are in a higher energetic state, having some of their hydrogen bonds broken. Thus it takes energy to insert a nonpolar material into the water phase. The amino acids can be classified solely on the basis of their hydrophobicity or hydrophilicity. In this case, we have three categories: the hydrophilic amino acids include aspartic acid, glutamic acid, lysine, and arginine; the hydrophobic amino acids include valine, leucine, isoleucine, methionine, phenylalanine, and tryptophan; the “neutral” amino acids, those that are neither strongly hydrophilic nor strongly hydrophobic, include glycine, alanine, serine, histidine, proline, threonine, cysteine, glutamine, and asparagine. PEPTIDE BONDS LINK AMINO ACIDS TOGETHER IN PROTEINS Proteins are formed as a linear, unbranched chain of amino acids. The amino acids are covalently linked by a peptide bond formed between the amino group of one amino acid and the carboxyl group of the next. The formation of the peptide bond is a dehydration reaction, as shown in Figure 2.3.8. Cells make proteins by the sequential addition of amino acids to the carboxyl terminus of the growing chain. This is accomplished on the ribosome when specific tRNAs (transfer RNAs) with their specific bound amino 131 132 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION NH2 NH2 C CH2 NH CH2 CH2 CH2 CH2 + NH2 HN CH2 H2N CH2 O C C H2N C OH H NH+ CH2 O C H2N C OH H OH H Arginine (Arg) R Lysine (Lys) K C O Histidine (His) H FIGURE 2.3.4 The chemical structures of the basic amino acids, lysine, arginine, and histidine. These contain ionizable NH groups so that at neutral pH, these residues would contribute positive charge to the protein. OH CH3 OH HC CH2 H2N C OH O H2N C C CH2 O C H2N C C OH OH H H Serine (Ser) S H Threonine (Thr) T OH Tyrosine (Tyr) Y O NH2 C O C NH2 CH2 H2N O C CH2 CH2 O C H2N C O C OH H Asparagine (Asn) N H OH Glutamine (Gln) Q FIGURE 2.3.5 The chemical structure of the polar amino acids. The polar amino acids contain groups that can form hydrogen bonds with water. These groups are soluble in water. As a result, those portions of proteins which contain large numbers of polar residues will usually be exposed to water. They help solubilize proteins in solution, so we might expect soluble proteins to be coated with polar, acid, or basic R groups. Prot ein St ructur e H3C H3C CH3 H O H2N C H2N C OH H Glycine (Gly) G C C H2N OH H C Alanine (Ala) A CH2 O H 2N C OH H H2C CH CH3 CH O CH3 CH3 C C H 2N OH H Leucine (Leu) L Valine (Val) V H3C O CH C O C OH H Isoleucine (Ile) I CH3 NH S CH2 SH CH2 H2C CH2 HN C CH2 O H2N C H Proline (Pro) P OH C CH2 O C H H 2N OH Cysteine (Cys) C C H CH2 O C H 2N OH Methionine (Met) M C H CH2 O C H2N OH Phenylalanine (Phe) F C O C H OH Tryptophan (Trp) W FIGURE 2.3.6 Chemical structures of the nonpolar amino acids. PROTEIN FUNCTION CENTERS ON THEIR ABILITY TO FORM REACTIVE SURFACES n-Octanol Hydrophobic molecules Proteins perform a variety of functions in cells, and these can be broadly classified as structural, catalytic, transport, or regulatory. All of these functions require the surface of the protein to interact with other surfaces. FIGURE 2.3.7 Assessment of hydrophobicity by the partition coefficient between octanol and water. A hydrophobic molecule is dissolved in n-octanol and then an aliquot of the solution is placed in contact with water and shaken. The material will distribute itself between the two phases. The ratio of the concentrations at equilibrium in the organic phase to the water phase defines the partition coefficient, which is an equilibrium constant that can be used to calculate the free energy of transfer from the organic phase to the water phase. Hydrophobic materials will have a higher concentration in the n-octanol phase; hydrophilic materials will have a higher concentration in the water phase. As discussed in Chapter 1.5, enzymes provide a surface on which biochemical reactions can occur. These reactions can occur in the water phase, but only slowly because of the large activation energy required. Binding of the substrates to the enzyme surface alters the path of the reaction, enabling it to proceed more quickly. This lower energy pathway is produced by the interaction of the protein surface with the substrates of the biochemical reaction. The details of the protein surface enable catalysis, the speeding up of a reaction without appearing in the overall stoichiometry of the reaction. The surface of the protein must closely match the surface of the substrates, which means that it cannot match other substrates well, because the protein surface cannot match two different surfaces. Thus the protein surface simultaneously enables catalysis while it confers specificity— the enzyme works only with specific substrates. acid bind to the appropriate triplet codon of the mRNA. The amino acid is attached to the tRNA via its carboxyl group. When the peptide bond is formed, the entire growing chain is transferred to the free amino terminal of the next amino acid. The ribosome then moves one frame (nucleotide triplet) and the process is repeated (see Chapter 2.2). In the same way that protein enzymes interact with biochemical substrates on the surfaces of the proteins, structural proteins interact with other components by virtue of their surfaces. The proteins that make up connective tissue are sticky. They bind to themselves and to a variety of other proteins. Their surfaces make them sticky. Closely matching surfaces allow proteins to interact with other proteins, thereby allowing one protein to Water 133 134 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION H3C CH3 CH H2N H3C CH2 O C C C H2N OH H O C H2N CH3 CH O C C CH2 NH O C C OH H OH H H + Valine Phenylalanine Peptide bond H2O FIGURE 2.3.8 Formation of a peptide bond between two amino acids. The overall reaction is shown. The actual reaction involves many intermediary steps, catalyzed by enzymes. COOH Ser Leu Ile Val Glu Gln Cys NH2 Gly Thr Tyr Cys S S Ile His S Leu Cys Lys Gln Pro Thr Leu Cys Ser S Thr Tyr Glu Phe Asn COOH Gly Gln Ser Asn Tyr Cys His Val Phe Asn Arg Glu S NH2 S Leu Phe Gly Gly Cys Leu Val Val Tyr Glu Ala Leu FIGURE 2.3.9 Primary structure of insulin. Insulin is a protein hormone secreted by the pancreas in response to high blood glucose. It causes peripheral tissues to take up the glucose, thereby lowering the plasma glucose concentration back toward normal. Insulin is synthesized as a larger, single polypeptide chain that is modified by excision of two peptides to form the A chain, with 21 amino acids, and the B chain with 30. This is one example of posttranslational modification that occurs with many proteins. regulate another through the binding together of matching surfaces. THERE ARE FOUR LEVELS OF DESCRIPTION FOR PROTEIN STRUCTURE Proteins take on their remarkably diverse functions because they can fold to form specific shapes and because some of the amino acids that make up the proteins have inherent chemical reactivity. These shapes provide a surface for the binding of materials, and this binding originates all of the functions of proteins. The shape of proteins has four levels of description: 1. 2. 3. 4. Primary Secondary Tertiary Quaternary. THE PRIMARY STRUCTURE OF A PROTEIN IS ITS AMINO ACID SEQUENCE The primary structure of a protein refers to its amino acid sequence. It defines the chemical connectivity of the constituent amino acids. In 1953, Sanger determined the amino acid sequence of the A and B chain of insulin. This was the first protein whose entire amino acid sequence was determined and the work was a major milestone in biochemistry. The chemical structure of insulin is shown in Figure 2.3.9. THE SECONDARY STRUCTURE OF PROTEIN REFERS TO THE FOLDING OF AMINO ACIDS IN ADJACENT SEQUENCES Proteins have three regular secondary structures: the α-helix, the β-sheet, and the β-turn. These are complicated, three-dimensional structures. The α-helix was first postulated by Linus Pauling, Robert Corey, and Herman Prot ein St ructur e Cα H Amino N of next amino acid Carboxyl C of one amino acid Distributed pi electrons in double bonds require trigonal geometry of both C–O bond and C–N bond.This makes O, C, N, H and both Cα nuclei coplanar Cα H Cα Cα H Cα Cα Electrons establish a rapid equilibrium between two alternate forms of bonding. This is called resonance. This is equivalent to partial double bonds for both the C–O and C–N bonds FIGURE 2.3.10 Planar peptide bond. The peptide bond involves a combination of an carboxyl group of one amino acid with the amino group of the next amino acid. The CN bond and C 5 O bond resonate: the electrons alternate between a single CN and double C 5 N bond, and simultaneously between a double C 5 O bond and a CO2 bond. This is equivalent to a distributed electron density over both bonds. This restricts free rotation about the CN bond, locking all six nuclei into a single plane. Branson (The Structure of Proteins: Two HydrogenBonded Helical Configurations of the Polypeptide Chain, Proc. Natl. Acad. Sci. USA 37:205211, 1951) on the basis of a planar peptide bond and linear hydrogen bond length of 0.272 nm, and no requirement for an integral number of amino acids per turn of the helix. The planar nature of the peptide bond was key. Pauling realized that resonance between the carbonyl oxygen and the CN bond would produce a partial double bond character to the CN bond that would prohibit free rotation about its axis. The H bonding would be trigonal and therefore planar. The carboxyl C would be similarly planar, and because both the N and the carboxyl C were connected, all six nuclei connected to the peptide bond would be coplanar. This idea is shown in Figure 2.3.10. Although the CN bond cannot freely rotate, due to its partial double bond nature, the NCα bond is not so restricted. This Cα nucleus is connected to the next C that participates in a peptide bond, and this CαC bond can also rotate. The result is that the next peptide bond, which also forms a plane, is rotated relative to the first. Two dihedral angles are defined to describe this rotation: ϕ is the dihedral angle about the NCα bond and ψ is the dihedral angle about the CαC bond. These are shown diagrammatically in Figure 2.3.11. Pauling and co-workers used the idea of the stiff plane of the peptide bond and the known length of the hydrogen bond to deduce the structure of the alpha helix, shown in Figure 2.3.12. In this structure, the polypeptide backbone traces a right-handed helix (going from the amino terminus to the carboxy terminus). Note that the C 5 O bonds point towards the carboxy terminus, where they hydrogen bond with the amino hydrogen. The protein forms a rod with the side groups of the amino acid sticking out more or less radially. Rotation around the N–Cα bond is called phi, φ Cα H N φ ψ C The orientation of successive planes of the peptide bonds are determined by φ and ψ Cα Rotation about the Cα–C bond is called ψ FIGURE 2.3.11 Rotation of successive peptide bonds. Each peptide bond defines a plane. Between each bond there is allowable rotation around the NCα bond (defined as ϕ) and rotation about the CαC bond (defined as ψ). The result is a change in direction of the polypeptide chain. The second major secondary structure in proteins is the beta sheet, shown schematically in Figure 2.3.13. These result from sideways hydrogen bonding of a linear chain of amino acids whose peptide plane is bent at the Cα carbon. There are two ways for this to be accomplished: parallel beta sheets join segments of the polypeptide chain whose amino to carboxyl direction is going in the same direction. Antiparallel beta sheets join polypeptide chains of opposite N to C polarity. Structures not fitting into these categories are often called “random coils,” although they may not be random at all. There are four main principles involved in the formation of these secondary structures. Hydrogen Bonding Stabilizes Structure The polar centers of the carbonyl oxygen and the amide nitrogen are hydrogen bonded to other structures in the 135 136 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION protein, either another amide bond or a polar side group of an amino acid. These hydrogen bonds stabilize protein structure and are instrumental in forming the α-helix. Carboxy terminus Hydrophilic Groups Face Water; Hydrophobic Groups Face Away from Water The protein will fold so that highly charged groups and polar groups are on the outside of the protein facing water. Here the interaction between these hydrophilic groups and the hydroxyl groups of water stabilizes the structure. Carbon Side Groups Cannot Occupy the Same Space Hydrogen This is a simple way of saying there is steric hindrance in the structure. Atoms in the amino acids exhibit repulsive forces when placed too close together. These forces ultimately arise from the interpenetration of atomic or molecular electron orbitals. Nitrogen Oxygen R groups Single bond Hydrogen bonds C–C and C–N bonds in the alpha helix backbone 0.54 nm = 3.7 amino acid residues Electrostatic Interactions Stabilize Structures Groups with opposite electric charge attract each other and this attractive force can stabilize the threedimensional arrangement of amino acids. Similarly, groups with the same sign electric charge repel each other and this repulsive force can also stabilize the structure by preventing closer movement of the electrically charged groups. TERTIARY STRUCTURE DEALS WITH THE THREE-DIMENSIONAL ARRANGEMENT OF ALL OF THE AMINO ACIDS Amino terminus FIGURE 2.3.12 The alpha helix. The dark bonds indicate those in the polypeptide backbone that are directly involved in the peptide bonds. The dashed lines indicate hydrogen bonds that bond successive turns of the helix with each other. These hydrogen bonds connect the carbonyl of one amino acid with the amino hydrogen of the fourth amino acid down the chain. The tertiary structure of proteins deals with how the regional structures are put together in space. For example, the α-helices may be oriented parallel to each other or at right angles. So the tertiary structure refers to the folding of the different segments of helices, sheets, turns, and the remainder of the protein into its native three-dimensional structure. Commonly, membrane proteins are anchored in the membrane by hydrophobic Carboxy terminus Carbon Amino terminus N C Hydrogen Nitrogen Parallel Oxygen Carboxy terminus R groups Amino terminus N C Antiparallel Carboxy terminus Single bond Hydrogen bonds C–C and C–N bonds in the beta sheet backbone Amino terminus C N FIGURE 2.3.13 Structure of the beta sheet. Strings of amino acids are bonded laterally through hydrogen bonds. The arrangement can be between strings that have the same amino to carboxy orientation, called parallel, or between strings with opposite orientation, called antiparallel. The planar peptide bonds line up to form a kind of pleated sheet. Prot ein St ructur e alpha helices that may be far removed in the primary sequence but closely apposed in the tertiary structure of the protein. the interconversion of disulfide bonds until the right ones are formed. CHAPERONES AND CHAPERONINS HELP PROTEINS FOLD QUATERNARY STRUCTURE REFERS TO THE INTERACTION OF A PROTEIN WITH OTHER PROTEINS As they are made on the ribosome, proteins begin to fold up. Sometimes the primary structure of the protein alone can determine the proper final shape, and denaturing the protein (adding materials or heat that disrupts its shape) is reversible. In other cases, proteins need help in determining their shape, and they are synthesized on a kind of scaffold that ensures that they fold properly. These scaffolds are generally other proteins called chaperones. There are two kinds of chaperones. Molecular chaperones bind to unfolded or partially unfolded proteins and stabilize their structure, preventing them from being degraded. Chaperonins directly facilitate the folding of proteins. Molecular chaperones are members of the Hsp70 family of proteins; hsp means “heat shock protein” because these increase after heat stress to an animal that would denature proteins. Complexes of eight Hsp60 molecules form a barrel-shaped chaperonin that aids protein folding. The quaternary structure refers to the interaction of a protein with other proteins or other components of the cell. This association of the protein with other elements of the cell can alter its three-dimensional shape and its activities because the close proximity of another protein’s surface can alter the shape of the protein. Many proteins exist in the cell in complex macromolecular assemblies in which quaternary structure enables or regulates the function of the component proteins. POSTTRANSLATIONAL MODIFICATION REGULATES AND REFINES PROTEIN STRUCTURE AND FUNCTION Proteins that come off the ribosome are not yet finished. The cell processes the newly made proteins in several ways, including: G G G G PROTEOLYTIC CLEAVAGE Cells make many proteins in precursor form with longer primary sequences than the finished product. Proteolytic cleavage forms the final protein by chopping off unwanted parts of the protein. formation of disulfide bonds folding into the functional form cleaving the proteins at specific sites chemical modification. PROTEINS ARE CHEMICALLY MODIFIED IN A VARIETY OF WAYS PROTEIN DISULFIDE ISOMERASE CATALYZES DISULFIDE EXCHANGE Cells chemically modify proteins after translation in a variety of ways, some of which are shown in Figure 2.3.14. These include: Proteins often are stabilized by disulfide bonds between cysteine residues in the protein. These cysteines are not necessarily close to each other on the primary sequence but must be close in the tertiary structure of the protein. The protein disulfide isomerase catalyzes G acetylation methylation G NH2 CH3 CH2 S CH CH2 CH2 H3C N O H3C C CH2 O NH C H N-acetylation C CH2 O NH C NH+ H2 C H2C C H N CH O C H 5-Hydroxy lysine CH2 O H N C H2C C H 3-Methyl histidine 3-Hydroxy proline N CH CH2 O C H OH C C CH OH C OH O O OH OH C CH2 HN C O C H OH Gamma carboxy 4-Hydroxy proline glutamic acid FIGURE 2.3.14 Examples of some chemical posttranslational modifications of proteins. The N-terminal amino acid of many proteins (about 80% of them) is acetylated. Lysine groups can also be acetylated (not shown). Acetylation may regulate the life span of proteins as nonacetylated proteins rapidly degrade. Hydroxylation of lysine in the 5 position occurs in collagen. In collagen, proline is converted to 3 hydroxyproline or 4 hydroxyproline. Histidine, particularly in actin, is methylated. Arginine can also be methylated (not shown) and this modification is part of the histone code (see Chapter 2.2). Glutamic acid residues in prothrombin (a protein involved in clotting of blood) and in bone proteins are carboxylated at the γ-side chain carbon to form gamma carboxy glutamic acid. The close proximity of the two carboxyl groups in gamma carboxy glutamic acid creates a binding site for calcium ions, and this confers Ca21 sensitivity to the coagulation process. The gamma carboxylation reaction requires vitamin K, which is necessary for normal blood coagulation. 137 138 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Rest of the protein GPI-linked protein Mannose O – O P O Mannose H2 C H2 C O HC NH2 O Mannose O O P O Mannose N-acetyl galactosamine O– N-acetyl glucosamine Inositol O – O P O H H C C H O H C Extracellular face H O O C C O CH2 CH2 H2C HC O 2 CH2 H2C CH2 H2C CH2 H2C CH2 H2C CH2 H2C CH2 Membrane bilayer lipid core H2C CH2 H2C CH2 H2C CH2 H2C CH2 H2C CH3 CH3 H3C H3C FIGURE 2.3.15 Anchors of membrane proteins. Some membrane proteins are anchored in the membrane by attachment of hydrophobic parts such as myristic acid or palmitic acid. Myristic acid is a 14-carbon hydrocarbon chain with a carboxyl group on one end that can covalently attach to the N-terminal glycine of membrane proteins. Palmitic acid is a 16-carbon saturated monocarboxylic fatty acid and attaches to proteins similar to myristic acid. The process of attaching myristic acid or palmitic acid is called myristoylation or palmitoylation. Farnesyl is a polymer of three 5-carbon units called isoprene. Farnesyl attaches covalently via a thioester bond to cysteine residues somewhere in the middle of the protein and helps anchor some proteins in the membranes. Other proteins link to GPI or glycosylphosphatidylinositol, a glycosylated membrane lipid that helps keep proteins in the membrane. G G G G G CH2 CH3 C H2C HC CH2 CH2 H2C H2C C CH2 CH3 H2C HC Farnesyl H 2C C CH2 CH3 H 2C HC CH2 CH2 H2C H2C C CH2 Cysteine residue of membrane protein O C N H O hydroxylation gamma carboxylation glycosylation myristoylation or palmitoylation phosphorylation. Proteins in the endoplasmic reticulum are often glycosylated, meaning that sugars or sugar derivatives are covalently attached to the proteins. N-linked glycosylation CH N-terminal glycine of membrane protein NH S CH2 HC Myristic acid CH2 CH2 H2C 2HC C C NH NH Intracellular face O O occurs when carbohydrate branches are added to the side chain NH2 of asparagine through N-acetylglucosamine. O-glycosylation occurs on the side chain OH of serine, threonine, or hydroxylysine and the connecting carbohydrate is N-acetylgalactosamine. Proteins stick in membranes because they are anchored there by a variety of posttranslational modifications, as shown in Figure 2.3.15. Prot ein St ructur e PROTEIN ACTIVITY IS REGULATED BY THE NUMBER OF MOLECULES OR BY REVERSIBLE ACTIVATION/ INACTIVATION The cell can alter the activity of its component proteins by altering the number of copies of the protein, or by activating or inactivating the proteins that are already present. Regulating the number of protein molecules requires synthesis of new protein molecules or degradation of existing ones. This takes time and degrading existing proteins wastes energy. Reversible activation or inactivation of proteins can achieve rapid regulation that also conserves energy. Cells use phosphorylation/dephosphorylation of proteins to regulate their activities. Attachment of a phosphate group changes the charge on a local region of the protein, which alters its threedimensional shape and changes its activity. Serine, threonine, and tyrosine residues are the targets for these phosphorylation reactions (see Figure 2.3.16). An enzyme that phosphorylates proteins is called a protein kinase; one that dephosphorylates proteins is called a protein phosphatase. Cells contain a variety of protein kinases and protein phosphatases (see Figure 2.3.17). OH –O P O Protein state A O OH –O P OH O CH2 C P H3C O O Protein kinase Protein phosphatase O HN –O ATP Pi CH O C H N C CH2 O C NH C O C ADP Protein state B PO4 H H H Serine (Ser) S Threonine (Thr) T Tyrosine (Tyr) Y FIGURE 2.3.16 Phosphorylation sites of proteins. Serine, threonine, and tyrosine all have hydroxyl groups that can be esterified with phosphate. This alters the shape and charge of the protein surface in that area, leading to changes in protein activity. FIGURE 2.3.17 Phosphorylation cycle for protein regulation. Some proteins have phosphorylation sites that can be covalently linked to phosphate from ATP through the action of a variety of protein kinases. This alters the local charge of the protein, which in turn changes its shape and its activity. The protein returns to its dephosphorylated state by the action of protein phosphatases. Clinical Applications: Protein Folding Diseases With the advent of the microscope in the mid-1800s, Pasteur, Koch, and many others formulated the germ theory of infectious diseases. We now know that microscopic viral, bacterial, protozoan, or parasitic agents cause a long list of diseases: smallpox, polio, rabies, HIV, yellow fever, anthrax, bubonic plague, syphilis, tuberculosis, cholera, gonorrhea, malaria, sleeping sickness, schistosomiasis, to name but a few. Each of these infectious agents contains DNA or RNA that codes for the organism’s own proteins and enables replication of its nucleic acid. Viruses do this by using the host’s machinery. Bacteria, protozoans, and parasites are self-contained organisms that use the host’s environment for their own reproduction. Investigation of a group of diseases called transmissible spongiform encephalopathies (TSE) required revolutionarily new thinking about infection. These diseases include CreutzfeldtJakob disease, kuru, GerstmannStraussler syndrome (GSS), and fatal familial insomnia (FFI) in humans, and scrapies and bovine spongiform encephalopathy (BSE) in animals. Kuru is a neurological disease among the Fore, a linguistic group of people in Papua New Guinea. In ritual cannabalism, these people ate the bodies of dead relatives. After long incubations, they developed partial paralysis and loss of motor control and eventually died. Because the disease seemed to run in families, a genetic cause was first proposed, but later rejected. Carleton Gajdusek, a pediatrician and virologist, was unable to transmit the disease to any animal, including primates. Igor Klatzo, a neuropathologist, examined tissues sent by Gajdusek and found that kuru was a unique disease without precedent, the closest condition being CreutzfeldtJakob disease. In a bizarre twist, William Hadlow, a veterinarian neuropathologist, saw some of Klatzo’s photomicrographs in an exhibit at the Wellcome Medical Museum in London, and noted a startling resemblance between neurohistological changes in kuru and those in scrapie, a neurological disease in sheep first described in 1732. It was known to be infectious, but the infectious agent was not yet identified. Hadlow wrote to Gajdusek and published a letter in The Lancet (Continued) 139 140 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Clinical Applications: Protein Folding Diseases (Continued) in 1959. The similarity of kuru to infectious scrapie prompted Gajdusek to further attempt inoculating animals with kuru. In 1966, he and his co-workers transmitted kuru to chimpanzees and then CreutzfeldtJakob disease to chimpanzees in 1968. Gajdusek earned the 1976 Nobel Prize in Medicine. Although it was initially assumed that the infectious agent must be some virus, investigators failed to identify any virus or any immunological response to one. The agent was not destroyed by UV irradiation or nucleases that typically inactivate nucleic acids. Tikvah Alper and her co-workers in 1966 found that the infectious agent of scrapie was too small to provide a nucleic acid code. In 1967, J.S. Griffith proposed that a protein alone could be the infectious agent of TSE if it was a pathogenic form of a host protein that could convert normal host protein to the pathogenic form. Finally, in 1982, Stanley Prusiner identified the infectious agent as a protein devoid of nucleic acids. He coined the term “prion” which stands for proteinaceous infectious particle. Prusiner earned the 1997 Nobel Prize in Medicine. SUMMARY The core of amino acids is their asymmetric alpha carbon. Attached to one side is an amino group. On the other side is a carboxyl group. The third group is hydrogen and the fourth group is a variable group, usually referred to as a side chain or R group. These variable groups define the class of the amino acids. Glutamic acid and aspartic acid have a carboxyl group on the side chain. At neutral pH, this group ionizes and therefore has a negative charge. Lysine, arginine, and histidine are basic amino acids because their side chains have a basic chemical character. At neutral pH, these are positively charged. Serine, threonine, and tyrosine have hydroxyl groups that confer polar character. Asparagine and glutamine have an amide group that is also polar. Nonpolar amino acids include glycine, alanine, valine, leucine, isoleucine, proline, cysteine, methionine, tryptophan, and phenylalanine. Some of these are highly hydrophobic, such as tryptophan, isoleucine, leucine, and phenylalanine. Proteins are made by the formation of peptide bonds between the amino group of one amino acid and the carboxyl group of another. Because of this bond, the basic character of the amino group is neutralized, and the acidic character of the carboxyl group is neutralized, and the character of the chain of amino acids is determined by the sequence of the side chains. We describe protein structure on four levels: the primary sequence describes the linear sequence of amino acids along the peptide chain backbone, proceeding from the amino terminus to the carboxy terminus. The protein folds into secondary local structures such as α-helices, β-sheets, and β-turns. The arrangement of these secondary structures in three-dimensional space produces the tertiary structure. Combination of proteins with other structures produces macromolecular complexes with quaternary structure. Some proteins spontaneously fold Prions are infectious proteins. They “reproduce” by causing normal cellular prion protein (PrPC) to fold up differently, converting it into the pathogenic, or scrapie, isoform (PrPSc). The native PrPC appears to have three α-helices and two short β-strands; PrPSc has two α-helices and much more β-sheet. This transition from α-helix to β-sheet is the fundamental event underlying prion diseases. Proteolysis of PrPSc produces a smaller, protease-resistant molecule of about 142 amino acids (Prp 2730) which polymerizes into amyloid that presents itself in the disease state. Ritualistic cannabalism transmitted kuru among the Fore people of New Guinea; industrial cannabalism spread BSE (“mad cow disease”) in Europe. There is more than one bad way to fold a protein. Increasing numbers of patients have contracted a new variant of CreutzfeldtJakob disease (vCJD) from prion-tainted beef. Because of its long incubation time, we do not yet know the price of mad cow disease. Current thinking that the disease can be eradicated completely by control of infection is wrong. The disease can spontaneously appear without infection. into their “native” shape, whereas others are assembled on a kind of scaffold that helps them fold up properly into their active form. Denaturation of proteins occurs when the protein loses its normal shape. In some cases this is reversible, but usually loss of the proper folding causes irreversible loss of function. Hydrogen bonding, electrostatic interactions, hydrophobic interactions, and steric hindrance all help stabilize proteins in their secondary structures. Disulfide bonds between cysteine side chains help stabilize higher order structure. Proteins undergo posttranslational modifications after they are synthesized. These include N-glycosylation or O-glycosylation, proteolytic cleavage, hydroxylation, methylation, acetylation, γ-carboxylation, covalent attachment of hydrophobic molecules that anchor proteins in membranes, and phosphorylation of specific hydroxyl groups on side chains of serine, threonine, and tyrosine. Protein function depends on the way their surfaces interact with the surfaces of other materials—substrates, structural elements, or other materials to which the proteins bind. Catalytic activity or structural or regulatory roles of proteins depend on the close match of their surface with the surface of the things they bind to. This also determines the specificity of the protein’s action. In general, activity of proteins in the cell can be regulated by altering the amount of protein or by altering its intrinsic activity. Reversible regulation can be achieved by phosphorylation/dephosphorylation of proteins. REVIEW QUESTIONS 1. Name the two acidic amino acids. What makes them acidic? Name three basic amino acids. What makes them basic? 2. Name five polar amino acids. What makes them polar? Name the nonpolar amino acids. Prot ein St ructur e 3. What does “hydrophobic” mean? What does “hydrophilic” mean? What is the partition coefficient? How does it measure hydrophobicity? 4. What is a peptide bond? Where do you find it? 5. What is the primary structure of a protein? 6. What is an α-helix? β-sheet? What interactions among protein side chains stabilize these structures? 7. What is “posttranslational modification”? Name four different kinds of posttranslational modification. 8. What residues are most often acetylated? What residues are hydroxylated? What is “γ-carboxylation”? What amino acid is γ-carboxylated? What is glycosylation? What is myristoylation or palmitoylation? 9. How can proteins be anchored in hydrophobic membranes? 10. Describe the phosphorylation/dephosphorylation cycle for regulating activity of proteins in the cell. 141 2.4 Biological Membranes Learning Objectives G G G G G G G G G G G G G Describe fatty acids and what is meant by saturated and unsaturated fatty acids Distinguish between cis and trans arrangement around a double bond in an unsaturated fatty acid Describe the constituents of phosphatidic acid, phosphatidylcholine, phosphatidylethanolamine, phosphatidylserine, and phosphatidylinositol Identify hydrophilic and hydrophobic groups in membrane lipids Recognize the steroid ring structure of cholesterol Recognize the chemical structures of cardiolipin, sphingosine, sphingomyelin, and ceramide Define surface tension Describe how amphipathic lipids reduce surface tension Describe motion in the plane of a lipid bilayer Describe the fluid mosaic model of biological membranes Distinguish between integral and peripheral proteins Describe caveolae and clathrin-coated pits Describe how secreted proteins are synthesized on the ER membrane BIOLOGICAL MEMBRANES SURROUND MOST INTRACELLULAR ORGANELLES As discussed in Chapter 2.1, cells contain a variety of subcellular organelles and the hallmark of most of them is that they are surrounded by a membrane. These membranes divide the cell into several compartments in which enzymes and substrates are sequestered away from the rest of the cell. This separation into compartments is required for the functioning of these organelles. Maintenance of this compartmentalization requires selective transport of materials across the membranes. Table 2.4.1 lists the various subcellular organelles with their approximate contributions to the cell volume and membrane area. The proportions of cell volume and area represented by the subcellular organelles vary markedly with cell type and activity. BIOLOGICAL MEMBRANES CONSIST OF A LIPID BILAYER CORE WITH EMBEDDED PROTEINS AND CARBOHYDRATE COATS The composition of biological membranes varies enormously among the different subcellular organelles, but all biological membranes share a basic structure. The core of the membrane is a lipid bilayer. Embedded in this core are a variety of proteins that carry out many of the activities of the membrane, including selective membrane transport, and some of both the lipids and proteins have carbohydrate coats. This basic structure is shown in Figure 2.4.1. The rest of this chapter expands on this general description. ORGANIC SOLVENTS CAN EXTRACT LIPIDS FROM MEMBRANES Gorter and Grendel, in 1925, provided early evidence for the lipid bilayer structure of membranes when they extracted the lipids from erythrocytes with acetone and spread them over the surface of water. They noted that the area occupied by the lipids was about twice the calculated area of the surface of the erythrocytes. Because the only membrane in the erythrocytes is the plasma membrane, they concluded that the membrane was a lipid bilayer. This confirmed earlier observations by Ernst Overton in the 1890s that there is an excellent correlation between the ability of a number of solutes to enter cells and their solubility in olive oil. Overton concluded that the surface of the cell was made up of lipids similar to olive oil. BIOLOGICAL MEMBRANES CONTAIN MOSTLY PHOSPHOLIPIDS Organelles can be isolated by cellular disruption and differential centrifugation, as described in Appendix 2.1.A1. The lipids in the membranes can be extracted into an organic phase, typically a chloroform/methanol mixture, because these lipids are hydrophobic (see Figure 2.3.7), and then the lipids can be separated into their component classes by chromatography, and the amounts can be measured. Approximate lipid composition of some membranes is given in Table 2.4.2. 142 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00013-6 Biological Membra nes However, it should be kept in mind that the lipid composition varies not only with the kind of organelle but also with the kind of cell. TABLE 2.4.1 The Relative Area and Enclosed Volumes of the Major Subcellular Membranes in a Typical Liver Cell Membrane Percent of Total Cell Volume Enclosed Plasma membrane Approximate Percent of Number per Total Cell Cell Membrane 100 1 2 22 1700 39 Rough ER 9 1 35 Smooth ER 2 1 16 Golgi apparatus 4 1 7 Nuclear inner membrane 6 1 0.2 Lysosomes 1 300 0.4 Peroxisomes 1 400 0.4 Mitochondria PHOSPHOLIPIDS CONTAIN FATTY ACYL CHAINS, GLYCEROL, PHOSPHATE, AND A HYDROPHILIC GROUP Table 2.4.2 shows that phospholipids comprise the most abundant class of lipids in membranes. Each consists of four parts: a glycerol backbone, a phosphate esterified to one end of the glycerol molecule, a polar head group attached to the phosphate, and two fatty acids esterified to the other two hydroxyl groups of the glycerol. The glycerol phospholipids form the major subclass of the phospholipids, and they contain phosphate esterified to the C-3 hydroxyl group of glycerol and two fatty acids esterified to the C-2 and C-1 hydroxyl groups. Fatty acids consist of two distinct regions (see Figure 2.4.2): a long hydrocarbon chain and a carboxylic acid group (COOH). The most frequent length of the carbon chain is 16 or 18 carbon atoms. As discussed in Chapter 1.4, carbon can form four single bonds with bond angles that approximate those between the center of a tetrahedron and its vertices. In a hydrocarbon chain, two of these bonds link a carbon with adjacent carbons and two remain for other bonds. When hydrogen is covalently bound to all of the two remaining bonding orbitals, the hydrocarbon is saturated. Unsaturated fatty acids contain one or more double bonds between carbon atoms. Those containing more than one double bond are polyunsaturated. The variety of phospholipids is produced from the variety of fatty acids and from the different polar head groups. The most common lipid constituent of membranes is phosphatidylcholine, with other phospholipids making major contributions. The components of the simplest phospholipid, phosphatidic acid, are shown in Figure 2.4.2. Values were estimated by quantitative electron microscopy. Carbohydrate coats Proteins Lipid bilayer core FIGURE 2.4.1 Basic structure of biological membranes. Membranes consist of a lipid bilayer core to which proteins are attached in a variety of ways. In addition, some lipids and proteins have carbohydrate groups attached to them. A variety of polar groups can be attached to the phosphate of phospholipids. These groups include serine, inositol, and ethanolamine, which can be methylated to form choline. The resulting phospholipids are shown in Figure 2.4.3. TABLE 2.4.2 Approximate Lipid Composition of Different Cell Membranes Lipids Percentage of Total Lipids by Weight Plasma Membrane Phosphatidylcholine Mitochondrial Membrane Endoplasmic Reticulum Membrane 24 39 40 Phosphatidylethanolamine 7 35 17 Phosphatidylserine 4 2 5 Cholesterol 17 3 6 Sphingomyelin 19 0 5 7 0 0 22 21 27 Glycolipids Other 143 144 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Glycerol H H 1 C OH Palmitic acid (C16:0) OH H 2C Phosphate OH -O C 3 OH H Stearic acid (C18:0) OH CH2 CH2 HC CH2 CH2 H2C H2C CH2 CH3 H2C CH2 CH2 H2C H2C Nonpolar tail HC H2C CH2 CH2 CH2 CH2 CH2 H2C CH2 H2C H2C CH2 HC CH2 H2C H3C CH3 C O CH2 H2C H2C CH2 H2C H2C H CH2 CH2 CH2 CH2 3 CH2 H2C H2C H2C C H2C H2C CH2 Polar head H CH2 HC CH2 O O H2C H2C H2C H2C 2C CH2 CH2 CH2 C O C H2C H2C H2C 1 CH2 CH2 CH2 H O H2C H2C H H CH2 CH2 CH2 H2C OH O C H2C O-O P O- Oleic acid (C18:1, n9) CH2 CH2 H2C Phosphatidic acid OH O C O C O P OH H H2C CH3 H2C CH2 H2C CH2 H3C FIGURE 2.4.2 Components of a simple phospholipid, phosphatidic acid. Glycerol forms the backbone. Each of the three carbons in glycerol covalently bonds to a hydroxyl group (OH). Rotation about the CC single bonds can position the C1 OH or C3 OH in any orientation. The central carbon, C2, is an asymmetric carbon when C1 and C3 are differently substituted, and will show stereoisomerism when different groups attach to the two ends. To distinguish the 1 and 3 positions, the numbering nomenclature shown in the figure is used. The structure of inorganic phosphate is also shown. Fatty acids form the hydrophobic core of phospholipids. The chemical structure of three common fatty acids: palmitic acid, stearic acid, and oleic acid is shown. The single carboncarbon bonds allow free rotation about the axis connecting the two carbon nuclei. Each carbon atom forms bonds that nearly line up the centers of two tetrahedrons with their respective apices. Palmitic acid is a 16-carbon saturated fatty acid, meaning that every carbon’s bonds other than the carboxyl carbon are fully occupied with single bonds to carbon or hydrogen. In the nomenclature of fatty acids, it is designated 16:0, indicating a 16-carbon fatty acid with no double CC bonds. Stearic acid is a 18-carbon saturated fatty acid, designated as 18:0. Oleic acid is a mono-unsaturated fatty acid, meaning that one pair of carbon atoms are joined by a double bond. In this case, the double bond is between carbons 9 and 10, numbering down from the carboxyl carbon. Its designation is 18:1 Δ9, where the 1 indicates one double bond and the Δ9 indicates the double bond begins at Carbon 9. There is an alternate number system starting from the terminal methyl group of the fatty acid, instead of starting at the carboxyl group. This nomenclature uses the prefix omega (ω); thus oleic acid is 18:1 ω 2 9. Here the 18 stands for the length of the hydrocarbon chain, 1 indicates the number of double bonds, and ω 2 9 indicates the position of the double bond numbering from the methyl end. Substitution of the omega with the letter n is becoming popularized, with the same meaning. As discussed in Chapter 1.4, double bonds produce kinks in the hydrocarbon chain due to restricted rotation about C 5 C double bonds. Oleic acid has a cis orientation of the H atoms around the double bond, meaning the two hydrogens are on the same side of the double bond. The structure of a phosphatidic acid is shown on the right. This particular phosphatidic acid has an oleic acid molecule esterified to C-2 of glycerol and a palmitic acid molecule esterified to C-1. The fatty acids differ from molecule to molecule, but this is a typical arrangement in which saturated fatty acids occupy the C-1 position and unsaturated or polyunsaturated fatty acids occupy the C-2 position. Phosphatidic acid illustrates a common property of this class of lipids in that it consists of spatially separated water-soluble polar or hydrophilic groups and water-insoluble nonpolar or hydrophobic groups. Phosphatidic acid Phosphatidyl ethanolamine Phosphatidyl choline Polar group H 1 C C 3 CH2 C O CH2 H2C H2C CH2 H2C H2C H2C CH2 H2C CH2 H2C CH2 H2C CH3 CH2 CH2 H2C CH2 H2C CH2 CH2 H2C CH2 H3C CH2 H 2C CH2 H 3C Fatty acid: palmitic acid C16:0 H 2C CH3 H 2C CH2 CH2 H 2C CH2 H 3C H 2C CH3 CH2 H 2C CH2 H 2C CH2 H 2C CH2 CH2 H 2C CH2 H 2C CH2 HC CH2 HC Nonpolar tail HC H 2C H 2C CH2 CH2 C O CH2 CH2 H 2C 3 H H 2C HC CH2 H 2C CH2 CH2 H 2C H2C CH3 CH2 H 2C CH2 H 2C H C O H 2C H 2C CH2 HC C 2 CH2 CH2 CH2 CH2 H 2C H2C CH2 H 2C CH2 H 2C H 2C Polar head O CH2 H 2C H 2C HC CH2 H 2C CH2 CH2 CH2 C O C H 2C H 2C H O C O CH2 CH2 1 H H 2C H H 3 CH2 CH2 H 2C H C O H 2C H 2C CH2 HC O CH2 H 2C H 2C 2C CH2 CH2 CH2 HC CH2 HC H2C CH2 H 2C 1C O C H 2C H 2C H O C O CH2 CH2 CH2 CH2 HC CH2 H 2C H2C H H 2C H H 3 CH2 CH2 CH2 CH2 H C O H 2C O P O- O O CH2 CH2 CH2 H2C 2C O C H 2C H2C 1C O C O H2C CH2 CH2 H CH2 H2C CH2 3 O O C H H C 2C O H H H CH CH OH HC OH CH C CH O H OH O P O- CH2 OH Phosphate O P O- O O H H 1C O O C H2C H H 2C O H2C O H H Glycerol O P O- O O C HC H 2C CH2 OH OH NH2 + H 2C CH2 Phosphatidyl inositol Polar group CH3 H3C N CH3 NH2 O-O P O- Phosphatidyl serine CH3 H3C H 2C CH2 H 2C CH2 H 3C Fatty acid: oleic acid C18:1 n9 FIGURE 2.4.3 Chemical structures of some common glycerophospholipids. The phosphate group in phosphatidic acid is esterified to the hydroxyl group of several other hydrophilic molecules including ethanolamine, choline, serine, and inositol. These form the lipids shown. Each of these are named for the hydrophilic group and the fatty acids, as in 1-palmitoyl, 2-oleolyl phosphatidylcholine. 146 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION PLASMANYL PHOSPHOLIPIDS AND PLASMENYL PHOSPHOLIPIDS USE FATTY ALCOHOLS INSTEAD OF FATTY ACIDS A major subclass of the phospholipids use fatty alcohols instead of fatty acids, forming an ether linkage with glycerol instead of an ester. These are called plasmanyl glycerol phospholipids. Most of these are modified to contain a vinyl ether linkage, in which the alcohol group is doubly bonded to the rest of the fatty alcohol chain. These are termed plasmenyl glyercol phospholipids or plasmalogens. Their structures are shown in Figure 2.4.4. They make up 1520% of the total phospholipids of cell membranes. SPHINGOLIPIDS USE SPHINGOSINE AS A BACKBONE AND ARE PARTICULARLY RICH IN BRAIN AND NERVE TISSUES Sphingolipids are present in many membranes but they are particularly rich in brain and nerve tissues. There are three classes of sphingolipids: sphingomyelin, cerebrosides, and gangliosides. Sphingomyelin is the only one of these that is a sphingophosphatide. It is analogous to the glycerophosphatides except that it contains sphingosine instead of glycerol as the core structure that links the hydrophilic phosphate and choline to the hydrophobic hydrocarbon chains. Sphingosine is a derivative of the amino acid, serine. The chemical structures of sphingosine and sphingomyelin are shown in Figure 2.4.5. The fatty acid amide of sphingosine alone is called a ceramide (see Figure 2.4.5). Ceramides can be linked through the hydroxyl group to sugar groups to form Diacyl glyercol phospholipid H H 1 C O O C R1 H 2C CARDIOLIPIN IS TWO GLYCEROLIPIDS LINKED BACK TO BACK The structure of cardiolipin is shown in Figure 2.4.6. It consists of two phosphatidic acid molecules linked through another glycerol. Mitochondrial membranes are particularly rich in cardiolipin. CHOLESTEROL CONDENSES MEMBRANES Cholesterol, shown in Figure 2.4.7, is the most abundant steroid in animal tissues. All steroid hormones are derived from cholesterol. It possesses a rigid ring structure that attracts normally flexible phospholipid chains to itself, causing membranes to become more rigid in the vicinity of this molecule. PHOSPHOLIPIDS IN WATER SELF-ORGANIZE INTO LAYERED STRUCTURES All of the glycerophospholipids and sphingolipids we have discussed are characterized by the spatial separation of a polar head group, consisting of ionized groups, hydroxyls and carbonyl oxygens, and a long tail consisting mainly of hydrocarbons. The polar head group is capable of interacting with water and each of the materials there individually is water soluble. This (Plasmalogen) Plasmenyl glycerol phospholipid X O P O– –O O H C H H 1 3 O OTHER LIPID COMPONENTS OF MEMBRANES INCLUDE CARDIOLIPIN, SPHINGOLIPIDS, AND CHOLESTEROL Plasmanyl glycerol phospholipid X –O another class of sphingolipids, the cerebrosides. The sugar part contains a number of hydroxyl groups and is hydrophilic. In some cases the sugar part is quite large and branched, forming a ganglioside. Both cerebrosides and gangliosides form another class of lipids called glycolipids because they incorporate sugar derivatives. H C O H 2C X O P O– –O O H C O H H C 2C H 1 3 H O O P O– O H C 3 O C O CH2 C O CH C O R2 CH2 R2 CH R2 R1 Fatty acids X = polar head group: ethanolamine, choline or inositol H R1 Vinyl ether bond Fatty alcohol FIGURE 2.4.4 Plasmanyl glycerol phospholipids and plasmenyl glycerol phospholipids. Some long-chain hydrocarbons have an alcohol and not a carboxyl group at their end. These can be joined to glycerol through an ether linkage, forming a plasmanyl glycerol phospholipid which is typically enriched in the sn-2 position with polyunsaturated fatty acids. In most cases, the carbon adjacent to the ether is joined in a double bond, forming a vinyl ether bond that characterizes the plasmenyl glycerol phospholipids, also known as the plasmalogens. Biological Membra nes Sphingomyelin CH3 H3C N CH3 Choline + H 2C CH2 Sphingosine O P O- Ceramide O Serine H HO C H HO H C H C C H HC H HO NH2 HC H2C CH2 H2C Different from glycerol CH3 CH3 C O CH2 H 2C H 2C H 2C CH2 H 2C H 2C CH2 HC H 2C CH2 CH2 H 2C CH3 CH2 H 2C CH2 H 2C H 2C Nonpolar tail HC CH2 HC Amide linkage CH2 CH2 CH2 H 2C H2C H CH2 CH2 CH2 CH2 H C CH2 CH2 H 2C H2C O HN H 2C HC CH2 CH2 Amide linkage H 2C H 2C C H 2C CH2 CH2 C CH2 CH2 CH2 H HC H 2C H 2C H2C HC H 2C H 2C CH2 H C O CH2 CH2 CH2 H2C C HN H 2C H O H CH2 CH2 H C HC HC HO Polar head H 2C CH2 H2C H2C CH2 CH2 H3C H3C FIGURE 2.4.5 Chemical structures of sphingosine, ceramide, and sphingomyelin. Note that sphingosine does not have glycerol as a core structure and joins a hydrocarbon through an amide linkage instead of an ester to form sphingomyelin. region of the molecule is hydrophilic, meaning water loving. The hydrocarbon tail is not soluble in water; it is hydrophobic or water hating. Hydrophobic parts are also described as lipophilic or fat loving. Molecules having both of these separate domains are said to be amphipathic, from the Latin and Greek amphi meaning having two sides. These two sides are illustrated by the space-filling model in Figure 2.4.8. These two separate domains of phospholipids are crucial to their behavior in cells. When placed in water, the polar heads of these molecules remain associated with water and form hydrogen bonds with it; the long hydrocarbon, nonpolar tails repel the water and associate with each other, forming a self-organized structure, the lipid bilayer. Much of this behavior of the lipids resides in the nature of water. To see this, we need to learn more about the behavior of water at hydrophobic interfaces. SURFACE TENSION OF WATER RESULTS FROM ASYMMETRIC FORCES AT THE INTERFACE At the airwater interface, water molecules are subjected to asymmetric forces, as shown in Figure 2.4.9. These molecules have lost some of their bonds connecting them to the bulk phase and are, therefore, in a higher energy state than water in the bulk phase. These molecules are partially evaporated. Thus it takes energy to promote water molecules to the interface and the energy of the surface increases with its area. At constant temperature and pressure, the change in surface energy is the change in the Gibbs free energy, G. The change is given by: ½2:4:1 dG 5 γ dA where dA is the increment in area, dG is the increment in Gibbs free energy, and γ is the surface tension. Since the energy has units of force 3 distance, the tension has units of force per unit length. Typical units of γ are dynes cm21. In SI units, γ is expressed in N m21 or J m22. The surface tension is a measure of how much more a water molecule at the surface is attracted to the bulk water phase because of the increase in intermolecular forces compared to the surface. WATER “SQUEEZES OUT” AMPHIPATHIC MOLECULES Recall that amphipathic molecules consist of spatially separated water-loving or hydrophilic head group and a 147 148 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Cardiolipin Phosphatidylcholine Water CH H2C H H C 2C H 1 CH2 O –O P O O P O– O O O H H C 3 H2C CH2 H2C CH2 CH HC H2C CH CH2 H2C CH2 H2C H2C CH2 CH2 H2C Nonpolar tail H2C H2C H2C CH2 CH2 H3C H2C H2C CH2 FIGURE 2.4.8 Space-filling model of phosphatidylcholine. Nonpolar surfaces are shown in gray or white. Polar surfaces are dark blue (oxygen), light blue (phosphorus), or intermediate blue (nitrogen). Charged surfaces attract water by dipoledipole interactions. The hydrophilic or water-loving parts of the molecule are all concentrated at one end. The hydrophobic (water-hating) or lipophilic (fat-loving) groups are located at the opposite end. CH2 CH2 CH2 CH3 CH2 H2C H2C CH2 H2C CH2 CH2 CH2 H2C H2C H2C H2C HC Fatty acids CH2 CH2 CH2 Hydrophilic groups – Hydrophobic groups (= Lipophilic groups) H2C H2C CH2 + Glycerol O C O H2C CH2 CH2 C1 CH2 CH2 H2C C2 H2C H2C H2C H O C H2C CH2 H2C H O C O CH2 CH2 Phosphate Polar head H O O C Choline H C 3 H O OH Air CH3 H3C FIGURE 2.4.6 Chemical structure of cardiolipin. Steroid nucleus Cholesterol 18 12 11 13 C 19 14 1 2 9 10 A B 5 3 4 Water 17 D 16 15 H3C CH3 8 CH3 7 Molecules at the surface have fewer attractive forces than molecules in the bulk phase; they are partially evaporated and have a higher energy 6 HO FIGURE 2.4.7 Chemical structure of the steroid nucleus and cholesterol. The steroids are all derived from the steroid nucleus, a perhydrocyclopentanophenanthrene nucleus of four rings: three fused rings in the phenanthrene arrangement and a five-carbon ring attached. Therefore, the nucleus has the name “cyclopentano” to refer to the fivecarbon ring; “phenanthrene” refers to the three six-membered rings and “perhydro” indicates that the double bonds of phenanthrene are saturated with hydrogen. The steroid nucleus is numbered as shown. FIGURE 2.4.9 Asymmetric forces of water molecules at the airwater interface. In the bulk phase of liquid water, the intermolecular forces acting on any water molecule are on average equal in all directions. In the air, intermolecular interactions are markedly reduced. In order to evaporate, energy must be added to the water molecules to break the attractive intermolecular forces. At the airwater interface, water molecules are subject to more intermolecular forces than those water molecules in the air phase, and fewer forces than water molecules in the bulk liquid phase. Therefore, water molecules at the surface have more energy than those in the bulk phase. In a sense, they are partially evaporated, having lost some but not all of their intermolecular bonds. If we were to increase the surface area, we would have to put in energy proportional to the area of increase. Biological Membra nes Hydrophobic fatty acyl tails Water has no attractive forces to hydrocarbon chains; being near them is the same as being partially evaporated—they have a higher energy; it takes energy to insert lipid into water Hydrophilic phosphate Air Hydrophilic choline Water Because it takes energy for the lipid to be in the middle of the water phase compared to at the surface, most lipids are found at the hydrophobic/hydrophilic surface where hydrophilic groups interact with hydrophilic water and hydrophobic groups interact with other hydrophobic groups. In essence, the water squeezes out the lipid to the air/water interface Water binds to hydrophilic head groups of phospholipids—they have the same number of attractive forces and have the same energy as water in the bulk phase hydrophobic or water-fearing hydrocarbon tail. Dissolving an amphipathic molecule in the bulk water phase disrupts the self-association of water, creating an interface within the bulk phase that requires energy. Because of this, the amphipathic molecule is “squeezed out” of the bulk water phase to the surface of the solution adjacent to the air. In this situation, the water molecules at the surface interact with hydrophilic groups in the amphipathic molecule, which lowers the energy of the surface, and the hydrophobic groups in the amphipathic molecule can interact with other hydrophobic groups through London dispersion forces. These ideas are shown diagrammatically in Figures 2.4.10 and 2.4.11. The free energy change for the transfer of phospholipid from the bulk phase to the surface is given as ½2:4:2 ΔGbulk.surface 5 Gsurface 2 Gbulk Since Gsurface is less than Gbulk, the free energy change for the transfer is negative, and therefore the transfer occurs spontaneously. AMPHIPATHIC MOLECULES SPREAD OVER A WATER SURFACE, REDUCE SURFACE TENSION, AND PRODUCE AN APPARENT SURFACE PRESSURE When amphipathic molecules such as phosphatidylcholine or oleic acid are dissolved in a volatile organic solvent (e.g., hexane, decane) and then layered over water, FIGURE 2.4.10 Water “squeezes out” phospholipids. The hydrocarbon parts of a phospholipid molecule within the bulk aqueous phase cannot form hydrogen bonds with the adjacent water molecules, so these water molecules cannot form as many hydrogen bonds as the other water molecules in the bulk phase. The set of water molecules surrounding the hydrocarbon are essentially partially evaporated and have a higher energy than the water molecules in the bulk phase. Therefore, to insert the hydrocarbon in the water takes energy. Phospholipids on the surface of the water, on the other hand, are in a lower energy state. The self-association of water thus “squeezes out” the phospholipid to the surface. the organic solvent evaporates and leaves a thin film of the lipid. As shown in Figures 2.4.10 and 2.4.11, these amphipathic molecules are squeezed out to the surface of the water, forming a layer a single molecule thick. The lipids form a monolayer. These lower the surface tension according to Eqn [2.4.1] because they lower the energy required to move water molecules to the surface (see Figure 2.4.11). This lowering of the surface tension can be measured using a Langmuir trough, as shown in Figure 2.4.12. This device has two barriers on the surface of the water. One barrier is fixed, the other is movable. Since the monolayer decreases the surface tension, the movable barrier feels a net force in the direction of the clean surfaces. By lowering the surface tension, the lipids appear to exert a surface pressure defined as ½2:4:3 π 5 γ0 2 γ where π is the surface pressure, γ 0 is the surface tension of the clean surface, and γ is the surface tension in the presence of the monolayer. Surface pressure has the same units as the surface tension. PHOSPHOLIPIDS FORM BILAYER MEMBRANES BETWEEN TWO AQUEOUS COMPARTMENTS Two monolayers can orient themselves back to back to form a bilayer between two aqueous compartments, as shown in Figure 2.4.13. This is the low energy state for this situation for the reasons we have already described. 149 150 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION If the phospholipids were dispersed throughout the solution, the surface area between the water phase and the hydrophobic, hydrocarbon phase would be large. Since it takes energy to produce this surface, the dispersed hydrocarbon is a high-energy state compared to the condensed one. A secondary cause of this spontaneous organization of the lipids is the attractive interactions among the hydrocarbon chains. LIPID BILAYERS CAN ALSO FORM LIPOSOMES The macroscopic lipid bilayer shown in Figure 2.4.14 is unstable to mechanical forces. When disrupted, the membrane breaks apart to form spherical bilayers separating an internal and external watery compartment. These hollow spheres are called liposomes (see Figure 2.4.15). These structures can also be generated directly from phospholipids by soaking them in water and adding sonic energy. Phospholipid at the air–water interface lowers the energy of the surface by providing attractive forces to the surface molecules; the hydrocarbon chains also attract each other by London dispersion forces Phosphatidyl choline Air Macroscopic planar bilayer membranes can be formed across a narrow aperture separating two solutions, as shown in Figure 2.4.14. These membranes are useful because they allow the study of single membrane channels incorporated into the membrane. ALTHOUGH LIPIDS FORM THE CORE, MEMBRANE PROTEINS CARRY OUT MANY OF THE FUNCTIONS OF MEMBRANES Phosphatidyl inositol So far we have been describing membrane lipids that form the core of biological membranes. These contribute to the barrier function of membranes, but many other functions of biological membranes are performed by protein constituents of the membranes. These functions include: Water Molecules at the surface have fewer attractive forces than molecules in the bulk phase; they are partially evaporated and have a higher energy G G FIGURE 2.4.11 Lowering of the surface energy by amphipathic molecules. Water at its ordinary surface is in a higher energy state—it takes energy to break the hydrogen bonds that binds the water molecules in the bulk aqueous phase. When phospholipids are present at the surface, water binds to the hydrophilic groups of the phospholipids, thereby reducing the energy needed to form additional surface. Because the energy of the surface is related to its area by the surface tension, reducing the energy of the surface per unit area is the same as reducing the surface tension. G G Transport: Cells must be able to move things into and out of the cell. Signal transduction: Cell must have mechanisms for responding to signals from other cells or from within the cell. These may be chemical, electrical, or mechanical signals. Recognition: Cells attach to other cells and to extracellular structures. They must be able to recognize where they should form attachments. Attachment: Cells must anchor themselves to the extracellular matrix or to each other. Often these attachments also provide a signaling pathway. Hydrophobic tails Fixed barrier FIGURE 2.4.12 The Langmuir trough. A clean water surface has a fixed barrier and a movable barrier. When only clean water forms the surface, the surface tension is γ0. Adding a lipid film reduces the surface tension to γ. The movable barrier experiences a net force toward the clean surface without lipid. Thus the lipid appears to exert a surface pressure. Air phase Movable barrier γ Water phase Polar head groups γ0 Biological Membra nes Phosphatidylcholine Cholesterol Phosphatidylethanolamine Aqueous layer Insulating hydrocarbon layer Aqueous layer Phosphatidic acid FIGURE 2.4.13 Structure of the lipid bilayer. Only some lipids are shown and in expanded format for clarity. In reality, the lipid bilayer forms a closed surface that approximates a plane. The interior of the lipid bilayer is fluid, consisting of hydrocarbon chains that are saturated (the straight chains in the figure) or unsaturated (bent chains in the figure). The phospholipids form the bulk of the bilayer. Lipids such as cholesterol have a rigid backbone that partially stiffens and solidifies the membrane. Cholesterol may accumulate in heterogeneous patches of membrane called lipid rafts. Barrier Water Water Narrow aperture FIGURE 2.4.14 Planar lipid bilayer between two aqueous compartments. Phospholipids were dissolved in hexane and “painted” over a small aperture drilled in a Lexan partition that separated two electrolyte solutions. After thinning, with the passive removal of hexane through the aqueous phase, a lipid bilayer forms between the two compartments. G Movement or force production: Cells often must move or transmit a force from inside the cell to the extracellular matrix. This requires connection of the cell’s cytoskeleton through the membrane to the extracellular matrix. FIGURE 2.4.15 A schematic drawing of a cross-section of a liposome. These are small structures about 50150 nm across. Because each bilayer is about 7.510 nm across, the thickness of the bilayer occupies a considerable portion of the entire liposome volume and the enclosed volume, the lumen, is small. Because of the high curvature and small size, the area on the outside of the liposome is nearly twice the area on the inside, and therefore significantly more lipid faces the outside compared to the inside surface of the liposome. Liposomes may find use someday to deliver drugs to specific locations within the body by incorporating recognition signals into the lipid bilayer. MEMBRANE PROTEINS BIND TO MEMBRANES WITH VARYING AFFINITY The proteins that perform the various functions listed above can be loosely classified according to how tightly they are bound to the membrane. Loosely bound proteins are called peripheral proteins, also sometimes called extrinsic proteins. They can be released from the membrane by relatively gentle procedures such as washing with a salt solution. Other proteins are called integral proteins, also sometimes called intrinsic proteins. These are tightly bound by the membrane and can be released only by resorting to drastic measures such as dissolving the membrane with a detergent. By coating the hydrophobic parts of the membrane proteins with hydrophilic material, detergents solubilize the membrane proteins. Examples of useful detergents in membrane research include the ionic sodium dodecyl sulfate (SDS) and the nonionic family of Triton detergents. Figure 2.4.16 illustrates integral and peripheral proteins. Proteins are held in membranes by the same kinds of interactions that hold lipids in the bilayer: hydrophobic and hydrophilic interactions. Many proteins have sequences of amino acids that penetrate all the way across the membrane. These are transmembrane proteins. In the parts of the protein exposed to lipid, hydrophobic amino acid side chains appose the lipid core: valine, leucine, isoleucine, phenylalanine, tryptophan, and methionine. Those parts of the protein exposed to water have a preponderance of hydrophilic amino acids: aspartic acid, glutamic acid, lysine, and arginine. The neutral amino acids can be present in 151 152 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Peripheral protein Glycolipids GPI anchors protein in the membrane Myristyl anchor Lipid bilayer core FIGURE 2.4.16 Schematic arrangement of membrane proteins in the lipid bilayer. Darker areas of proteins are predominantly hydrophilic; lighter areas denote hydrophobic areas. Hexagons signify hydrophilic carbohydrates. Proteins are usually anchored by hydrophobic sequences of amino acids; less often they are anchored by covalent attachment to hydrophobic materials such as GPI, farnesyl, palmitic acid, or myristic acid. Farnesyl anchor Integral, transmembrane protein either domain. Some proteins have multiple sequences that cross the membrane, with domains facing each of the watery solutions on the two sides of the bilayer. The sequences that cross the membrane are typically arranged as alpha helices, with the hydrophobic amino acids facing the hydrophobic lipid core. Other proteins may bind to the membrane by covalently attached hydrophobic groups. These include myristic acid (14:0 fatty acid), palmitic acid, (16:0) farnesyl, and glycosylphosphatidylinositol (GPI) (see Figure 2.4.16). Lateral diffusion Stretch Flip-flop Flexion Rotation LIPIDS MAINTAIN DYNAMIC MOTION WITHIN THE BILAYER Researchers have labeled phospholipids with probes that are sensitive indicators of molecular motion and have tracked the mobility of lipids and proteins within the plane of the bilayer. The results of these experiments show a variety of molecular motions, shown in Figure 2.4.17. This has given rise to the fluid mosaic model of biological membranes. The term fluid mosaic model describes a dynamic system in which the lipids form a plane that gradually curves around to form a closed surface—there are no exposed lipid edges. Lipid motion in the plane of the membrane includes rotation, flexion, stretch, and lateral diffusion in the plane. All of these movements are rapid. The proteins in the membrane can also move, unless they are anchored by binding to other components of the membrane. There is a gradient of fluidity of motion from the polar head group to the center of the hydrophobic interior of the membrane, being most fluid in the center and most anchored at the head group. Lipid movement from one side of the membrane to the other—the “flip-flop” FIGURE 2.4.17 Possible motions of lipids within a bilayer. Lipid molecules can move laterally within the plane of the membrane, rotate about their long axis, flex within the fluid interior of the membrane, or move from one side of the bilayer to the other. Most of these motions are fast, but the “flip-flop” reaction is very slow, occurring less than once in 2 weeks for any individual lipid molecule. Exchange of neighbors occurs very fast, on the order of 107 times per second. This rapid exchange gives rise to a rapid lateral movement within each half of the bilayer. reaction—is slow because it requires lipids to bring their hydrophilic head group through the lipid phase, which is energetically costly. Because of the very slow movement of lipids from one half of the bilayer to another, membranes can maintain an asymmetric composition, but it must actively make and sustain it. In most biological membranes, the two half bilayers differ in their composition. Cells add lipids only to the cytoplasmic side of membranes. An enzyme called a “scramblase” flips lipids from the cytoplasmic half to the extracellular half, but this enzyme is not very specific. The different composition of membranes arises from a second enzyme, a “flippase”, that flips only some phospholipids from the cytoplasmic half to the extracellular half. Biological Membra nes The fluidity of membrane lipids depends on their composition. Saturated fatty acids are known to be very stiff compared to unsaturated fatty acids. Saturated fatty acyl chains can be packed closely together, whereas the kinks produced by cis double bonds make it difficult to pack these chains close together. Because of this, unsaturated fatty acids promote fluid movement within the bilayer. Cholesterol is also a very rigid molecule. Cholesterol molecules orient themselves in the membrane with their steroid nucleus adjacent to the hydrophobic tails and the hydroxyl group adjacent to the polar head groups. The steroid nucleus is a flat, plate-like structure that partially immobilizes the nearby fatty acyl chain, thereby reducing the fluidity of the bilayer and increasing the mechanical stability of the bilayer. The motions of the lipids in the bilayer makes it appear as a two-dimensional fluid. It is fluid within the plane of the membrane, but relatively rigid perpendicular to this plane. The membrane proteins more or less “float” in this lipid see like so many icebergs in the North Atlantic. This combination of fluid lipids and iceberg proteins was the origin of the descriptive term, fluid mosaic model. LIPID RAFTS ARE SPECIAL AREAS OF LIPID AND PROTEIN COMPOSITION Lipid rafts are microdomains in biological membranes that contain different proportions of lipids and proteins from the rest of the membrane. They were discovered when portions of membranes were found to be less easily solubilized by detergents. Detergents are chemicals that dissolve membranes by providing their constituents with flotation devices: they bind the hydrophobic domains and coat them with water-soluble material. These detergent-resistant areas of membrane accumulate cholesterol and sphingolipids. Sphingolipids generally contain longer and straighter fatty acyl chains. These attract each other more forcefully than do unsaturated fatty acids, because the straight chains can pack more closely without the kinks in their chains. These aggregate into the raft microdomain. Because these chains are straighter, the membranes are also thicker at the rafts. The plasma membrane is thought to have many such rafts about 70 nm in diameter. CAVEOLAE AND CLATHRIN-COATED PITS ARE STABILIZED BY INTEGRAL PROTEINS The surface of cells forms a variety of specializations that curve inwardly, forming an indentation of the membrane. Caveolae are one of these. Caveolae are a subset of lipid rafts, but not all lipid rafts are caveolae. Caveolae are 6080 nm pits in the membrane that contain some 140150 oligomeric caveolin molecules. There are three mammalian caveolins: CAV1, CAV2, and CAV3, all have parts that bind to membranes. Their oligomeric structure is stabilized by a family of cytoplasmic proteins called cavins. Caveolar membranes are enriched in both cholesterol and phosphatidylserine (another phospholipid in which the head group is serine instead of ethanolamine, choline, or inositol). Depletion of the cholesterol or mechanical flattening of caveolae causes dissociation of cavin from caveolin (see Figure 2.4.18). Flattening of the caveolae occurs upon stretch of skeletal muscle, cardiac myocytes, endothelial cells, and fibroblasts. It may be that caveolae are involved in the sensing or response to mechanical stretch. Membranes can also form clathrin-coated pits that are involved in receptor-mediated endocytosis, in which parts of the membrane invaginate and pinch off, forming an interior vesicle with enclosed extracellular material. Clathrin consists of three heavy chain subunits (CHC17 or CHC22) and three light chains (CLC) that trimerize to form a triskelion, the unit of clathrin. These units then associate on membranes to form a clathrate (lattice) structure. The lattice structure consists of a number of pentagons and hexagons. The clathrin protein itself has a curvature to it, and this imparts a curvature to the clathrate and stabilizes the budding part of the membrane. The membrane is then pinched off by another protein complex called dynamin (see Figure 2.4.19). CAV1 Shear stress Cavin complex Plasma membrane FIGURE 2.4.18 Caveolae response to stretch. Caveolae are indentations or pits in the plasma membrane that are stabilized by a network of integral proteins that include oligomers of caveolin (CAV1). Cytoplasmic proteins called cavins stabilize the caveolin structure. When the membrane is stretched, the caveolae flatten and cavin dissociates from the caveolin. This may be part of how cells sense or respond to stretch. H+ ATP Clathrin heavy chains Light chains Clathrin coat Triskelion structure Dynamin FIGURE 2.4.19 Clathrin-coated pits and endocytosis. Clathrin consists of a trimer of heavy chains each of which binds a light chain. These selfassemble to form a clathrin coat that stabilizes budding membranes. The budding membranes are pinched off through the actions of dynamin. 153 154 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION mRNA 1. Protein synthesis begins on free ribosomes 40S 3. Complex of ribosome, peptide, and SRP binds to SRP receptor on the ER SRP Ribosome 60S 1 Signal sequence 4. Complex transfers to a translocon and elongation resumes NH2 2 3 ER membrane α 3' β 6 and 7. Chaperones bind to the elongating chain and help fold it properly SRP receptor 2. SRP recognizes the signal sequence, binds to it and the large ribosomal subunit, and stops elongation 4 Translocon 5 6 7 Folded protein Signal peptidase Cleaved signal sequence 5. Signal peptidases cleave off the signal sequence Chaperone protein FIGURE 2.4.20 Mechanism of synthesis of secretory proteins into the ER lumen in the pancreas. Synthesis begins on free ribosomes in the cytoplasm (1). The initial N-terminus of the protein contains a signal sequence that is recognized and bound to by cytosolic SRP, signal recognition particle, a nucleoprotein consisting of RNA and a set of six separate proteins (2). Binding of SRP stops elongation. The complex of nascent polypeptide, ribosome, and SRP is bound to the ER by the SRP receptor consisting of a peripheral α-subunit and an integral β-subunit (3). The SRP receptor transfers the ribosome to a translocon, a complex of proteins that spans the ER membrane and provides an aqueous channel for the protein across the membrane (4). The SRP dissociates from the complex in the process. The polypeptide chain grows and translocates across the ER membrane at the same time. The signal sequence is cleaved off (5) by signal peptidases in the ER lumen. The growing polypeptide chain is eventually completed and folds up into its active conformation (7). The folding is assisted by chaperone proteins and enzymes such as Hsp-70, protein disulfide isomerase, and peptidyl prolyl isomerase (6). Source: Modified from Lodish et al., Molecular Cell Biology, 4th Ed, W.H. Freeman, New York, NY, 2000. SECRETED PROTEINS HAVE SPECIAL MECHANISMS FOR GETTING INSIDE THE ENDOPLASMIC RETICULUM The synthesis of membrane proteins or proteins destined for secretion poses a problem for cells because the ribosomes on which the proteins are made are in the cytosolic compartment but their products are either in the membranes or in the extracellular compartment. This problem is solved by an elaborate mechanism, shown in Figure 2.4.20. SUMMARY Lipids readily dissolve in organic solvents such as chloroform while they are sparingly soluble in water. Several classes make up biological lipids including fatty acids, phospholipids, and steroids. They are typically made up of long chains of hydrocarbons or carbon rings. The major constituent of biological membranes are the phospholipids. These consist of a polar head group connected covalently to a nonpolar tail. The polar head group in turn consists of a hydrophilic group like inositol, serine, choline, or ethanolamine linked to a phosphate group, which in turn is esterified to glycerol. All of these are highly water soluble. The other hydroxyls of the glycerol are esterified to two fatty acids, which are highly water insoluble. Thus these phospholipids spatially separate hydrophilic and hydrophobic parts. When placed in water, the hydrophilic parts associate with the water while the hydrophobic parts associate with other hydrophobic molecules. When placed on top of water, these amphipathic molecules form a lipid monolayer. The surface tension of the water results from asymmetric forces on the surface from the bulk water phase and the air phase. Because the hydrophilic parts of the lipids attract water molecules on the surface, they reduce the asymmetry in forces. Accordingly, lipids reduce the surface tension. Experimentally, this appears as a surface pressure. Folding such a monolayer back on itself produces a bilayer membrane. This consists of a double layer of molecules in which the hydrophilic domain faces the water phase and the hydrophobic domain faces the interior of the membrane, occupied by other hydrophobic parts of lipid molecules. Other lipid aggregates include the liposome. The liposome is a bilayer that forms a hollow sphere. As in all molecules, chemical bonds in lipids can stretch, rotate, and flex. These motions along long chains produce motion within the hydrophobic interior of membranes. The hydrophobic chains are relatively well anchored at the polar head group, so there is a gradient of fluidity in the membrane, it being most fluid in the center and less at the periphery. Double bonds in Biological Membra nes the hydrocarbon chains make a kink in the chain that disallows close packing of the chains. Thus double bonds promote fluidity within the bilayer. Saturated fats, those that contain no double bonds, make membranes more rigid. Cholesterol is a rigid molecule composed of a plate-like steroid nucleus and a hydrophobic tail. It generally makes membranes more rigid. Lipids freely move in the plane of the membrane while motion across the membrane, the “flip-flop” reaction, is slow. Cells take advantage of this slow flip-flop to maintain asymmetric distributions of lipids in the two halves of the bilayers. Proteins embed in the membrane and move around, something like icebergs floating in a lipid sea. Thus the membrane is described as a “fluid mosaic.” However, thicker microdomains of the membrane contain concentrations of sphingolipids and cholesterol. These microdomains are called lipid rafts. Proteins can be loosely associated with membranes or tightly bound. The loosely bound proteins are called peripheral or extrinsic proteins and the tightly bound proteins are called integral or intrinsic proteins. Proteins are held in the membrane either by hydrophilichydrophobic interactions between their amino acids and the lipid and water phases or by attachment of hydrophobic groups such as myristic acid, palmitic acid, farnesyl, or GPI. Many proteins and lipids have carbohydrate coats. The synthesis of secreted proteins requires the synthesis of an endoplasmic reticulum (ER) signal sequence that is recognized by a signal recognition particle (SRP), which then binds to a receptor on the ER membrane. This enables transfer of the cytosolic ribosome to the ER membrane and subsequent simultaneous translation and translocation of protein across the ER membrane. Once inside the ER, the synthesized protein undergoes posttranslational modification. REVIEW QUESTIONS 1. Name the major membranes in the cell. Which membrane accounts for most of the membranes in the cell? 2. What is a saturated fatty acid? What is an unsaturated fatty acid? What effect does unsaturation have on the structure of the fatty acid? What do fatty acids attach to in phospholipids? 3. What is a phospholipid? What are the major types of phospholipids? Which chemical groups on the phospholipid are hydrophilic? Which groups are hydrophobic? What is the significance of spatial separation of hydrophilic and hydrophobic character in lipids? 4. What is the general structure of cardiolipin? What is sphingosine? What is sphingomyelin? What is a ceramide? How do these differ from phosphatidylcholine? 5. What is surface tension? What are its units? What do amphipathic lipids do to the surface tension? Why? 6. Name the various degrees of freedom of lipid motion in a bilayer. Which is the slowest? Which is the fastest? 7. What is meant by the term “fluid mosaic model”? 8. What is an integral protein? What is a peripheral protein? In what ways can proteins be anchored to membranes? 9. What is a liposome? What is a planar lipid bilayer? 10. What are lipid rafts? 11. What is meant by “caveolae”? What is a clathrin-coated pit? 12. How do secreted proteins get inside secretory vesicles? What is meant by “signal sequence”? 155 Problem Set 2.1 Surface Tension, Membrane Surface Tension, Membrane Structure, Microscopic Resolution, and Cell Fractionation 1. Consider a soap film stretched over a wire frame, one end of which is movable. Experimentally, one observes that there is a force exerted on the movable member as indicated in Figure 2.PS1.1. Clearly, this force depends on the dimensions of the wire frame. Therefore, we express the force per unit length as γ. Write an expression for the work performed in expanding the film a distance dx. Rewrite this in terms of the area increment, dA, by which the film is expanded. 2. Consider a soap film again in the form of a bubble as shown in Figure 2.PS1.2. The surface tension can be thought of as either the force per unit length or the energy per unit area. The minimal energy form for a soap film is the minimum area for a given volume. This is the sphere. So, in the absence of other effects, including gravity, the soap bubble should be a sphere. A. What is the total surface energy of the sphere? Remember that the variable we have been using for surface tension is γ. B. If the radius were to decrease by dr, what would be the change in the surface energy? C. Since shrinking decreases the surface energy, at equilibrium the tendency to shrink must be balanced by a pressure difference across the film, ΔP. At equilibrium, the work against this pressure for an increment in radius dr is exactly equal to the decrease in surface energy. That is, at equilibrium the free energy change is zero. Otherwise, the bubble would not be stable and it would shrink. What is the work that must be done against this pressure difference? Hint: Pressure is force per unit area, so the total force must be the area times the pressure. Work is force times distance. D. Equate the pressurevolume work in part C to the surface energy decrease in part B. From this equation, derive an expression for ΔP in terms of γ and r. This result is a famous equation, the Law of Laplace, which finds application in respiratory physiology and cardiovascular physiology. 3. When heart cells are exposed to a hypotonic medium, they swell and measurements show that their volume has increased. Measurements of their membrane capacitance, however, do not change. How can this happen? 4. Liposomes form structures 100 nm across their outside diameter. The average density of the lipids used to form the liposomes is 0.89 g cm23. Assume that the thickness of the bilayer is 8 nm. A. What is the volume of the lipid shell? What is its mass? B. What is the ratio of the outer surface area to the inner surface area of the liposomes? C. What is the enclosed volume of the liposome? Po Pi dr Movable barrier Surface tension l Direction of expansion r Soap film 156 FIGURE 2.PS1.1 Soap film on a wire frame. The soap film exerts a force per unit length on the movable barrier. This force is the surface tension. To expand the film, we must do work. FIGURE 2.PS1.2 A soap bubble of radius r. Because the surface tension results in an inwardly directed force, the bubble will tend to collapse unless there is a pressure difference across the membrane that prevents its collapse. © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00014-8 Problem Set D. Using the answers to A and C, what is the overall density of the liposomes? Assume that the liposome is filled with water with density 5 1 g cm23. E. How many liposomes can be derived from 100 mg of lipids? F. A drug is soluble to 5 mM. If the liposomes were formed in a solution of this drug, and therefore the enclosed volume included the drug to this concentration, how many moles of drug would be contained in the liposomes derived from 100 mg of lipids? 5. Isolated cardiac sarcoplasmic reticulum (SR) vesicles have an average outside diameter of about 150 nm. The membrane itself is about 10 nm thick. The enclosed volume can be estimated by measuring the efflux of passively loaded tracer materials such as mannitol, and the result gives 5 μL mg21 SR protein. A. How many vesicles are there per mg of SR protein? B. What is the surface area of the vesicles per mg of SR protein? C. If the SR Ca-ATPase Ca21 uptake activity is 4 μmol min21 mg21, what is the uptake activity per unit surface area? 6. The method of measuring the surface tension of a liquidair interface is the drop weight method. In this method, drops are allowed to form at the end of a tube of known radius, and a number of them are collected and then weighed so that the weight per drop can be determined accurately. The weight per drop is given by Tate’s Law (1864): ½2:PS1:1 W 5 2πrγ where r is the radius of the tube. This equation uses the idea that the surface tension is the force per unit length and that the maximum force that can be used to support the weight of the forming drop is the circumference of the tube times its surface tension. In practice, the weight of the drop is less than that given by Tate’s Law because some of the liquid supported by the tube remains after the drop falls. More detailed analysis makes use of a correction factor such that ½2:PS1:2 W 0 5 2πrγf where W0 is the actual weight per drop and f is the correction factor. It turns out that the correction factor f varies with rV 21/3, where V is the volume of the drop. Approximate values of the correction factor are given in Table 2.PS1.1: A. Using a tip with an outside diameter of 0.40 cm and an inside diameter of 0.20 cm, 20 drops of an organic liquid weighed 0.80 g. The density of the liquid was 0.95 g cm23. This liquid wet the tip. (Hint: This goes to determine whether you use the inside or the outside diameter!) Use the appropriate correction factor from Table 2.PS1.1, TABLE 2.PS1.1 Correction Factors for the Drop Weight Method rV21/3 f 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 0.7256 0.6828 0.6515 0.6250 0.6093 0.6000 0.5998 0.6098 0.6280 0.6535 0.640 0.603 0.567 0.535 Source: Data from A.W. Adamson, Physical Chemistry of Surfaces, Interscience, New York, NY, 1967. and the drop weight, to calculate the surface tension of the organic liquid. B. Using a tip with and outside diameter of 0.21 cm and an inside diameter of 0.18 cm, 20 drops of a liquid had a mass of 0.766 g. The density of the liquid was 1.00 g cm23. This liquid wet the tip. (Hint: This goes to determine whether you use the inside or the outside diameter!) Use the appropriate correction factor from Table 2.PS1.1, and the drop weight, to calculate the surface tension of the organic liquid. 7. The surface tension of pure water is 72.0 dyne cm21. 1 dyne cm21 is equivalent to 1 mN m21, which is the SI unit for surface tension. When dipalmitoyl lecithin is spread at the surface pressure is 50 Å2 mol21, 11 mN m21. What is the surface tension when dipalmitoyl lecithin is spread on the surface? 8. The tension in a biological membrane can be measured in a variety of ways. One way is called the pipette aspiration technique. In this technique, a specially manufactured micropipette is attached to a vesicle by light suction. These pipettes typically have a open diameter of 12 μm and have a square end. Application of increasing suction draws the vesicle into the pipette, forming a cylindrical part within the pipette and a spherical part outside of it (see Figure 2.PS1.3). A. Assume that the Law of Laplace holds for both the spherical part of the vesicle outside the pipette and the hemisphere within the pipette. Write the two equations relating Pv, Po to Dv and T, the tension in the vesicle, and Pv, Pp to Dp. There is only one tension in the membrane, which is the same everywhere. B. Defining ΔP 5 Po 2 Pp, use the answer in part A to solve for ΔP in terms of T, Dv, and Dp: 157 158 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Pv TABLE 2.PS1.2 Data for the Dilatation of Sarcolemma Vesicles Using the Micropipette Aspiration Technique Dv Dp Pp L Po FIGURE 2.PS1.3 Pipette aspiration technique. A vesicle obtained from a “bleb” on a cell when the cell is exposed to hypotonic medium is excised and attached to a micropipette by application of suction. Increasing the suction draws the vesicle into the pipette a distance L and reduces the diameter of the remaining vesicle. By LaPlace’s law, this increases the tension in the membrane. Continual decreases in Pp eventually causes the vesicle to rupture. The critical tension for rupture can be determined in this way. C. Solve part B to express T in terms of ΔP, Dp, and Dv D. A giant sarcolemmal vesicle was obtained from a rabbit muscle and subjected to the pipette aspiration technique (J.A. Nichol and O.F. Hutter, Tensile strength and dilatational elasticity of giant sarcolemmal vesicles shed from rabbit muscle, J Physiol. 493:187198 (1996)). The pipette diameter (Dp) was 19 μm and the vesicle diameter was 66 μm. At a pipette suction of 8 cm H2O, what is the tension in the membrane? E. What is the total area of the membrane in terms of Dv and Dp and the length L? F. When suction pressure is increased, L increases and Dv decreases. The increased tension stretches the membrane, causing a dilatation. The elastic area expansion modulus is defined as T ΔA K5 A where ΔA is the expansion of the membrane area due to dilatation and T is the tension. Considering your answer for part E, write an expression for the increase in area attributable to membrane dilatation, ΔA, in terms of an initial condition with Dvi and Li and a final condition with Dvf and Lf. G. In practice, the increase in the length of the projection is not long enough to cause an easily measurable difference in Dvf compared to Dvi. Assuming that the volume of the vesicle plus projection remains constant, express Dvf in terms of Dvi, ΔL, and Dp. 9. Using the micropipette aspiration technique described in Problem #8, the following data were obtained for the tension and area dilatation for sarcolemma vesicles obtained from rabbit skeletal muscle (see Table 2.PS1.2): The vesicle ruptured at the last point recorded. Tension (mN m21) ΔA/A Tension (mN m21) ΔA/A 1.4 2.6 3.9 5.1 6.6 7.9 0.0027 0.0058 0.0078 0.0112 0.0127 0.0160 9.1 10.5 11.9 13.1 14.5 0.0178 0.0212 0.0248 0.0261 0.0296 J.A. Nichol and O.F. Hutter, Tensile strength and dilatational elasticity of giant sarcolemmal vesicles shed from rabbit muscle, J Physiol. 493:187198 (1996) Calculate the elastic area expansion modulus (see Problem 2.PS1 problem #8 for a definition of the elastic area expansion modulus). 10. For light of wavelength 5000 Å (5500 nm), calculate the theoretical maximum resolution of an optical microscope. 11. Typically the energy of the electron beam in an electron microscope is known, because the voltage through which the electrons are accelerated is known. One electron volt is the energy gained by an electron when it is accelerated across a potential of 1 V. One electron has a charge of 1.602 3 10219 C. So the electron volt is 1.6310219 V C51.6310219 J (1 J51 V C51 N m). The rest mass of the electron is 9.109 3 10231 kg. A. Using this information, and assuming that all of the energy are converted to kinetic energy, calculate the momentum of a 150-keV electron (the denominator in Eqn (1.2.A1.2)). (Hint: Kinetic energy E 5 p2/2m.) B. Calculate the wavelength of an electron having a kinetic energy of 150 keV. C. Using the result of (B), calculate the theoretical resolving power of an electron microscope using a 150-keV electron beam. 12. In a Sorvall T-865 fixed angle rotor, the distance to the axis of rotation is 3.84 cm at the top of the tube and 9.10 cm at the bottom of the tube. Calculate the RCF at 20,000 rpm at the top and bottom of the tube. 13. We are centrifuging a collection of particles with diameter 150 nm and average density of 1.10 g cm23 through a water solution with density 1.0 g cm23 at 20,000 rpm. The viscosity of the water is 1 3 1023 Pa s, where Pa is the pascal 5 1 N m22. A. At a distance of 8.5 cm from the axis of rotation, what is the net force on a particle? B. From Stoke’s equation, calculate the frictional coefficient. C. What is the particle’s terminal velocity? D. What direction is the net force? E. What causes this net force? Problem Set 14. In eukaryotic cells (cells with a nucleus), ribosomes have two major subunits, a 60s and a 40s. If we assume both are spheres and have the same average density, what are their relative sizes? 15. Intestinal cells make a calcium-binding protein called calbindin. Calbindin has a molecular weight of about 9000 Da. Its synthesis requires the active form of vitamin D, 1,25(OH)2 cholecalciferol, to turn on the gene for the protein. When 1,25(OH)2 cholecalciferol is given to vitamin-D-deficient people, it takes about 45 minutes for the intestinal cells to make the first complete calbindin. A. Identify the major steps that could account for the 45-minute lag in appearance of calbindin. B. Eukaryotic cells (cells with a nucleus) attach amino acids to new proteins at the rate of about 2 per second. Is the synthesis of the protein the major part of the lag? (Hint: The average molecular weight of an amino acid is about 100 Da. You can estimate how many amino acids are in the protein from this information—you could look it up because its sequence is known, but we are just doing a “back of the envelope” calculation here.) 16. We have a double-stranded DNA segment of 1000 base pairs. Its nucleotide composition is randomly distributed among A, T, C, and G. Assume that each hydrogen bond in the double strand has an energy of 4 kcal mol21 (1 J 5 0.239 cal). A. How many hydrogen bonds are there in the segment? (Hint: Consult Figure 2.2.3 for the numbers of hydrogen bonds for each base pair.) B. What is the total energy necessary to pry apart the two strands, assuming that hydrogen bonding is the only force keeping them together? (It is not.) C. Assume that hydrogen bonds break when they are stretched 0.2 nm. How much force is necessary to break one? Pump 2 Pump 1 Flow rate R1 Reservoir S1 at C1 D. If all the hydrogen bonds in our DNA segment were to be ruptured all at once, how much force would be necessary? 17. The molecular weight of the protein myosin is 525,000 g mol21. Its sedimentation coefficient is 6.4S (this S is the Svedberg, not seconds) in water with a density of 1.0 g cm23, and its partial specific volume is 0.73 cm3 g21. Calculate the frictional coefficient, β. This is sometimes called f. A. From the molecular weight and the specific volume, calculate the radius myosin would have if it were spherical. B. From the radius calculated in (B), determine the frictional coefficient from Stoke’s equation. This is called f0. The viscosity of water at 25 C is 1 3 1023 Pa s, where Pa is the pascal 5 1 N m22. C. If the protein were spherical we would expect f 5 f0. Is myosin spherical? 18. Consider the device shown in Figure 2.PS1.4 that consists of two reservoirs, S1 and S2, that initially contain the limiting concentrations C1 and C2, respectively, where C1 , C2. A pump removes fluid from S1 at rate R1 and places it in reservoir S2. Therefore, as soon as the pumping starts the concentration in S2 begins to change. A magnetic stir bar rapidly mixes reservoir S2, and a second pump withdraws fluid from S2 at rate R2 and places it in a centrifuge tube. The total volume of S1 5 S2 and both are one-half of the capacity of the tube, so that when all of the solutions are pumped into the tube, the tube is filled. Show that for R2 5 2 3 R1, the gradient is linear in volume from C2 at the bottom of the tube to C1 at the top. 19. The SR is a specialized endoplasmic reticulum in skeletal, cardiac, and smooth muscle cells. It contains a Ca-ATPase pump that actively pumps Ca21 ions from the cytosol to an enclosed compartment within the SR, its lumen. The activity of the SR can be estimated by the rate of oxalate-supported Ca21 uptake. This activity is useful because it can also be measured in homogenates of the tissue. Evidence suggests that the Flow rate R2 Reservoir S2 at C2 Stirrer Centrifuge tube with gradient FIGURE 2.PS1.4 One way to make a gradient. Two reservoirs have the limiting concentrations C1 and C2. Pump 1 removes fluid from reservoir S1 at rate R1 and places it in reservoir S2, initially at C2 but then becomes diluted with fluid from S1. Pump 2 removes fluid from S2 at rate R2 and places it in a centrifuge tube. The gradient that is formed depends on the values of R1 and R2 and the volumes of the reservoirs. Both S1 and S2 begin with identical volumes equal to one-half of the volume delivered to the centrifuge tube. 159 160 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION oxalate-supported Ca21 uptake rate is only due to the SR and other organelles—the surface membrane or mitochondria—do not contribute to it. The left ventricles from a set of dogs were removed under general anesthesia, weighed and homogenized in 3 volumes of buffer (3 mL of buffer for every g wet weight of heart) and homogenate protein, volume and oxalatesupported Ca21 uptake rate were measured. The homogenate was subjected to differential and then sucrose-gradient centrifugation to isolate membrane vesicles of the SR. The following was obtained from an average of 10 preparations: heart homogenate volume: 4.1 mL g21 wet weight of heart, heart homogenate protein: 46.1 mg mL21, heart homogenate oxalatesupported Ca21 uptake rate: 119 nmol min21 (mg homogenate protein)21 isolated SR oxalatesupported Ca21 uptake rate: 3.44 μmol min21 (mg SR protein)21. A. Calculate the homogenate protein per g wet weight of heart tissue. B. Calculate the total homogenate Ca21 uptake rate per g wet weight of tissue. C. Assuming that the SR is 100% pure, how much SR is there, in mg of SR protein, per g of wet weight of heart tissue? (Hint: think about the units in the calculation.) Passive Transport and Facilitated Diffusion Learning Objectives G G G G G G G G G G G 2.5 transport. The cell must be able to control what goes across the membranes and how fast. There are three main mechanisms for transport: Describe the microporous membrane as a model of passive membrane transport Describe the lipid bilayer model of passive membrane transport Define the permeability of a membrane Describe how the permeability depends on the microscopic character of the membrane for a porous membrane Describe how the permeability depends on the microscopic character of the membrane and solute for a dissolution model of passive transport Be able to determine the free energy change for passive transport Distinguish between facilitated transport and diffusional transport on the basis of saturability, specificity, rates, and competition Write an equation showing the rate of facilitated transport as a function of its solute concentration with zero-trans concentration. Identify the variables and describe their meaning Distinguish between an ionophore and a channel Describe what is meant by channel gating Distinguish between voltage-gated channels and ligandgated channels MEMBRANES POSSESS A VARIETY OF TRANSPORT MECHANISMS As described in Chapter 2.4, membranes serve as effective barriers to the free movement of materials, thereby dividing the cell into compartments. This compartmentalization is necessary. In muscle, for example, it allows for the control of contraction by releasing Ca21 ions from a store (the specialized endoplasmic reticulum of the muscle cell) into the cytoplasmic compartment. Relaxation is then brought about by removing Ca21 ions from the cytosol back into the storage compartment. In another example, compartmentalization allows mitochondria to transduce the energy of oxidation of foodstuffs into the chemical energy of ATP. However, compartmentalization does not make sense if material absolutely cannot travel between the compartments. What is necessary is selective transport and regulated A. Passive transport 1. Diffusion 2. Facilitated transport B. Active transport 1. Primary active transport 2. Secondary active transport C. Osmosis. In this chapter, we will consider passive transport across two types of hypothetical membranes: a microporous membrane and a lipid bilayer membrane. The mechanisms of passive transport differ considerably between these two models, but the overall form of the equations is similar. In Chapter 2.6, we will consider active transport and then in Chapter 2.7, we will discuss osmosis. A MICROPOROUS MEMBRANE IS ONE MODEL OF A PASSIVE TRANSPORT MECHANISM Here we introduce the porous membrane as a model for biological membranes as shown in Figure 2.5.1. We consider here that a microporous membrane separates two solutions of different concentrations but the same pressure. The pores allow solute particles to pass, but the rest of the membrane that lacks pores is impermeable to the solute, and also to solvent water. We assume that the solute particles are small compared to the pores. First we write the flux, the flow per unit area within a single pore. This is governed by Fick’s Laws of Diffusion given as js 5 2D ½2:5:1 @CðxÞ @x @CðxÞ @2 CðxÞ 5D @t @x2 where js indicates the flux within the pore. We use the lower case “j” purposefully to distinguish it from Js, which we will use to signify the macroscopic flux across the entire membrane. The top equation is Fick’s First Law of Diffusion; the bottom equation is Fick’s Second Law of Diffusion. Here we are concerned only with flux across the membrane, in one direction, and the 161 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00015-X 162 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Membrane Pore CL Constant slope = (CL – CR)/ (0 – δ) δ js Pore CR n = N pores/unit area FIGURE 2.5.1 Schematic of the hypothetical microporous membrane. In this model, the membrane is a thin sheet of thickness δ. It is pierced by many cylindrical pores oriented perpendicular to the surface of the membrane. The radius of each pore is a and the number of pores, N, per unit area A is n 5 N/A. This model might not pertain to some cellular membranes, but it may describe some extracellular membranes such as the basement membrane, which supports many cells, especially epithelial cells, or it may represent the filtration membrane present in the kidney. one-dimensional forms of Fick’s laws apply to this situation. Let CL be the concentration on the left side of the membrane and CR be the concentration on the right. We can arrange it so that the volumes of the two baths are so large that CL and CR are effectively kept constant. Under these conditions, the solute flow will come to a steady state or stationary value. This means that neither the fluxes nor the concentration of solute changes with time. Fick’s Second Law of Diffusion becomes ½2:5:2 05D @2 CðxÞ @x2 Note that this situation cannot be literally true, as solute is moving from one compartment to another, so there must be some changes in C(x) with time. However, C(x) can be so nearly constant that we can ignore the very slight error. The solution to this equation is that @C(x)/@x is constant. This means that the concentration within the pores is linear with x. We solve this equation by two successive integrations, incorporating the boundary conditions that at x 5 0, C(x) 5 CL and at x 5 δ, C(x) 5 CR (see Figure 2.5.2). The concentration is written as CL 2 CR CðxÞ 5 CL 1 ½2:5:3 x 02δ and the concentration gradient is @CðxÞ CL 2 CR ΔC ½2:5:4 52 52 @x δ δ The flux in the pore is given from Eqns [2.5.1] and [2.5.4] as ½2:5:5 js 5 D ΔC δ The total flow of solute per pore, qs, is given by the area of the pore times the flux within the pore: ½2:5:6 qs 5 πa2 js δ x=δ x=0 FIGURE 2.5.2 Cross-section of a microporous membrane in the vicinity of a pore. Superimposed on the cross-section is a graph of the concentration gradient. The left compartment has a higher concentration (CL) than the right compartment with concentration CR. Under this situation, the flux through the pore is to the right. The total flow across an area A of the membrane containing N pores is ½2:5:7 Qs 5 Nqs 5 Nπa2 js The macroscopically observed flux across the membrane is the total flow of solute (Qs) divided by the macroscopic area of the membrane. Js 5 ½2:5:8 Qs A 5 Nπa2 js 5 nπa2 js A 5 nπa2 D ΔC δ According to this equation, the observed macroscopic solute flux across the membrane is linearly related to the concentration difference by a coefficient that includes the thickness of the membrane (δ), the density of pores in the membrane (n), the radius of the pores (a), and the diffusion coefficient of the solute (D). Often many of these parameters are not known with accuracy and we lump all of these terms together to write ½2:5:9 Js 5 pΔC where p is the permeability of the membrane to the solute. This phenomenological coefficient has the units of cm s21 and includes all of the microscopic parameters of the membrane: n πa2 D δ In this model, the permeability increases when the size of the pores increases, when the number of pores per ½2:5:10 p5 Passive Tr ansport and F acilit ated Di ffusion unit area of membrane increases, when the thickness of the membrane decreases, and when the diffusion coefficient of the transported solute increases. Which of these can be regulated? Typically membranes do not regulate their thickness, nor can the diffusion coefficient be altered. Channels can act like pores and they can be gated. That is, the channels can be opened or shut. This has the effect of controlling the area through which materials can be transported and this is a common way of regulating ion transport. Another way of physiologically regulating passive transport is by controlling the number of pores (or channels) in a membrane. The distinction between pores and channels lies in the substrate. We think of pores as holes in a substrate that will not collapse, as if we drilled a tiny hole in a thin plastic sheet. Lipid bilayers, however, will not support a watery void in their interior. Channels are proteins embedded in the membrane that line a watery pathway across the membrane and prevent its collapse by providing mechanical support on the sides of the pathway. The watery path across the membrane is made by the proteins that form the channel. DISSOLUTION IN THE LIPID BILAYER IS ANOTHER MODEL FOR PASSIVE TRANSPORT Consider now a markedly different model of the membrane. In this case, there are no pores, but we envision that a molecule may penetrate from the left to the right of the membrane by dissolving in the lipid bilayer core of the membrane, diffusing across the lipid, and then being extracted back into the aqueous phase on the other side of the membrane. This model of passive transport is shown schematically in Figure 2.5.3. The dissolution of the solute in the lipid membrane is described quantitatively by a constant called the partition coefficient (see Chapter 2.3): ½2:5:11 ks 5 equilibrium C in the lipid phase equilibrium C in the water phase Partition into membrane Partition into solution Diffusion 1 2 3 CR CL Lipid bilayer FIGURE 2.5.3 Cartoon of the lipid bilayer model of passive transport. The left and right compartments are separated by a lipid bilayer membrane. Solute, shown here as blue spheres, moves across the membrane in three well-defined steps. In step 1, the particle partitions itself into the lipid phase of the membrane. In step 2, the material diffuses across the lipid bilayer. In step 3, the material partitions itself back into the aqueous phase on the right side of the membrane. Overall transport rates are determined by the rates of steps 1, 2, and 3. diffusion across the lipid bilayer. This equation can be rewritten as ½2:5:13 Js 5 ks Ds;lipid ΔC δ This last equation is identical in form to that derived earlier in the microporous membrane model: ½2:5:9 Js 5 pΔC In the case of the dissolutiondiffusionsolution model, we identify the permeability as ½2:5:14 p5 ks Ds;lipid δ If equilibrium is reached quickly at both the left and right surface of the membrane, then diffusion through the lipid phase would limit the rate of transport. Let the concentration on the left side of the membrane be CL and the concentration on the right side of the membrane be CR. The concentration immediately inside the membrane on the left, by Eqn [2.5.11], will be ksCL, and the concentration on the right inside the membrane will be ksCR. The steady-state flux through the lipid phase is ks CL 2 ks CR Js 5 Ds;lipid ½2:5:12 δ Equation [2.5.14] neatly sums up many experimental observations concerning the permeability of materials through biological membranes. These are briefly summarized in Overton’s rules: where Ds,lipid is the diffusion coefficient of the solute in the lipid phase and δ is the thickness of the lipid phase. We write Js here because the entire area of the membrane is available for dissolution of the solute and A. The permeability is proportional to the lipid solubility. B. The permeability is inversely proportional to molecular size. Once again, the permeability is a single phenomenological parameter that relates the flux to the concentration difference. It incorporates all of the microscopic parameters of the membranesolute pair into a single parameter. In this case, these microscopic parameters are the partition coefficient, the thickness of the membrane, and the diffusion coefficient of the solute in the membrane. 163 164 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Thus we expect ethanol to permeate cell membranes quite easily because it is small and it is lipid soluble. On the other hand, glucose is larger and it is not readily lipid soluble and so it requires another mechanism to enter the cell. This mechanism is the carrier. Overton’s rules derive from two main components of Eqn [2.5.14]: the dependence of p on ks means that lipid solubility is a direct determinant of the permeability and the dependence on Ds means that size is an inverse determinant of permeability. Lipid solubility here depends on the hydrophobicity or lipophilicity of the solute. Different chemical groups in any molecule confer hydrophilic or hydrophobic character to those parts of the molecule, as described in Chapters 2.3 and 2.4. In particular, electric charge make a solute highly hydrophilic and not lipophilic. Thus charged or ionized solutes are generally highly impermeable by this dissolution mechanism of transport. Solutes enriched in hydroxyl groups, carboxyl groups, and amino groups are generally not easily permeable through this mechanism unless they are also very small. FACILITATED DIFFUSION USES A MEMBRANE-BOUND CARRIER For some valuable materials, the membrane permeability is not large enough for the cell’s needs and facilitated diffusion is used to carry solute across the membrane. One possible way a carrier could operate is shown in Figure 2.5.4. The transformations that occur to allow this transport are not known in detail. It is likely that the SL EL 1 E.SL 4 2 E.SR 3 ER SR FIGURE 2.5.4 Schematic diagram of carrier-mediated passive transport. The carrier is designated as “E.” In this scheme, an integral protein molecule in the membrane binds to a solute molecule on one side of the membrane. The carrier molecule then undergoes a transformation that has the effect of changing the side of the membrane that is accessible to the binding site. The solute molecule then dissociates from the carrier on the opposite side. The carrier then returns to its original shape. carrier provides something like a pore for the solute, but the pore is specifically designed to fit the solute. In this case, the solute molecule never dissolves in the lipid bilayer but is protected from it by a pocket of the protein carrier. The mechanism involved in facilitated diffusion shown in Figure 2.5.4 involves a sequence of four reactions: SL 1 EL -E 2 SL E 2 SL -E 2 SR ½2:5:15 E 2 SR -ER 1 SR ER -EL 222222222 Sum : SL -SR The sum of these four reactions is the movement of solute from the left side of the membrane to the right side. The carrier concentration does not enter into the overall stoichiometry of the reaction, because its presence on both sides of the reactions cancels itself out. It acts as a catalyst for transport because it determines its rate without being altered by the process. FACILITATED DIFFUSION SATURATES WITH INCREASING SOLUTE CONCENTRATIONS Facilitated diffusion can be distinguished from a purely diffusional mechanism because facilitated diffusion is saturable and it is specific. Plots of the flux versus the concentration are not linear, as you would expect from the diffusion mechanisms as shown in Eqns [2.5.8] and [2.5.13]. The rate increases with concentration but only up to a point. This is due to the fact that there are only so many carrier molecules in the membrane. When they are all busy, there can be no further increase in the rate of transport. This is analogous to the ferrying of people across a river that is too deep for most of them to wade. Although some can wade, most must cross only by ferry. When there are not too many people, the ferries can accommodate them easily and the transport rate will increase with each increase in the number of people waiting on shore. When the crowd on the shore gets too great, however, the ferries become full and the rate of transport can be increased further only by the number of brave souls who can wade the river (diffuse across the membrane) or by increasing the number of ferries. Figure 2.5.5 shows the kinetics of a saturable transport mechanism. These curves often closely resemble the hyperbolic plots characteristic of MichaelisMenten enzyme kinetics and can be fit to ½2:5:16 Qtrans 5 Qmax C Km 1 C where Qtrans is the flow of transported material across the membrane in moles per unit time, Qmax is the maximum flow, C is the concentration of the transported Passive Tr ansport and F acilit ated Di ffusion competitive inhibitors. This competitive inhibition is closely related to that observed in enzyme kinetics. In other cases, a compound sharing some similarity with the natural substrate may bind to the carrier but not be transported. If the binding is at the transport site, such a compound might inhibit transport. 12 Qmax 10 Qtransport 8 PASSIVE TRANSPORT OCCURS SPONTANEOUSLY WITHOUT INPUT OF ENERGY 6 Km 4 The chemical reaction for the overall transport is written as ½2:5:17 2 SoluteL -SoluteR and the free energy change for the reaction is 0 0 2 4 6 [S] 8 10 12 ½2:5:18 ΔG 5 GR 2 GL Substituting in with the chemical potential, we obtain FIGURE 2.5.5 Graph of a saturable transport mechanism. The rate of transport in moles per unit time per unit area is plotted against the concentration of material on the feed side [S]. A maximum transport rate, Qmax, can be identified. The substrate concentration at halfmaximal transport is used to characterize the affinity of the transport mechanism for its substrate. ΔG 5 nðΔμ0 1 RT ln CR Þ 2 nðΔμ0 1 RT ln CL Þ CR ½2:5:19 5 n RT ln CL solute on the feed side of the membrane, keeping the concentration on the opposite side at zero, and Km is a constant characteristic of the carrier. The term “Km” comes from MichaelisMenten kinetics, and these carriers almost certainly do not have the mechanism first proposed by Michaelis and Menten. Nevertheless, the term “Km” has come to mean “the concentration of substrate at half-maximal activity.” Sometimes this is referred to as Kt, the concentration at half-maximal transport. Eqn [2.5.16] is a simplified version of the exact solution of the kinetics of the scheme shown in Figure 2.5.4. where n is the number of moles of solute moving from left to right. Here there is no electrical work term because the charge on the molecule, z, is zero. If CR . CL, ΔG calculated from Eqn [2.5.19] will be positive. This means that the opposite process will occur. That is, solute will move from the right to the left, opposite to the direction shown in Eqn [2.5.17]. If ΔG 5 0, then no net movement occurs and CR 5 CL. If CR , CL, then ΔG calculated according to Eqn [2.5.19] will be negative and the reaction will proceed as written, with solute moving from the left to the right side of the membrane. Thus thermodynamics tells us what process can occur and with what change in free energy, but it does not give us an expression for the permeability or the rate at which the process will occur. FACILITATED DIFFUSION SHOWS SPECIFICITY Another distinguishing feature of carrier-mediated facilitated diffusion is its structural specificity. The parts of the carriers that bind transported solute are specifically designed for that solute and not others. For example, most cell membranes in the human contain carriers for glucose. They will transport D-glucose but not its enantiomer (mirror image compound) L-glucose. The carrier for glucose will not transport amino acids and vice versa. FACILITATED DIFFUSION SHOWS COMPETITIVE INHIBITION The specificity of carrier-mediated facilitated diffusion also gives rise to competitive inhibition. Compounds that closely approximate the shape of the natural substrate may also bind to the carrier and be transported across the membrane. Since the carrier cannot carry both compounds at the same time, the transport of the natural substrate is reduced by the presence of Diffusion through aqueous pores or through the lipid barrier of membranes or by facilitated diffusion is called passive transport because none of these mechanisms requires “outside” energy. These flows occur spontaneously. What this means is that the energy that drives them is contained within the solutions themselves. This does not mean that they occur rapidly, but only that they occur naturally without the addition of any “outside” force. The rate at which they occur depends on the mechanisms of transfer. The analysis of the mechanism gives us additional information, such as what determines and regulates the rate. The overall ΔG for facilitated diffusion is the ΔG for the sum, which is the same in Eqns [2.5.15] and [2.5.17]. The net ΔG 5 nRT ln CL/CR. Thus the participation of the carrier, which remains unchanged by the transport reaction, does not alter the reaction energetics at all, whereas it does alter the reaction kinetics. The carrier is a catalyst. It speeds up the reaction, which in this case is transport, without entering into the stoichiometry of the reaction. This is an example of how thermodynamic 165 166 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Example 2.5.1 Specificity of Transport There are a variety of transporters for glucose that are called GLUT (for glucose transporter). The GLUT-1 transporter imports glucose into a variety of cell types. This is an integral membrane protein with a molecular weight of 45 kDa. Its Km for glucose is 1.5 mM. It will also transport L-glucose with a Km of 3000 mM. Glucose is typically about 100 mg% in the extracellular fluid. At what fraction of Qmax will glucose be transported at this concentration? The normal plasma [glucose] is given as 100 mg%, which is 100 mg of glucose per 100 mL of plasma. This is Pg 5 100 mg/ 100 mL 3 1000 mL L21 5 1 g L21. Since the molecular weight of glucose is 180 Da, its gram molecular weight is 180 g mol21 and the normal plasma glucose concentration is Pg 5 1 g L21 =180 g mol21 5 0:0056 M 5 5:6 mM The rate of transport, assuming zero-trans glucose, is given by Eqn [2.5.16] as Qtrans 5 ½5:6 mM=ð1:5 mM 1 5:6 mMÞQmax 5 0:789 Qmax At plasma [glucose], the transporters are nearly saturated and the rate of transport could be increased mostly by affecting the number of transporters, i.e., increasing Qmax. What would the L-glucose rate of transport be at the same concentration as D-glucose? Qtrans L-glucose 5 ½5:6 mM=ð3000 mM 1 5:6 mMÞQmax 5 0:0019 Qmax D-Glucose is transported almost 400 times more quickly than L-glucose. Example 2.5.2 Effect of Substrate Concentration on Flux GLUT-1 glucose transporter is the most ubiquitous form of glucose transporter. It is present in high amounts in erythrocytes and endothelial cells, the bloodbrain barrier and in the proximal straight tubule of the nephron. If its Km for glucose is 1.5 mM, how much would transport increase if plasma glucose were increased from 80 mg% to 120 mg%? First, we convert the plasma glucose concentrations to mM. 80 mg% means 80 mg per 100 mL or 80 mg/0.1 L 5 800 mg/L. The molecular weight of glucose is 180 g, so this concentration is equivalent to 0.8 g/180 g mol21/L 5 4.44 mM. Similarly, 120 mg% is 6.67 mM. GLUT2 is another glucose transporter that is present in beta cells of the islets of Langerhans, in the pancreas, and also in the kidney, intestine, and liver. Its Km for glucose is much higher, 17 mM. How much would transport increase if plasma glucose were increased from 80 mg% to 120 mg%? We can use the molar concentrations for glucose that we used before for the GLUT1 calculations. The transport rates are calculated at 80 mg% glucose as Q 5 Qmax 3 4:4 mM=ð17 mM 1 4:44 mMÞ 5 0:207Qmax and at 120 mg% glucose, it would be If the Km for glucose is 1.5 mM, then the transport rate at 80 mg % glucose would be Q 5 Qmax 3 6:66 mM=ð17 mM 1 6:66 mMÞ 5 0:282Qmax Q 5 Qmax 3 4:4 mM=ð1:5 mM 1 4:44 mMÞ 5 0:75Qmax Here the transport rate increases 36% when blood glucose increases 50%. And at 120 mg% glucose it would be Q 5 Qmax 3 6:66 mM=ð1:5 mM 1 6:66 mMÞ 5 0:82Qmax The transport rate increases 9% when the blood glucose increases by 50%. These GLUT1 transporters are insensitive to changes in blood glucose. analysis of the reaction is independent of the mechanism: it tells us about the energetics without telling us anything about the reaction’s path or its rate. The saturability of carrier-mediated facilitated diffusion distinguishes it from the other passive transport mechanisms that show a linear relationship between flow and the concentration difference across the membrane. Neither mechanism can concentrate solute. Flow of material always occurs from the side with the higher concentration to the side with the lower concentration. Thus GLUT1 and GLUT2 serve different functions. GLUT1 operates close to maximal rates relatively independently of blood glucose levels. GLUT2 increases transport almost proportionately with blood glucose, so that GLUT1 is used for basal glucose transport into metabolizing tissues, whereas GLUT2 finds use as part of the sensor apparatus for glucose concentrations in blood. When the concentrations on the two sides of the membrane are equal, no further flow occurs because the two solutions are in equilibrium. IONS CAN BE PASSIVELY TRANSPORTED ACROSS MEMBRANES BY IONOPHORES OR BY CHANNELS So far we have considered passive diffusion of nonelectrolytes. Suppose now that the diffusing species Passive Tr ansport and F acilit ated Di ffusion are electrically charged. Charged species are poorly soluble in the lipid phase, and so they cannot merely dissolve in the lipid on one side of the membrane, diffuse across, and then enter the compartment on the opposite side of the membrane. They need either carriers or channels to get across. IONOPHORES CARRY IONS ACROSS MEMBRANES OR FORM CHANNELS Fungi and bacteria make a class of poison called ionophores. These are molecules that allow ions to cross membranes. The fungi and bacteria make these compounds to kill off competition by disrupting the permeability barrier of their competitors’ membranes. These ionophores are of two types: carriers and channel formers. Figure 2.5.6 illustrates these two types of ionophores. An example of a carrier is A23187. This material is commercially obtained from Streptomyces chartreusis and has weak antibiotic activity against gram-positive bacteria. It is particularly active for divalent cations with a specificity of Mn21 .Ca21 .Mg21 .Sr21 .Ba21 .Li1 .Na1 .K1. It is predominantly used as a carrier for Ca21. Other examples of natural molecules that act as carriers include Diffusable ionophore binds ligand and carries it across the bilayer 1 An example of a channel former is gramicidin A. This is an antibiotic polypeptide containing 15 amino acids that is isolated from the bacterium Bacillus brevis. The molecule appears to form a pore by linking two molecules of gramicidin A across the bilayer. The gramicidin pore appears to behave like a water-filled pore. Other examples of pore-forming antibiotics include amphotericin and nystatin. Amphotericin makes a channel by interacting with cholesterol in cell membranes. ION CHANNELS A variety of integral membrane proteins form channels for ions. These ion channels exhibit some of the characteristics of carriers in that they are highly selective. These channels exhibit other characteristics such as gating. Gating refers to the fact that these channels act as if they have gates that are opened sometimes, allowing ions to cross the membrane, and are closed at other times, preventing ions from moving. The percent of the time the channels are opened is referred to as the open probability, po, and can be regulated in various ways. Some channels open when another molecule binds to the channel. These are ligand-gated channels. Other channels sense the local potential, probably through the presence of charged groups on the channel, and open or close depending on the potential. These are voltage-gated channels. A cartoon of these types of channels is shown in Figure 2.5.7. Voltage gates are usually charged; they move in response to the local electric field 4 CL valinomycin (a K1 ionophore) and nigericin (an H1 ionophore). 3 CR – + + + 2 Lipid bilayer + Interstitial fluid Channel-forming ionophore provides an aqueous path across the membrane FIGURE 2.5.6 How ionophores work. Some ionophores increase the passive diffusion across a lipid bilayer by providing a hydrophilic pocket that binds a solute and sequesters it away from the hydrophobic lipid interior. These ionophores generally show specificity of transport because the pocket binds some ions better than others. For these types of ionophores, the ionophoreligand complex is believed to diffuse across the lipid bilayer, carrying the ligand with the ionophore. Other ionophores form an aqueous channel across the lipid bilayer. These channel-forming ionophores are less specific but still show specificity due to the size and shape of the channel. – + E – – + Lipid bilayer + – + – + – + – + – + – Other parts of the channel confer specificity Cytosol Ligand binding to ligand-gated channels opens or closes a gate FIGURE 2.5.7 Voltage- and ligand-gated channels. Voltage-gated channels typically have a highly charged part of the protein that responds to the local electrical field produced by the separation of charge on the two sides of the membrane. Changes in this electric field alter the disposition of the gate to either open or close access to a hydrophilic pathway across the membrane. Ligand-gated channels bind a regulatory ligand that alters the shape of the channel so as to open or close its pathway. 167 168 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Examples of voltage-gated and ligand-gated channels abound, and a full description of them here is premature because we have not yet studied membrane potential or action potentials. Voltage-gated Na1 and K1 channels allow ions to flow across the membrane only under specific circumstances. The flow of ions is an electric current, because charged ions are moving. The membrane itself is a tiny capacitor, as we saw in Chapter 1.3. The currents going through these channels can discharge the capacitor, changing the voltage across the membrane, or they can charge it back up again. In this way, opening of the fast Na1 channel in neurons causes a brief, pulse-like change in the voltage across the neuronal cell membrane. The membrane potential is reestablished by a later opening of the K1 channels. These actions produce the nerve impulse, the brief change in nerve cell membrane potential that propagates down the nerve and activates its target—either another neuron, a muscle fiber, or some secretory cell. Other voltage-gated channels include the T-, L-, and N-type Ca21 channels. The designation “T” signifies that this channel opens “transiently”; the “L” stands for “long-lasting”; and the N indicates that this type of channel is “neuronal.” The channels generally open upon depolarization of the cell membrane and they specifically transport Ca21. The consequence is that the [Ca21] inside the cell increases in the vicinity of the channel, and this Ca21 binds to cellular elements to change their activity. Ligand-gated channels may be present on the surface membrane and also on interior membranes. The endoplasmic reticulum of many cells contains a large tetrameric protein called the IP3 receptor. This receptor forms a channel for Ca21 across the ER membrane and opens in response to IP3 (inositol trisphosphate) that is liberated from the surface membrane as part of signal transduction. Gating by IP3 causes Ca21 release from the ER and the increased cytoplasmic [Ca21] alters cellular activity. Many ligand-gated channels are present on the surface membrane and respond to neurotransmitters or hormones. These channels are gated by the binding of a chemical rather than by the voltage difference across the membrane. Acetylcholine is the neurotransmitter involved in skeletal muscle neurotransmission. When activated, the motor neuron nerve terminal releases acetylcholine near the skeletal muscle membrane. The acetylcholine binds to nicotinic acetylcholine receptors, so named because of their sensitivity to nicotine, on the muscle membrane. Binding of acetylcholine opens a large conductance pathway mainly for Na1. This causes a depolarization of the muscle membrane that propagates along the muscle surface, eventually activating muscle contraction. WATER MOVES PASSIVELY THROUGH AQUAPORINS Passive transport of water across biological membranes also occurs through water channels. These are tiny pores formed by proteins called aquaporins. There are a variety of aquaporins and they are present on virtually every cell membrane. AQP1 has a molecular weight of 29 kDa and forms a channel by the association of four monomers. In some membranes, the number of aquaporins is physiologically regulated so that water movement through the cell can be regulated. This is particularly important in the kidney, because the kidney has the final job of retaining water when it is scarce and excreting it when it is in excess. Although water obeys Fick’s Laws of Diffusion, its movement is dominated by pressure-driven flow. SUMMARY Materials cross biological membranes by a variety of mechanisms including passive transport, active transport, and osmosis. Passive transport mechanisms require no input of metabolic energy. Because of this, passive transport always entails the movement of materials from regions of high concentration to regions of low concentration. The free energy change per mole in the reaction SL-SR is Δμ 5 RT ln½CR =CL where CR and CL are the concentrations of S on the right- and left-hand sides of the membrane, respectively. If CR , CL, then Δμ , 0 and the reaction proceeds from the high concentration (CL) to the low concentration (CR). Passive diffusion across membranes is characterized by a linear relationship between the rate of transport and the concentration difference across the membrane: Js 5 pΔC where Js is the macroscopically observed flux and p is the permeability. This equation holds true if we envision the membrane as a microporous membrane in which diffusion occurs through tiny pores, or if we envision the solute as dissolving in the lipid bilayer and diffusing across it. The dependence of p on the microscopic characteristics of the membrane differs in these two models. For a microporous membrane p 5 nπa2 D=δ where n is the number of pores per unit area, a is the radius of the pore, D is the diffusion coefficient of the solute, and δ is the thickness of the membrane. For a solute dissolving in the lipid bilayer p 5 KDlipid =δ where K is the partition coefficient of the material in the lipid phase, Dlipid is the diffusion coefficient in the lipid bilayer, and δ is the thickness of the bilayer. Some membrane proteins bind solutes and provide an alternative path across membranes. The alternative path facilitates the diffusion of the solute across the membrane. These proteins are carriers for the solutes. The kinetics of transport shows specificity, saturation, Passive Tr ansport and F acilit ated Di ffusion and competition with similar solutes. The overall transport rate often obeys an equation of the form Qtrans 5 Qmax C=½Km 1 C 3. where Qmax is the maximum transport rate, typically limited by the number of carriers in the membrane, and Km is a measure of the dissociation constant of the carrier for the solute. 4. Passive transport mechanisms include lipid dissolution and diffusion, facilitated diffusion, ligand-gated channels, voltage-gated channels, diffusion-mediated ionophores, and pore-forming ionophores. 5. 6. REVIEW QUESTIONS 1. Why is the gradient for a diffusive process linear at steady state? 2. For a microporous membrane, what effect would increasing the number of pores have on diffusive flux across a membrane? What effect would increasing the size of the pores have? What effect should result from increasing the 7. 8. 9. 10. size of the diffusing solute? Increasing the temperature? For a solute dissolution model, what effect would increasing the partition coefficient have? Increasing the particle size of the diffusing solute? Increasing the temperature? Why should two models as different as the microporous membrane and solute dissolution model have identical relationship between J and ΔC? Under what conditions is the free energy for transfer across a membrane for a solute equal to zero? How does the function relating rate of transport to concentration differ between facilitated diffusion and simple diffusion? Why does simple diffusion not show specificity or competition? What two general mechanisms are used to regulate the open or closed state of channels? What is a channel? How would you determine Km for facilitated diffusion? Qmax? 169 2.6 Active Transport: Pumps and Exchangers Learning Objectives G G G G G G G G G Write the equation for the electrochemical potential for an ion Give the approximate concentration of the major ions inside and outside of a heart cell Be able to calculate the free energy change for ion movement across the cell membrane under given conditions of membrane potential and ion concentrations Be able to calculate the free energy change for ion transport when coupled to other processes Distinguish between active and passive transport Distinguish between primary and secondary active transport Give an example of a primary and secondary active transport mechanism Distinguish between P-type, V-type, F-type, and ABC-type active transporters Define the terms: symport, cotransporter, antiport, and exchanger THE ELECTROCHEMICAL POTENTIAL DIFFERENCE MEASURES THE ENERGETICS OF ION PERMEATION Here we consider the transport of four ions: Na1, K1, Cl2, and Ca21, across the surface membrane of heart cells at rest. These ions have different concentrations inside and outside of the cell, and are also subjected to electrical forces because there is a potential difference across the cell membrane and thus an electric field within the membrane. The cardiomyocytes maintain a resting membrane potential. We will learn later how this is established, but for now it is enough to know that there is a separation of charge in the outside and inside compartments. At rest, there is an accumulation of negative charges inside the cell and positive charges outside. This separation of charges gives rise to the membrane potential, which is always taken as the difference between the electrical potential inside and outside the membrane: ½2:6:1 Δψ 5 ψi 2 ψo The membrane potential is sometimes identified with the variables ψ, Vm, or Em. The situation is described in Figure 2.6.1. Let us first take the case of Na1. The concentration of 1 170 Na outside the cell is about 145 mM and inside the cell it is 12 mM. The higher concentration outside of the cell favors a net Na1 flow from outside to inside, driven by diffusion. Further, the negative potential inside the cell also favors Na1 movement into the cell. We write the free energy change per mole for the movement of Na1 from out to in as ½2:6:2 Δμ 5 μi 2 μo Recall here that the free energy change is the free energy of the final state (inside the cell, in this case) minus the free energy of the initial state (outside the cell in this case). When we are dealing with the free energy per mole, we write μ. This is an intensive property which is defined by the conditions and not by the extent of the cell or its membrane or of the amount of material being transported. The free energy itself is an extensive property that depends on how much material is being transported. We can insert the definition of electrochemical potential that we justified earlier (see Eqn [1.7.14]): ½2:6:3 μx 5 μ0x 1 RT ln Cx 1 zx ℑψx where the subscript x denotes substance x to avoid confusion with the subscript i or o which denotes the inside or outside of the cell, respectively. Substituting in for the conditions of the inside and outside of the cell, we find Δμ 5μi 2μo 5μ0Nai1RT ln½Na1i1ℑψi2μ0Nao2RT ln½Na1o 2ℑψo ½Na1i 5RT ln 1ℑðψi 2ψo Þ ½Na1o ½2:6:4 The standard free energy per mole (μ0) cancels out because it is independent of condition. The μ0 is the part of the electrochemical potential that incorporates the chemical energy involved in the formation of bonds. Since in this case there is no chemical transformation (the Na1 ion is not chemically transformed in any way; it is simply transported from one side of the membrane to the other), Δμ0 5 0. Also, the z in the formula for electrochemical potential is the integer charge, which is 11 for the Na1 ion. Remember here that ln is the natural logarithm, not logarithm base 10. The calculation of Δμ for the conditions of the cell shows that Δμ , 0 (see Example 2.6.1). This means that the reaction as written is spontaneous: it will occur in the direction written passively, without additional forces. Our analysis of the energy does not tell us © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00016-1 Active Tr ans port: Pumps a nd Exchangers anything about how fast the process will occur, because the thermodynamic analysis is independent of the mechanism, and the mechanism is what determines the rate. What we can say, however, is that a channel for Na1 that opens in this membrane will cause a rapid inflow of Na1 from the extracellular fluid into the cell. Net flow would not occur in the opposite direction. What we can also say is that if the membrane has any non-zero permeability to Na1, the flux will be from the extracellular fluid (the outside) into the cell. Now let us consider what happens with Ca21. In a completely analogous way, we write the change in free energy for Ca21 entry into the cell as Δμ 5 μi 2 μo 5 RT ln½Ca21 i 1 2ℑψi 2 RT ln½Ca21 o 2 ℑψo 5 RT ln ½Ca21 i 1 2ℑðψi 2 ψo Þ ½Ca21 o Membrane 160 150 [Na+]o 145 mM 140 130 Concentration (mM) 120 110 [Cl– ]o100 mM 100 90 80 70 60 50 40 30 20 10 0 [K+]o 4 mM [Ca2+]o1.2 mM + + + + + + + + + + + + + + + + + + + + + + + + + + + + Cytosol – – – – – – – – – – – – – – – – – – – – – – – – – – – – Note that this process (see Example 2.6.2) has a negative Δμ, so it also occurs spontaneously. In this case, the free energy change is much more negative. This is due to the fact that the concentration gradient for Ca21 contains more energy and the electrical energy gain is twice as great because each Ca21 ion has twice the charge of an Na1 ion. A channel for Ca21 on the cell membrane would let Ca21 into the cell under these conditions. Now let us calculate the free energy change for K1 entry into the cell. The formula is: ½2:6:6 ½2:6:5 Extracellular fluid Here the 2 in the equation arises because Ca21 has two positive charges per ion. The two charges correspond to zx in Eqn [2.6.3]. The electrical force on a Ca21 ion is twice the force on an Na1 ion in the same electric field and thus movement produces twice the energy. [K+] i 155 mM Voltmeter Δμ 5 μi 2 μo 5 RT ln½K1 i 1 ℑψi 2 RT ln½K1 o 2 ℑψo 5 RT ln ½K1 i 1 ℑðψi 2 ψo Þ ½K1 o In this case, Δμ is positive (see Example 2.6.3). This means that K1 under these conditions does not passively enter the cell. Rather, the spontaneous process is K1 exit from the cell. If a channel specific for K1 was to open in the membrane, K1 would leave the cell. In this last example, let us calculate the free energy associated with Cl2 entry into the cell. Again we write the difference in the electrochemical potentials: –80 mV – Δμ 5 μi 2 μo 5 μ0Cli 1 RT ln½Cl2 i 1 ð2 1Þℑψi 2 μ0Clo 1 RT ln½Cl2 o 2 ð2 1Þℑψo ½Cl2 5 RT ln 2 i 2 ℑðψi 2 ψo Þ ½Cl o [Na+]i 12 mM [Cl–]i 5 mM [Ca2+]i 0.1 µM FIGURE 2.6.1 Concentrations of Na1, K1, and Ca21 and the resting membrane potential across the resting cardiac muscle cell membrane. The subscript “o” refers to the “outside” of the cell; “i” denotes the “inside” compartment. ½2:6:7 Note that the valence on the Cl2 ion is negative and it is entered that way in the equation. According to the calculation (see Example 2.6.4), at rest [Cl2] is distributed at equilibrium across the cell membrane. That is, there is no free energy change for Cl2 transport across the resting muscle membrane. These calculations show how the electrochemical potential calculates the energetics of transport, and further EXAMPLE 2.6.1 Free Energy of Na1 Transport For the conditions shown in Figure 2.6.1, calculate the free energy of transport of Na1 from outside to inside the cardiomyocyte. Inserting the values for the concentrations into Eqn [2.6.4] and using R 5 8.314 J mol21 K21 (51.987 cal mol21 K21; 1 J 5 0.239 cal) and T 5 310 K, and remembering that the faraday (ℑ) is 9.649 3 104 C mol21, we get Δμ 5 8:314 J mol21 K21 3 310 K ln½ð12 3 1023 MÞ=ð145 3 1023 MÞ 1 9:649 3 104 C mol21 3 ð2 0:080 VÞ 5 2 6:42 kJ mol21 2 7:72 kJ mol21 5 2 14:14 kJ mol21 5 2 3:38 kcal mol21 171 172 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION EXAMPLE 2.6.2 Free Energy of Ca21 Transport For the conditions shown in Figure 2.6.1, calculate the free energy of transport of Ca21 from outside to inside the cardiomyocyte. Inserting the values for the concentrations and membrane potential into Eqn [2.6.5], we get: Δμ 5 8:314 J mol21 K21 3 310 K ln½ð0:131026 MÞ=ð1:231023 MÞ 1 2 3 9:649 3 104 C mol21 3 ð2 0:080 VÞ 5 2 24:2 kJ mol21 2 15:4 kJ mol21 5 2 39:6 kJ mol21 5 2 9:46 kcal mol21 EXAMPLE 2.6.3 Free Energy of K1 Transport For the conditions shown in Figure 2.6.1, calculate the free energy of transport of K1 from outside to inside the cardiomyocyte. We can insert the values for [K1]i 5 155 3 1023 M and [K1]o 5 4 3 1023 M into Eqn [2.6.6] to calculate Δμ: Δμ 5 8:314 J mol21 K21 3 310 K ln½ð155 3 1023 MÞ=ð4 3 1023 MÞ 1 9:649 3 104 C mol21 3 ð2 0:080 VÞ 5 2 9:43 kJ mol21 2 7:72 kJ mol21 5 1 1:71 kJ mol21 5 1 0:41 kcal mol21 EXAMPLE 2.6.4 Free Energy of Cl2 Transport For the conditions shown in Figure 2.6.1, calculate the free energy of transport of Cl2 from outside to inside the cardiomyocyte. 2 23 We can insert the values for [Cl ]i 5 5 3 10 [Cl2]o 5 100 3 1023 M into Eqn [2.6.7] to calculate Δμ: M and show that both the electrical and diffusive forces enter into the equations that determine the direction of ion flow. Things get a bit more complicated and interesting when ion flow is coupled to other processes. ACTIVE TRANSPORT MECHANISMS LINK METABOLIC ENERGY TO TRANSPORT OF MATERIALS All of the ion movements we have discussed so far involve movement “down” the electrochemical gradient: the free energy was higher for the initial condition than for the final condition, and the free energy change Δμ 5 μfinalμinitial is negative (Δμ , 0). Passive diffusion through pores or the lipid bilayer, carriers, and channels are all passive. This does not mean that energy is not involved. What it means is that the energy does not derive from metabolism. The energy comes from the solutions themselves. However, cells also concentrate some materials by moving them from a region of low electrochemical potential to a region of higher electrochemical potential. This movement has Δμ . 0. It can occur spontaneously only when the Δμ 5 8:314 J mol21 K21 3 310 K ln½ð5 3 1023 MÞ=ð100 3 1023 MÞ 2 9:649 3 104 C mol21 3 ð2 0:080 VÞ 5 2 7:72 kJ mol21 1 7:72 kJ mol21 5 0 kJ mol21 5 0 kcal mol21 positive Δμ for transport is coupled to another process with a more negative Δμ. Typically this process is ATP hydrolysis. Primary active transport moves materials against an electrochemical gradient by the direct involvement of ATP hydrolysis. Examples of molecules that are involved in active transport include the ion pumps: Na, K-ATPase, Ca-ATPase, and H-ATPase. Secondary active transport moves materials against an electrochemical gradient by the indirect involvement of ATP hydrolysis. ATP is used to establish an electrochemical gradient for something, usually Na1, and the energy stored in the electrochemical gradient for Na1 is then used to pump material “uphill.” Examples of secondary active transport are the Naglucose cotransport in the intestinal epithelium and renal proximal tubule and the NaCa exchange in the heart surface membrane. NA,K-ATPASE IS AN EXAMPLE OF PRIMARY ACTIVE TRANSPORT The analysis of free energy changes on ion movement that we performed earlier in this chapter indicated that Active Tr ans port: Pumps a nd Exchangers EXAMPLE 2.6.5 Calculate the Free Energy for Operation of Na,K-ATPase Δμ For the conditions of the cell, calculate the free energy for the Na,K-ATPase. We also calculated 1 1:71 kJ mol21 . We have already calculated that Δμ for ATP hydrolysis is 257.1 kJ mol21 (see Chapter 1.7). Inserting the values for Δμ, we get Na1 o going into the cell: ΔμNao -Nai 5 We have calculated Δμ for 2 14:14 kJ mol21 . This is the opposite process of what the pump does. Thus Δμ for Na1 exit is ΔμNai -Nao 51 14:14 kJ mol21 . the distribution of Na1, K1, and Ca21 is far from equilibrium. This implies that the cell actively maintains these concentrations away from equilibrium, or else the equilibrium distribution would eventually occur. The mechanism responsible for maintaining the resting concentrations of Na1 and K1 is the NaK pump. This pump moves three Na1 ions out of the cell at the same time that it transports two K1 ions into the cell. The movement of these ions is coupled to the hydrolysis of ATP. The whole process is written as 1 1 1 ½2:6:8 ATP 1 3Na1 i 1 2K o .ADP 1 Pi 1 3Nao 1 2K i The Δμ for this entire process is equal to the sum of the Δμ for three separate processes: ½2:6:9a ATP.ADP 1 Pi ½2:6:9b 1 3Na1 i .3Nao ½2:6:9c 1 2K1 o .2K i The overall free energy for this coupled process is written as ΔμNa;K -ATPase 5 ΔμATP.ADP1Pi 1 3ΔμNai .Nao 1 2ΔμKo .Ki ½2:6:10 for K1 entry ΔμKo -Ki 5 as ΔμNa;K -ATPase 5 2 57:1 kJ mol21 1 3 3 14:14 kJ mol21 12 3 1:71 kJ mol21 5 2 11:26 kJ mol21 3Nao E2 3Nao P 2Ko E2 E2 P P 2Ko E2 2Ko E2 2Ki P Plasma membrane E1 3Nai E1 P ATP ADP 3Nai E1 E1 ATP 3Nai ATP 2Ki FIGURE 2.6.2 Modified post-Albers scheme for the reaction mechanism of Na,K-ATPase. Follow the reaction scheme in the direction of the arrows and you will see that the net reaction is the hydrolysis of ATP and the transport of three Na1 ions from in to out and two K1 ions from out to in. The processes are coupled in the reaction mechanism of the pump. The pump cannot hydrolyze ATP without binding Na1 and K1 in sequence. Neither can Na1 be transported without K1 transport and ATP hydrolysis. The coupling is made possible in part by the phosphorylation of the enzyme at an aspartic acid residue. This is shown by EP in the diagram. The formation of a phosphoenzyme is common to the P-type active transport pumps. The negative Δμ that we have calculated for the Na,KATPase indicates that the Na,K-ATPase reaction will occur spontaneously, but it will not tell us at what rate. The rate is a consequence of the mechanism of the pump. Recall that Δμ is the free energy per mole. In this case, it is the free energy per mole of completed reaction with the stoichiometry given by the overall reaction in Eqn [2.6.8]. NA,K-ATPASE FORMS A PHOSPHORYLATED INTERMEDIATE ΔμNa,K-ATPase 5 211.26 kJ mol21 (see Example 2.6.5) means that the net change in free energy per mole of reaction, not per mole of Na1 or K1 or ATP, is 211.26 kJ. This is excess free energy of ATP hydrolysis beyond that required to transport Na1 and K1. ATP hydrolysis has a total of 57.1 kJ of energy per mole of ATP that can be harnessed to do work, and the Na,KATPase uses 45.84 kJ of energy per mole of reaction to do electrochemical work, under cell conditions. According to this result, there is enough energy in ATP hydrolysis to drive the Na,K-ATPase reaction. These calculations hold for the resting cell under the conditions we have investigated. The free energy for the Na,KATPase reaction changes with changes in cellular [Na1], [K1], Δψ, [ATP], [ADP], or [Pi]. Changes in [ATP], [ADP], or [Pi] alter the energy available to the Na,KATPase to do work. Changes in [Na1], [K1], or Δψ alter the energy necessary for transport. The mechanism of ion pumps is generally complicated. It is useful to think of the enzyme as being characterized by a limited number of conformations to which we can give identifying labels. The reaction mechanism is then viewed as the sequential steps that occur among these conformations to achieve ATP hydrolysis and ion transport. Transformations between these conformations are determined by rate constants. Each step has two rate constants, one for the forward and the other for the reverse reaction. A simplified scheme for Na,K-ATPase is shown in Figure 2.6.2. THE NA,K-ATPASE IS ELECTROGENIC The overall reaction of the Na,K-ATPase shown in Eqn [2.6.8] indicates the stoichiometry of 3 Na1 being transported out of the cell and 2 K1 ions being transported into the cell. Thus the numbers of charges moving in 173 174 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION PMCA Na,K-ATPase 3Na+ 2K+ H,K-ATPase Ca2+ H + K+ Plasma membrane ADP + Pi ATP ATP 3Na+2K+ ATP Ca2+ ADP + Pi Ca2+ ADP + Pi ATP H+ ATP H+ K+ ADP + Pi ADP + Pi ATP Ca2+ ADP + Pi Vacuolar H-ATPase SERCA SPCA Vesicle membrane Ca2+ H+ Ca2+ Endoplasmic reticulum membrane FIGURE 2.6.3 Some types of primary active transport mechanisms. Some are located on the plasma membrane of specific cell types; others, such as the smooth endoplasmic reticulum Ca-ATPase (SERCA), are located on subcellular organelles. All of the primary active transporters hydrolyze ATP. PMCA is the plasma membrane Ca-ATPase; SERCA is the smooth endoplasmic reticulum Ca-ATPase; SPCA is the secretory pathway Ca-ATPase. the two directions are not equal, and each turnover of the pump corresponds to the movement of one net positive charge out of the cell. Thus the operation of the pump involves a transmembrane current and a net separation of charge. For this reason, the pump is called electrogenic—it generates an electric current and produces an electrical potential. As we will see later, its contribution to the resting membrane potential is small, but it is not zero. THERE ARE MANY DIFFERENT PRIMARY ACTIVE TRANSPORT PUMPS There are a variety of primary active transport pumps encoded by the human genome. These can be classified into four major groups. P-type ATPases: These all form phosphorylated intermediates in the pump mechanism, like that shown for the Na,K-ATPase in Figure 2.6.2. Among these are gastric H1-ATPase that is responsible for acidification of the stomach contents; Na1,K1-ATPase that is responsible for maintaining ionic gradients in most cells; PMCA (for plasma membrane calcium ATPase) responsible for pumping Ca21 out of cells; the SERCA family of pumps, where SERCA stands for smooth endoplasmic reticulum Ca21ATPase, which is responsible for removing Ca21 from the cytosol of a variety of cell types and placing it in storage in internal sacs within cells; and SPCA, the secretory pathway Ca-ATPase, which pumps Ca21 and other ions, such as Mn21 and Zn21, into the Golgi to bind to secretory proteins. The SPCA is distinguishable functionally from the SERCA pumps because the SERCA pumps are inhibited by thapsigargin at low concentrations whereas SPCA is not. Figure 2.6.3 shows some of these primary active transporters. V-type ATPases: Membranes of lysosomes and secretory vesicles contain a vacuolar-type H1-ATPase that pumps H1 ions from the cytoplasm into the vesicles. This V-type H1-ATPase differs from the gastric H1-ATPase in that it does not require K1. The structure and mechanism of V-type ATPases differs from the P-type active transporters. F-type ATPases: These are more commonly referred to as ATP-synthetases, because they usually work in the reverse mode to make ATP rather than hydrolyze it for the purpose of transport. The main example in the human is the F0F1ATPase of the inner mitochondrial membrane, which is discussed in Chapter 2.10. It uses the electrochemical gradient of H1 to make ATP, but it can also hydrolyze ATP. ABC transporters: The ABC here stands for “ATP-binding cassette.” This is a large family of proteins that engage in the primary active transport of a wide variety of solutes. THE NACA EXCHANGER AS AN EXAMPLE OF SECONDARY ACTIVE TRANSPORT According to our earlier calculation, Δμ for Ca21 entry into the heart muscle cell was 239.6 kJ mol21. If left to Active Tr ans port: Pumps a nd Exchangers itself Ca21 would slowly leak into the cell and disturb the resting [Ca21]i. Heart cells have at least two mechanisms for pumping the Ca21 back out. One of these, mentioned above, is a Ca-ATPase that directly couples ATP hydrolysis to the outward transport of a single Ca21 ion. This is the PMCA Ca21 pump. A second mechanism is called the NaCa exchanger. Its stoichiometry, or relative mole numbers, is 3Na:1Ca. The overall reaction is 21 1 21 3Na1 o 1 Cai .3Nai 1 Cao ½2:6:11 The Δμ for this entire process is equal to the sum of the Δμ for two separate processes: 1 3Na1 o .3Nai ½2:6:12 21 Ca21 i .Cao So we write ½2:6:13 ΔμNa;Ca exchange 5 3ΔμNao .Nai 1 ΔμCai .Cao According to this result (see Example 2.6.6), there is enough energy in the Na1 electrochemical gradient to drive Ca21 out of the cell. However, it requires coupling the entry of three Na1 ions for each Ca21 ion that exits the cell. If the Na1Ca21 exchange was to couple the movement of two Na1 to each Ca21, there would be insufficient energy to drive Ca21 efflux, and the exchanger would actually work in reverse mode, with the Ca21 gradient driving the efflux of Na21. The Ca21 efflux that occurs through the NaCa exchanger is “uphill,” meaning that it requires energy. Therefore, it is an active transport. The energy, however, does not come from ATP hydrolysis directly. It comes from the energy stored in the electrochemical gradient of Na1. This energy, in turn, comes from the operation of the Na,K-ATPase that establishes and maintains the Na1 and K1 gradients across the cell membrane. Therefore, the NaCa exchange is an example of secondary active transport. It requires energy from a source outside of the solutes themselves. The energy is supplied by the Na1 gradient. The Na1 gradient is established using ATP hydrolysis. SECONDARY ACTIVE TRANSPORT MECHANISMS ARE SYMPORTS OR ANTIPORTS The NaCa exchanger, NCX, described above is an example of an antiport. It has this description because the two materials being transported go in opposite directions. Such a device is also called an exchanger or a counter-transporter. There are a variety of antiport secondary active transport mechanisms, as summarized in Figure 2.6.4. Thus far we have largely discussed cationic transporters, those that transport the cations. Cations are ions that migrate toward the cathode, the negatively charged electrode, and so cations are positively charged ions. Devices also transport anions or negatively charged ions. An important example of these is the Cl2HCO32 exchanger. In the red blood cell membrane this exchanger is the AE1 protein, which comprises a large fraction of the integral proteins of the erythrocyte membrane. This anion transporter exchanges Cl2 for HCO32 in the ratio of 1:1. This EXAMPLE 2.6.6 Calculate the Free Energy for Operation of the NaCa Exchanger We have already calculated Δμ for Na1 entry as: ΔμNao -Nai 5 2 14:14 kJ mol21 We have also calculated Δμ for Ca21 entry as ΔμCao -Cai 5 2 39:6 kJ mol21 . ΔμNCX 5 3 3 ð2 14:14 kJ mol21 Þ 1 39:6 kJ mol21 5 2 2:8 kJ mol21 Symports (=cotransporters) Antiports (=exchangers) NCX AE Ca2+ 3Na+ HCO3– Cl– But in this case we are dealing with Ca21exit, which has Δμ 5 139.6 kJ mol21. The overall Δμ for the Na,Ca exchanger is thus NHE H+ Na+ NIS B0AT1 Neutral Na+ AA 2Na+ I– SGLT1 NKCC2 2Na+ Glucose Na+ 2Cl– K+ 2Na+ Glucose Na+ 2Cl– K+ Furosemide Bumetanide Plasma membrane Ca2+ 3Na+ HCO3– Cl– H+ Na+ Na+ Neutral 2Na+ I– AA FIGURE 2.6.4 Examples of secondary active transport. NCX means “Na calcium exchanger”; AE means “anion exchanger”; NHE means “Na H exchanger”; B0 is a specific name of a type of Naamino acid transporter, of which there are several types; NIS means “Na iodine symport”; SGLT means “sodium glucose linked transporter.” Many of these transporters exist in multiple forms or isoforms. NKCC2 means “Na, K two chloride cotransporter”. 175 176 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION exchanger is important, as you will see later, in helping the erythrocytes to carry waste CO2 (see Chapter 6.4). Other secondary active transport mechanisms transport two materials in the same direction and therefore are categorized as symports, also called cotransporters. Examples include the Naglucose transporter in the intestine and kidney membranes that transports glucose in the lumen of these organs into the absorptive cells lining the lumen. These cells also contain a number of Naamino acid transporters that do the same thing as the glucose transporter: they transport amino acids from the lumen into the cells. The amino acids and glucose transported in this way are then transported into the blood by another mechanism, usually by facilitated diffusion. Some examples of symports are shown in Figure 2.6.4. Secondary active transporters and facilitated diffusion proteins are classified in the family of solute carriers (SLC). There is a wide variety of these proteins, over 300 of them, and Figure 2.6.4 shows just a sampling of them. Appendix 2.6.A2 describes the nomenclature of these transport proteins. SUMMARY The movement of ions across cell membranes involves a change in free energy that depends on the concentration of the ion on both sides of the membrane, the charge on the ion, and the membrane potential. The free energy change can be calculated as Δμ 5 μfinal 2 μinitial where μ 5 μ0 1 RT ln C 1 zℑψ where R is the gas constant 5 8.314 J mol21 K21, z is the charge on the ion, ℑ is the faraday 5 9.649 3 104 C mol21, and T is the temperature in Kelvin. At rest in muscle cells, the free energy per mole (Δμ) for Na1 entry into the cell is negative, meaning that it occurs spontaneously. Similarly, Δμ for K1 exit is negative, whereas Δμ for Cl2 is near zero, implying that Cl2 distribution is near equilibrium. Cells maintain the Na1 and K1 gradients by actively pumping out Na1 ions and pumping in K1 ions. This movement of ions requires free energy that is supplied by the energy in the terminal phosphate bond of ATP. The Na,K-ATPase couples the outward movement of three Na1 ions and the inward movement of two K1 ions to the hydrolysis of 1 ATP molecule. The enzyme mechanism is responsible for this coupling. Under cell conditions, the overall Δμ for the Na,K-ATPase is negative because Δμ for ATP hydrolysis is more negative than the combined positive Δμ for Na1 and K1 transport “uphill.” The Na,K-ATPase is an example of primary active transport in which the transport of ions is directly linked to the hydrolysis of ATP. Other transporters couple the positive Δμ for solute transport with the negative Δμ for Na1 entry into the cell. These are secondary active transporters, because they use energy to concentrate materials but the energy is derived directly from the Na1 gradient and indirectly from ATP hydrolysis. Examples of secondary active transport include the surface membrane NaCa exchanger, Naglucose cotransport, and Naamino acid cotransport. Transporters carrying materials in the same direction are called symports or cotransporters. Those carrying materials in opposite directions are called antiports or exchangers. Multiple examples of both classes occur in the body. Clinical Applications: SGLT2 Inhibitors and Diabetes Mellitus Diabetes mellitus is characterized by an abnormally high plasma glucose concentration caused by insufficient production of insulin by the pancreas. Insulin is a protein hormone that increases glucose transporters in peripheral tissues (GLUT4) that remove glucose from the circulation. There are two major classifications of persons with diabetes. Those with Type 1 diabetes require insulin injections and have little or no production of insulin, generally caused by a destruction of the beta cells in the pancreas that produce the hormone. Persons with Type 2 diabetes generally produce insulin but the body cells are resistant to the hormone and the circulating levels of insulin are inadequate to lower plasma glucose levels. High blood glucose causes glycosylation of proteins, as measured clinically by HbA1c, glycosylated hemoglobin. Long-term control of plasma glucose is monitored by the HbA1c level. Diabetes mellitus is described in more detail in Chapter 9.4. The kidney produces urine by filtering large volumes of blood and then reabsorbing the desired materials and discarding the rest. Glucose is reabsorbed in the kidney in the proximal tubule of the nephron, the functional unit of the kidney, by SGLT2 present in the apical membrane of the tubule (see Figure 2.6.5). Final transport of glucose into the blood occurs over GLUT2, a facilitated transport mechanism. 90% of glucose reabsorption in the kidney occurs via the SGLT2 glucose uptake mechanism. A relatively new class of drugs inhibits the SGLT2 so that not all of the filtered glucose is reabsorbed, and glucose appears in the urine. These drugs include empagliflozin, canagliflozin, and dapagliflozin. The chemical structures of these drugs are shown in Figure 2.6.6. The term “diabetes” derives from the Greek meaning “to siphon” and this refers to the increased urinary flow in diabetic persons. This is caused by the inability of the kidney to reabsorb all of the glucose, and the excretion of this glucose in a larger volume of water. The effect of the SGLT2 inhibitors is to (1) lower blood glucose levels, (2) reduce HbA1c levels, (3) generally cause a slight loss of weight, and (4) tend to cause dehydration due to increased urine flow. These drugs are of limited value to persons with dysfunctional kidneys (Whalen K, Miller, S, and Onge E, The role of sodiumglucose co-transporter 2 inhibitors in the treatment of type 2 diabetes, Clin. Therap. 37:11501166, 2015). Active Tr ans port: Pumps a nd Exchangers Basolateral membrane Apical membrane Lumen Na+ SGLT2 Basement membrane ADP Na+ K+ ATP Peritubular capillary Na+ K+ FIGURE 2.6.5 Mechanism of reabsorption of glucose from the ultrafiltrate in the lumen of the nephron proximal tubule. SGLT2 transports 2 Na1 ions into the cell along with glucose. The energy is supplied by the electrochemical gradient for Na1 that is established by pumping Na1 out of the cell at the basolateral membrane. Glucose enters the blood through GLUT2, a facilitated diffusion carrier on the basolateral membrane. Glucose Glucose GLUT2 –4 mV 0 mV –7 0mV HO Empagliflozin O HO HO Cl O O OH O HO Canagliflozin HO HO O CH3 OH O F S HO Dapagliflozin HO HO O Cl OH O O CH3 FIGURE 2.6.6 Chemical structure of the SGLT2 inhibitors. Clinical Applications: Oral Rehydration Therapy One of most important public health issues in the world is the availability of clean drinking water. Contaminated water supplies carry cholera and other infectious agents that result in diarrhea and vomiting that cause dehydration that can be fatal, especially in children. Prior to the introduction of oral rehydration therapy (ORT), diarrhea was the leading cause of infant mortality in developing nations. ORT is estimated to have reduced world-wide infant deaths from 5 million per year to 3 million per year (2006 figures). However, diarrhea remains the second leading cause of death in children less than 5 years old (18%, after pneumonia at 19%). The WHO (World Health Organization) and UNICEF (United Nations Children’s Fund, shortened from the original United Nationals International Children’s Emergency Fund) jointly publish guidelines for the composition of oral rehydration solution (ORS). Its current formulation is: 2.6 g NaCl; 2.9 g trisodium citrate dihydrate (Na3C6H5O7∙2 H2O); 1.5 g KCl; 13.5 g glucose per L of solution. The molar ratio of sodium to glucose in this ORS is 1.0, and it is slightly hyposmotic. It should be made with clean water, but when clean water is not available other fluids may be substituted—but not sugar-containing fluids like fruit juices. The effectiveness of ORT relies on the SGLT1 system in the intestine that transports glucose along with 2 Na1 ions into the enterocyte and then into the blood. The transport of nutrients creates an osmotic reabsorption of water along with the solutes. This is enough to counteract the loss of fluid through diarrhea or vomiting. ORT was not used until the 1960s. Before that time, rehydration was accomplished by intravenous fluid administration. In the developing world, IV therapy is not widely available and ORT is a low-technology solution that is much more readily available. 177 178 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION REVIEW QUESTIONS 1. If a Na channel was to open on the surface of a cardiac cell, at rest, in which direction would Na ions travel? If the channels were K-specific, in which direction would K ions travel? If it were Ca21-specific, which way would Ca21 go? If it was a Cl2 channel, which way would Cl2 ions go? 2. Which part of Eqn [2.6.4] gives the energy due to diffusion? Which part gives the energy due to electrical forces? For Na1, which is bigger, the part due to concentration differences or the part due to electrical forces? Which part of the total energy, concentration or electrical, is bigger for Ca21? 3. Where does the energy come from for the movement of materials by passive transport? 4. Where does the energy come from for the movement of materials by primary active transport? 5. Where does the energy come from for the movement of materials by secondary active transport? 6. Do you think that the NCX can go in reverse? 7. Why cannot thermodynamics predict reaction rates? 8. What is a symport? What is an antiport? What is a cotransporter? APPENDIX 2.6.A1 DERIVATION OF THE USSING FLUX RATIO EQUATION The passive flux of an ion in a solution is given by Fick’s First Law of Diffusion with an external force, which was introduced in Eqn [1.7.19]. The onedimensional form of this equation is reproduced here: ½2:6:A1:1 Js 5 2 D @C D @ψ 2 Czℑ @x RT @x where J is the flux, in mol cm22 s21, D is the diffusion coefficient in cm2 s21, C is the concentration (converted to units of mol cm23), R is the gas constant (58.314 J mol21 oK21 5 1.987 cal mol21 oK21), T is the absolute temperature (oK), z is the integer of the charge on the ion ( 6 1 or 2, generally), ℑ is the Faraday (596,489 coulomb mol21) and ψ is the electrical potential, in joules (5volt-coulomb). The quantity @C/ @x is the magnitude of the gradient of C, and @ψ/@x is the magnitude of the electric field. The flux can be obtained by integrating this equation. To begin, we multiply both sides of the equation by an integrating factor, ρ, and we choose ρ so that the right-hand side of the equation becomes an exact differential: @C Czℑρ @ψ 1 ½2:6:A1:2 Js ρ 5 2 D ρ @x RT @x We choose ρ so that the terms in brackets are an exact differential. Thus we want ½2:6:A1:3 ρ @C Czℑρ @ψ dðρ CÞ @C @ρ 1 5 5ρ 1C @x RT @x dx @x @x From comparing the left of Eqn [2.6.A1.3] to the right, we see that multiplication by ρ transforms the equation into an exact differential if @ρ zℑρ @ψ 5 @x RT @x ½2:6:A1:4 We rearrange this to get @ρ zℑ 5 @ψ ρ RT ½2:6:A1:5 A solution to Eqn [2.6.A1.5] is ρ 5 eRT ψ zℑ ½2:6:A1:6 We may insert this result back into Eqn [2.6.A1.2], with Eqn [2.6.A1.3], to obtain zℑ ψ RT d Ce zℑ ½2:6:A1:7 Js eRT ψ 5 2 D dx which may be rewritten as zℑ Js eRT ψ dx 5 ½2:6:A1:8 zℑ ψ 2 D d CeRT If we are considering passive transport across a membrane, we can determine the passive flux by integrating this equation from x 5 0 (one side of the membrane) to x 5 δ, the other side of the membrane for a membrane with thickness δ: ðδ ð@ zℑ zℑ ψ ψ RT RT Js e ½2:6:A1:9 dx 5 2 D δ Ce 0 0 Here we limit ourselves to the steady-state condition. In this case, Js does not vary with distance across the membrane—it is constant. Therefore, Js may be removed from the integral and we get zℑ Ð@ 2 D 0 δ CeRT ψ Js 5 ½2:6:A1:10 Ð δ zℑ ψ RT dx 0 e The numerator in this equation is the integral of an exact differential and can be immediately evaluated between the boundaries. This gives zℑ zℑ ψðδÞ ψð0Þ 2 D CðδÞeRT 2 Cð0ÞeRT Js 5 ½2:6:A1:11 Ð δ zℑ ψ RT dx 0 e The denominator in this equation can be evaluated only if ψ(x) is known. However, generally ψ(x) is unknown. Ussing made the observation that the presence of an active transport mechanism would not obey Eqn [2.6.A1.11], because the flux would not be passive, and this equation describes the passive flux. He further made the observation that we don’t need to know how ψ varies with x if we take the ratio of the unidirectional fluxes. The unidirectional flux is the flux that you would observe if the concentration on the other side was zero. Active Tr ans port: Pumps a nd Exchangers We define here the unidirectional flux Ji-o to be the flux from inside to outside, with x 5 0 on the inside and x 5 δ on the outside. This flux is given from Eqn [2.6.A1.11] by setting C(δ) 5 0, and we obtain ½2:6:A1:12 zℑ eRT ψð0Þ D Cð0Þ Ji-o 5 Ð zℑ δ RT ψ dx 0 e We further define the unidirectional flux Jo-i to be the unidirectional flux from outside with x 5 δ to the inside, with x 5 0. This flux is also given from Eqn [2.6.A1.11] by setting C(0) 5 0; we obtain ½2:6:A1:13 Jo-i 5 zℑ 2 D CðδÞ eRT ψðδÞ Ð δ zℑ ψ eRT dx 0 Here the minus sign conveys a convention that the outward flux is taken as positive. Thus a negative flux simply means that the flux is directed inward, and a positive flux is directed outward. The magnitude of the fluxes is given by the absolute values of the fluxes. If we take the ratio of the unidirectional fluxes, the denominators in each cancel each other out, and we don’t need to do the integration that requires knowledge of ψ(x). Taking the ratio of the two unidirectional fluxes, we arrive at ½2:6:A1:14 zℑ Ji-o D Cð0Þ eRT ψð0Þ 5 zℑ Jo-i 2 D CðδÞ eRT ψðδÞ Using the notation that C(0) 5 C(i), the inside concentration, and C(δ) 5 C(o), the outside concentration, this equation can be simplified to ½2:6:A1:15 Ji-o CðiÞ zℑ ψð0Þ2ψðδÞ eRT 52 CðoÞ Jo-i Using the definition that the difference in potential across the membrane, ψ(0) 2 ψ(δ) 5 Em, the membrane potential, this last Eqn [2.6.A1.15] becomes ½2:6:A1:16 Ji-o CðiÞ zℑ Em eRT 52 CðoÞ Jo-i This last equation is the Ussing Flux Ratio Equation. It describes the expected ratio of the unidirectional fluxes if the ions are transported passively across the membrane. Deviations from the flux ratio equation are taken to indicate that the fluxes are not transported passively. That is, deviations from the expected flux ratio can be taken to indicate that an active transport mechanism is present. Ussing considered other possible deviations from the expected flux ratio such as single-file transport. Hans Ussing (19112000) derived his flux ratio equation in the late 1940s, soon after radioactive isotopes became available to measure unidirectional ion fluxes. He was the first to prove the existence of active transport mechanisms, using the frog skin as a model. The Na,K-ATPase was later discovered, in 1957, by Jens Skou, who earned a Nobel Prize for the discovery. APPENDIX 2.6.A2 NOMENCLATURE OF TRANSPORT PROTEINS HUGO NOMENCLATURE Proteins have “trivial” or common names that were generally first provided by their discoverers. These names have often been changed when a new function for the protein was discovered or its relationship to other proteins was discovered. Since these proteins derive from genes, it is now increasingly useful to use the name for the gene to also describe the resulting protein. The HUGO Gene Nomenclature Committee (HGNC) provides a unique identifier for each gene. This committee is a part of the Human Gene Organization (HUGO). Full and updated databases are available at www.genenames.org. CARRIER CLASSIFICATIONS The transport proteins as described in Chapters 2.5 and 2.6 can be generally classified as belonging to a small number of types. These are: Passive Transporters: Facilitated transporters Ion channels Water channels Active Transporters: Secondary active transporters: exchangers and cotransporters ATPase pumps (P, V, and F-types) ABC (ATP-binding cassette) transporters. These functional classifications do not map precisely onto the HUGO nomenclature. The HUGO classification lumps facilitated transporters and secondary active transporters into one group, the solute carriers, SLC. Carriers are often grouped into families based on sequence homology rather than functionality. An example of this is the sodiumiodine exchanger located in the thyroid gland, which is part of the sodiumglucose cotransport family. SOLUTE CARRIERS The solute carriers include the facilitated diffusional carriers such as GLUT1 and GLUT2 and the secondary active transport carriers such as NCX, NHE, SGLT, NIS, and AE. All members of the SLC superfamily are named according to SLC n X m where SLC indicates the superfamily, n is an integer that denotes the family, X is a letter that denotes the subfamily, and m is a second integer that denotes the isoform. As an example, the facilitated glucose carriers are all members of the SLC2A subfamily of which there are 14 members: SLC2A{1. . .14}. These correspond to their common names, as shown in Table 2.6.A2.1. There are an enormous number of these SLC genes. There are at least 52 families (n) with a total of 396 genes (Σmi) that encode for transporter proteins. It is 179 180 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION TABLE 2.6.A2.1 Nomenclature of the Facilitated Glucose Transporters (Augustin, R. “The protein family of glucose transport facilitators; It’s not about glucose after all.” Life 62:315333, 2010) HUGO Name Common Name Location and Function SLC2A1 GLUT1 Km for glucose 5 1.5 mM; ubiquitous but important for erythrocytes and uptake of glucose into the cerebrospinal fluid (CSF) SLC2A2 GLUT2 Km for glucose 5 17 mM; liver, kidney, intestine, β cells of pancreas; transports glucose out of intestine and kidney cells SLC2A3 GLUT3 Km for 2-deoxyglucose 5 1.4 mM; present in neurons, spermatozoa, placenta SLC2A4 GLUT4 Km for glucose 5 5 mM; skeletal and cardiac muscle, adipose tissue; rate-limiting step in insulin-stimulated glucose uptake into tissues SLC2A5 GLUT5 Km for fructose 5 6 mM; mainly jejunum of small intestine but also found in kidney, brain, muscle and fat; primarily a fructose transporter SLC2A6 GLUT6 brain, spleen, and leukocytes SLC2A7 GLUT7 Km for glucose 5 0.3 mM; apical membrane of intestine and colon; also transports fructose SLC2A8 GLUT8 Km for glucose 5 2 mM; testis, cerebellum, liver, spleen, lung, fat; intracellular without insulin-stimulated translocation to the surface SLC2A9 GLUT9 Km for glucose 5 0.6 mM but also transports urate; kidney and liver and β cells of pancreas; exchanges urate for glucose or fructose SLC2A10 GLUT10 Present in heart, lung, brain, liver, muscle, pancreas, placenta, and kidney; may be intracellular SLC2A11 GLUT11 Three isoforms (GLUT11{A,B,C}); in humans GLUT11 is exclusively expressed in slow-twitch muscle fibers SLC2A12 GLUT12 SLC2A13 GLUT13; HMIT SLC2A14 GLUT14 Does not transport sugar; H1-coupled myoinositol symporter; Km 5 0.1 mM for myoinositol; located intracellularly; highest in brain entirely unreasonable to attempt memorization of all of them. The list of these and other gene families is available in tabular format on the HUGO website. other solutes that need carrying, there are a large number of these kinds of carriers. They belong to the SLC1, SLC3, SLC6, SLC7, SLC36, SLC38, and SLC43 subfamilies. Note that some designations within a family can be confusing in that the function of the carrier is not related to its stated family! For example, GLUT 9 (SLC2A9) is primarily regarded as a urate transporter rather than a glucose transporter, but it is a member of the facilitated diffusion carriers for glucose. The anion exchanger shown in Figure 2.6.4, AE, also has three subtypes (AE13) that correspond to SLC4A13. These are all electroneutral Cl2 HCO32 exchangers. The SLC4 family contains 10 members (SLC4A{1. . .5; 7. . .11} that include Na1 and HCO32 or CO322 cotransporters. A second example of this is found in the sodiumglucose transport family. SGLT1 and SGLT2 are part of the SLC5A subfamily that includes 12 members (SLC5A {1. . .12}. SLC5A1 corresponds to SGLT1 and SLC5A2 corresponds to SGLT2, but SLC5A5 corresponds to the sodiumiodine exchanger (NIS), shown in Figure 2.6.4, that does not transport glucose! The NIS is grouped in this way because of structural similarities in the transport proteins that leads researchers to suppose that they belong to the same subfamily of transporters. SGLT1 is the main transporter in the intestine whereas SGLT2 is the main transporter in the kidney, although SGLT1 is also found in the kidney. The amino acid transporter shown in Figure 2.6.4 has three members (B0AT13) that correspond to genes SLC6A19, SLC6A15, and SLC6A18, respectively. These all carry neutral amino acids. Because there are so many amino acids and also a variety of neurotransmitters and The Na1Ca21 exchanger (NCX) of Figure 2.6.4 also has three members (NCX13) encoded by genes SLC8A13, respectively. There is also a mitochondrial Na1Ca21 exchanger, NCLX, as a gene product of SLC8B1. There are additional transporters for Ca21 including NCKX (Na1Ca21K1 exchanger, SLC24 family) and CCX (Ca21-cation exchanger). The Na1H1 exchanger (NHE) of Figure 2.6.4 has several members. NHE15 are all present on the surface membrane, and NHE3 and NHE5 recycle between the surface and intracellular membranes. They correspond to SLC9A15, respectively. NHE6, NHE7, and NHE9 appear to be located in intracellular membranes along the secretory pathway. They correspond to SLC9A6, SLC9A7, and CLC9A9, respectively. The NaK2Cl cotransporter shown in Figure 2.6.4 comes in two forms, NKCC1 and NKCC2, corresponding to genes SLC12A1 and SLC12A2. Active Tr ans port: Pumps a nd Exchangers There are a large number of other transporters of the SLC gene superfamily that transport neurotransmitters or ions, or vitamins, or metabolites across plasma membranes or internal membranes. A complete listing is available on the HUGO website. ATP-DRIVEN ION PUMPS Names for the genes for the ATP-driven ion pumps have the form ATP n X m complex. The mitochondrial F-type ATPase consists of an F1 and an FO subunit that themselves are complexes of additional subunits. These are encoded by the ATP5 family of genes, with subfamilies denoted by the letters A, B, C, D, E, F, G, H, I, J, L, and O. Somewhat differently, the V-type ATPase that acidifies lysosome contents by pumping in H1 ions is designated ATP6V0 and ATP6V1 for the V0 and V1 subunits, and letters corresponding to the subunits within V0 and V1. where ATP indicates the superfamily, n is an integer that denotes the family, X is a letter that denotes the subfamily, and m is an integer that denotes the member. Many of the ATP-driven pumps consist of multiple subunits, and these are generally organized by having the same family integer, n, and a different subfamily name. For example, the Na,K-ATPase has an α and a β subunit, and these are indicated as ATP1A and ATP1B. There are four varieties of each, so that the α subunit of the Na,KATPase corresponds to ATP1A{1. . .4} and the β subunit has genes ATP1B{1. . .4}. ABC TRANSPORTERS The Ca21 pumps are members of the ATP2 family. ATP2A{1,2,3} correspond to SERCA1, SERCA2, and SERCA3, respectively, that are located on the internal membranes of the cell, the endoplasmic reticulum, or the sarcoplasmic reticulum. The plasma membrane CaATPases, PMCA{1. . .4} are encoded by the ATP2B {1. . .4} genes, respectively. SPCA1 and SPCA2 are encoded by ATP2C1 and ATP2C2. The gastric H1K1ATPase is encoded by ATP4A (the α subunit) and ATP4B (the β subunit}. All of the above-mentioned pumps constitute the P-class of ATP-driven ion pumps. AQUAPORINS The F-type and V-type ATPase ion pumps are much more complex, consisting of multiple subunits and multiple copies of some of these subunits. Therefore, there is no one gene that encodes the entire operating ABC transporters hydrolyze ATP to transport a wide variety of substrates. There are 48 transporters classified in 7 families denoted in this case by letters alone: ABC X m where ABC denotes “ATP-binding cassette”, X is a letter indicating the family, and m is an integer indicating the member. Aquaporins are proteins that increase water movement across biological membranes. They have a molecular weight of around 30 kDa and associate as tetramers, although each monomer has a water channel. Several of the aquaporins will transport other small molecular weight, electrically neutral substrates such as glycerol or urea. There are a variety of aquaporins named AQP n where n is the member of the family. There are 14 members of the family, named AQP{1. . .12}. AQP0 has been renamed MIP for “major intrinsic protein” of the lens, and AQP12 has A and B subtypes. In the HUGO classification, aquaporins are considered to be a subtype of ion channels. 181 2.7 Osmosis and Osmotic Pressure Learning Objectives G G G G G G G G G G G G G G G G Define osmosis and osmotic pressure Write van’t Hoff’s limiting law for osmotic pressure dependent on concentration List the colligative properties of solutions and explain why vapor pressure depression perfectly predicts osmotic pressure Describe the osmotic coefficient for correction of nonideality in solutions Be able to calculate the predicted osmotic pressure for a solution Define hydraulic conductivity or hydraulic permeability Define the reflection coefficient Explain how the hydraulic conductivity depends on the microscopic parameters of the membrane Write an equation for volume flow across a microporous membrane in the presence of hydraulic and osmotic pressures Be able to calculate the hydraulic conductivity and reflection coefficient given appropriate data Describe the origin of the osmotic pressure for microporous membranes Describe how erythrocyte cell volume changes when placed in contact with solutions of varying [NaCl] or [glucose] Contrast the concepts of tonicity and osmolarity Describe the behavior of a perfect osmometer Explain why cells are not perfect osmometers Define RVD and RVI OSMOSIS IS THE FLOW OF WATER DRIVEN BY SOLUTE CONCENTRATION DIFFERENCES Probably no concept is more confusing to beginning students than osmosis and osmotic pressure, partly because it is defined backwards, as you will see. We begin with the experiments of Pfeffer in 1877. Pfeffer made a precipitation membrane in the walls of unglazed porcelain cups by reacting copper salts with potassium ferricyanide. He used the precipitation membrane that resulted to separate a sucrose solution on the inside of the cup from water on the outside. 182 He observed that water flowed from the outside to the inside. This is the primary observation of osmosis. Osmosis refers to the movement of fluid across a membrane in response to differing concentrations of solute on the two sides of the membrane. The word “osmosis” originates from the Greek, meaning “thrust” or “impulse.” Pfeffer also observed that the flow was proportional to the sucrose concentration inside the cup. When water was inside the cup, he observed that a pressure applied to the inside compartment would force water out of the cup, and this flow was proportional to the pressure. When sucrose was inside a closed cup, a pressure would develop inside the cup and this pressure was proportional to the sucrose concentration. He recognized this as an equilibrium state in which the outward filtration of water balanced the inward movement caused by osmosis. He defined the osmotic pressure as the pressure necessary to stop osmotic flow across a barrier that is impermeable to the solute. The osmotic pressure historically is given the symbol π (see Figure 2.7.1). This definition is the key: it defines the osmotic pressure as the pressure needed to stop osmotic flow rather than the pressure that drives osmotic flow. It is also defined only for a semipermeable membrane, one that is impermeable to solute but permeable to water. THE VAN’T HOFF EQUATION RELATES OSMOTIC PRESSURE TO CONCENTRATION Pfeffer’s data (see Figure 2.7.2) showed that the osmotic pressure of solutions was linearly related to their concentration. In 1887, van’t Hoff argued from Pfeffer’s results, and from thought experiments considering gases in equilibrium with solutions, that the osmotic pressure should be given by ½2:7:1 π 5 RTCs where R is the gas constant (0.082 L atm mol21 K21), T is the absolute temperature, and C is the molar concentration of solute in the inner compartment. Equation [2.7.1] is known as van’t Hoff’s Law. It gives the osmotic pressure due to the solute, s. In a mixture, the osmotic pressures due to each solute particle add up, much like Dalton’s Law of Partial Pressures in gases. The result is that X ½2:7:2 π 5 RT Cs © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00017-3 Pressure-driven flow: ΔP > 0 Osmotic flow: ΔP = 0 Qv = A LpΔP Qv = -A LpΔπ Osmotic pressure: Qv = 0 ΔP = Δπ Qv = A Lp(ΔP - Δπ) Applied pressure, PL Applied pressure, PL Piston Piston Piston Pure water Pure water Pure water Pure water Solution Solution Semipermeable membrane Semipermeable membrane Semipermeable membrane FIGURE 2.7.1 Equivalence of hydrostatic and osmotic pressures in driving fluid flow across a membrane. Left panel: An ideal, semipermeable membrane is freely permeable to water but is impermeable to solute. When the membrane separates pure water on the right from pure water on the left, application of a pressure, PL, to the left compartment forces water across the semipermeable membrane. The flow is linearly related to the pressure difference by the area of the membrane (A) and a proportionality constant, LP, that is characteristic of the membrane. This constant is variously called the hydraulic conductivity, hydraulic permeability, or filtration coefficient. Positive QV is taken as flow to the right. Middle panel: The ideal, semipermeable membrane separates a solution on the left from pure water on the right, and water moves to the solution side by osmosis. The flow, QV, is linearly related to the difference in osmotic pressure, Δπ, by the area of the membrane and the hydraulic conductivity, LP. The flow causes expansion of the left compartment and movement of the piston, which is assumed here to be weightless. Right panel: Application of a pressure, PL, to a solution so that ΔP 5 Δπ results in no net flow across the membrane. The osmotic pressure of a solution is defined as the pressure necessary to stop fluid flow when an ideal semipermeable membrane separates pure water from the solution. 3500 Pressure (mmHg) 3000 2500 2000 1500 1000 500 0 0 1 2 3 4 5 6 7 Sucrose concentration (%w/w) FIGURE 2.7.2 Plot of the data of Pfeffer (1877) for the osmotic pressure of sucrose solutions. A copper ferrocyanide precipitation membrane was formed in the walls of an unglazed porcelain cup. The membrane separated a sucrose solution in the inner chamber from water outside the cup. The inner chamber was then attached to a manometer and sealed. The linear relationship between the pressure measured with this device and the sucrose concentration was the experimental impetus for deriving van’t Hoff’s Law. [Data from Pfeffer, W. Osmotische Untersuchungen Studien zur Zellmechanik, Leipzig [translated by G.R. Kepner and E.J. Tadelmann: Osmotic investigations: Studies on cell membranes, 1985, Van Nostrand Reinhold, New York]. 184 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION The concentration in van’t Hoff’s law, ΣCs, refers to the concentration of osmotically active solute particles in the solution. Organic compounds such as glucose typically dissolve to form one particle for each molecule of solute, and for these compounds Cs is the same as the molar concentration. Strong salts, on the other hand, dissociate to form more than one particle for each mole of salt. NaCl, for example, dissociates in solution to form one Na1 and one Cl2 ion. The concentration of osmotically active particles is twice the concentration of NaCl. Similarly, CaCl2 dissociates nearly completely to form one Ca21 ion and two Cl2 ions, and the total concentration of particles is 3 times the concentration of CaCl2. The osmolarity of a solution equals ΣCs and is expressed in osmoles per liter to indicate clearly that we are referring to the number of osmotically active solute particles, called osmolytes, rather than the concentration of the solute. of work such as pressurevolume work. The general equation for the chemical potential is THERMODYNAMIC DERIVATION OF VAN’T HOFF’S LAW ½2:7:5 We have learned that all spontaneous processes (those that occur naturally without any additional forces) are accompanied by a decrease in free energy. We have also learned that the free energy per mole is the chemical potential (see Chapter 1.7). In the case of a solution separated from pure water by a semipermeable membrane, water movement will occur when there is a difference in the chemical potential of water on the two sides of the membrane, such that water movement results in a decrease in the free energy. At equilibrium, when the pressure applied to the solution is equal to the osmotic pressure, the chemical potential of water is equal on both sides of the membrane, and no net movement of water occurs. In the derivation that follows, we consider that a semipermeable membrane separates two compartments. On the right is pure water and on the left is some solution with concentration Cs. We know that there will be flow of water from the pure water to the solution side and that application of pressure, π, to the solution side will stop the flow. What is the relation between π and Cs? We can discover this by looking at the equilibrium condition when osmotic pressure balances hydrostatic pressure. At this point, the free energy change for water across the membrane is zero. THE CHEMICAL POTENTIAL INCLUDES PRESSUREVOLUME WORK The equation we have used for chemical potential of a solute, the free energy per mole, is ½2:7:3 μs 5 μ0s 1 RT ln Cs 1 z`ψ In this equation, z`ψ represents the work done, per mole, in moving the material from the standard state to the condition in which the material is placed. In this case, it is the electrical work. There are other kinds ½2:7:4 μs 5 μ0s 1 RT ln Cs 1 work terms where the work terms include all work (except concentration work, which is included explicitly in RT ln Cs) necessary to bring the material from the standard state to its present state. THE ACTIVITY CORRECTS THE CHEMICAL POTENTIAL FOR INTERACTIONS BETWEEN SOLUTE PARTICLES It turns out that Eqn [2.7.4] is an approximation. In our derivation of the general Fick’s Law, we did not consider some other kinds of interactions, such as solute molecules bumping into each other. The accurate equation is μs 5 μ0s 1 RT ln as 1 work terms where as is the activity of the solute. In the case of osmosis, the work term is the pressurevolume work and there is no electrical work term. At equilibrium, where the pressure across the semipermeable membrane is the osmotic pressure, the chemical potential of water on the two sides of the semipermeable membrane must be equal (because the free energy change at equilibrium is zero). Therefore, we write the equality of chemical potential for water on the left and right sides as μ0w 1 RT ln aw;L 1 V w PL 5 μ0w 1 RT ln aw;R 1 V w PR ½2:7:6 where the subscripts L and R refer to the left and right sides of the semipermeable membrane, μw0 is the chemical potential of liquid water in its standard state (pure water at 1 atm pressure), Vw is the volume of water per mole (the partial molar volume), P is the pressure, and aw is the activity of water. For an ideal solution, the activity of water is its mole fraction: ½2:7:7 aw 5 Xw 5 nw nw 1 ns where Xw is the usual variable denoting the mole fraction of water, and nw and ns are the moles of water and solute, respectively, in any aliquot of the solution. Substituting in for aw and canceling the μw0 on both sides of Eqn [2.7.6], we come to ½2:7:8 V w PL 2 V w PR 5 RT ln Xw;R 2 RT ln Xw;L Since the right-hand solution is pure water, Xw,R 5 1.0 and ln Xw,R 5 0. Thus we have ½2:7:9 V w ðPL 2 PR Þ 5 2 RT ln Xw;L Now the mole fractions of solute and water in a solution must sum to 1.0. This is expressed as ½2:7:10 Xw;L 1 Xs;L 5 1:0 ln Xw;L 5 lnð1 2 Xs;L Þ Os mos is an d Os motic P res sur e In dilute solutions, Xs,L ,, 1.0, so we may approximate ln(1 2 Xs,L) 2Xs,L. Substitution of this result into Eqn (2.7.9) gives ½2:7:11 ðPL 2 PR Þ 5 RT Xs;L Vw The left-hand side of Eqn [2.7.11] is just the osmotic pressure, π, which is equal to the extra pressure that must be applied to the left-hand side in order to establish equality of the chemical potential of water on the two sides of the membrane. From the definition of mole fraction, Eqn [2.7.11] becomes ½2:7:12 ðPL 2 PR Þ 5 π 5 π5 Water Solution Water RT ns V w nw Again for a dilute solution, nwVw V, the volume of the solution. Thus Eqn [2.7.13] gives ½2:7:14 Solution RT ns V w ns 1 nw For a dilute solution, ns ,, nw, so we approximate this result as ½2:7:13 Water vapor ns π 5 RT V Air This last equation is the van’t Hoff equation for the osmotic pressure. This thermodynamic derivation entails two assumptions: the solution is dilute enough to approach ideality and that the solution is incompressible so that the pressurevolume work is VwΔP. It is important to recognize that the van’t Hoff equation is not exact for physiological solutions. Rather, it is an approximation that is strictly true only for dilute ideal solutions. FIGURE 2.7.3 Equivalence of vapor pressure and osmotic pressure. Two beakers containing a solution or pure water are both placed in a single, sealed compartment. The air above the fluids contains air molecules plus water vapor. The partial pressure of water is its contribution to the total pressure, and it is proportional to the water concentration in the gas phase. The vapor pressure is defined as the partial pressure of water in equilibrium with the liquid phase. Water molecules will leave the liquid to moisten a dry gas. At equilibrium, molecules will evaporate from the liquid phase and condense from the gas phase at equal rates, so that a dynamic equilibrium is established. The vapor pressure of pure water is higher than the vapor pressure of the solution. This vapor pressure depression is one of the colligative properties of solutions. Because there is only one vapor pressure, it cannot simultaneously be in equilibrium with the water and with the solution. Thus water from the pure water beaker will continue to evaporate, because its vapor pressure is higher than that in the air, and water will continue to condense into the solution, because the partial pressure of water is higher than the vapor pressure of the solution. OSMOTIC PRESSURE IS A PROPERTY OF SOLUTIONS RELATED TO OTHER COLLIGATIVE PROPERTIES Thus osmotic pressure and vapor pressure depression are perfect predictors of each other because essentially they are the same phenomenon. or ½2:7:15 π 5 RTCs Osmotic pressure is closely related to some other properties of solutions, the colligative properties. These include the freezing point depression, the boiling point elevation, and the vapor pressure depression, all caused by dissolving solutes in a solution. The osmolarity is often determined from vapor pressure depression or freezing point depression, rather than from direct osmotic pressure measurements. The osmolarity is the concentration necessary to observe these phenomenon. To see the connection between osmotic pressure and vapor pressure depression, consider Figure 2.7.3. A solution placed in a sealed container with a source of pure water will gain water because its vapor pressure is lower than that of the water. This situation is formally equivalent to osmosis, where the semipermeable membrane is the intervening air between the two surfaces. THE OSMOTIC COEFFICIENT ϕ CORRECTS FOR THE ASSUMPTION OF DILUTE SOLUTION AND FOR NONIDEAL BEHAVIOR As noted above, the van’t Hoff equation makes two assumptions: the solution is dilute and it is ideal. The assumption of ideality enters when we equate the activity of water with its mole fraction. The assumption of dilute solutions allows us to identify ln(1 2 Xs) with 2 Xs. We can correct for both assumptions by identifying πobserved ϕ5 ½2:7:16 RTCs Here ϕ is the osmotic coefficient. The osmotic coefficient can be less than or greater than 1.0. 185 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION THE RATIONAL OSMOTIC COEFFICIENT CORRECTS FOR THE ASSUMPTION OF IDEALITY 3.5 Equation [2.7.9] gives the osmotic pressure in terms of the mole fraction of water: 2.5 ½2:7:17 π52 RT ln Xw Vw But the true equation for osmotic pressure is given by the manipulation of Eqn [2.7.6] as ½2:7:18 π52 3.0 π observed / π calculated 186 π52 RT RT p ln aw 5 2 ln 0 Vw Vw p Figure 2.7.4 shows the ratio of the observed osmotic pressure to that calculated by Eqn [2.7.15], Eqn [2.7.17], or Eqn [2.7.21]. These give an estimate of ϕ and g. The approximate values of ϕ at physiological concentrations for a variety of common solutes are given in Table 2.7.1. Units used in the calculation of osmotic pressure and appropriate values for R are given in Table 2.7.2. EQUIVALENCE OF OSMOTIC AND HYDROSTATIC PRESSURES As mentioned earlier, Pfeffer observed a linear relationship between the flow across the membrane and the pressure when water was on both sides of the membrane. This relationship can be described as ½2:7:22 g 1.5 1.0 0.5 0.0 0.0 0.5 g ln Xw 5 ln aw The value of g can be calculated from Eqns [2.7.17] and [2.7.19] as πobserved g5 ½2:7:20 RT 2 ln Xw Vw Thus the rational osmotic coefficient corrects for the discrepancy between the osmotic pressure and the osmotic pressure calculated from the mole fraction of water. This assumes ideality, in which the activity is equal to the mole fraction, but not dilution. Thus the calculations of osmotic pressure based on Eqn [2.7.17] are better than those calculated using van’t Hoff’s Law because calculations based on Eqn [2.7.17] are valid even for solutions that are not dilute. However, Eqn [2.7.17] still requires the assumption of ideal solution behavior, or that the activity of water is equal to its mole fraction. Eqn [2.7.18] gives the osmotic pressure without assuming either a dilute solution or ideality. It can be calculated from vapor pressure measurements as ½2:7:21 2.0 πobserved / (-RT/VW ln p/p0) RT ln aw Vw We introduce the rational osmotic coefficient, g, to make these two equations give the same result: ½2:7:19 φ QV 5 A LP ðPL 2 PR Þ 5 ALP ΔP JV 5 LP ðPL 2 PR Þ 5 LP ΔP 1.0 1.5 2.0 Sucrose concentration (M) 2.5 3.0 FIGURE 2.7.4 Osmotic coefficients as a function of sucrose concentration. The osmotic coefficient, ϕ, was calculated according to Eqn [2.7.16] by dividing the observed osmotic pressure by RTC (circles). The rational osmotic coefficient was obtained according to Eqn [2.7.20] by dividing the observed osmotic pressure by 2 RT/VW ln XW (squares). The correlation of vapor pressure to osmotic pressure was tested by dividing the observed osmotic pressure by 2 RT/VW ln p/p0 (triangles). This figure shows that the van’t Hoff equation is good for dilute solutions but fails at high solute concentrations, due to the failure of the assumptions that solutions are dilute and ideal. The osmotic coefficient corrects for these failures. The rational osmotic coefficient deviates significantly at higher sucrose solutions where solution behavior is further from ideal. The ratio of the observed osmotic pressure to that calculated from vapor pressure measurements is close to 1.0 over the entire concentration range. Data from Glasstone, S. Textbook of Physical Chemistry. Princeton: Van Nostrand, 1946. TABLE 2.7.1 Approximate Values of the Osmotic Coefficient for Common Solutes Under Physiological Conditions Solute Number of Particles Formed upon Solution Molecular Weight (g mol21) NaCl 2 58.4 0.93 KCl 2 74.6 0.92 Osmotic Coefficient (ϕ) CaCl2 3 111.0 0.85 Na2SO4 3 142.0 0.74 MgCl2 3 95.2 0.89 MgSO4 2 120.4 0.58 NaHCO3 2 84.0 0.96 Alanine 1 89.1 1.00 Mannitol 1 182.2 1.00 Glucose 1 180.2 1.01 Sucrose 1 342.3 1.02 where QV is the flow in units of volume per unit time and JV is the flux, or flow per unit area, in units of velocity. Here the positive flow is taken from left to right and pressure drives the flow. Thus if PL . PR, then Os mos is an d Os motic P res sur e TABLE 2.7.2 Units for the Calculation of Osmotic Pressure Pressure Units 1 atm Equivalent Gas Constant (R) Solute Osmolyte Concentration (ΣCs)a atm 1 0.082 L atm mol21 K21 mol L21 mm Hg 760 62.36 L mm Hg mol21 K21 mol L21 Pa 5 N m22 22 dyne cm 1.013 3 105 8.314 N m mol21 K21 1.013 3 10 8.314 3 10 dyne cm mol 6 7 mol m23 5 mol (1000 L)21 21 21 K mol cm23 a Osmolarity (osmol L21) is defined as the concentration of osmotically active particles, osmolytes, in mol L21. Therefore, the units osmoles and moles cancel in the calculation of osmotic pressure. Example 2.7.1 Calculate the Osmotic Pressure of 0.9% NaCl 0.9% NaCl means 0.9 g per 100 mL of solution. This is 9 g L21. We can convert this to molarity by dividing by the molecular weight, 58.4 g mol21: [NaCl] 5 9 g L21/58.4 g mol21 5 0.154 M. The osmotic pressure is calculated as π 5 RT ϕC 5 0.082 L atm mol21 K21 3 310 K 3 0.2866 mol L21 5 7.29 atm. This is an enormous pressure on the physiological scale. The effective osmolarity of this solution is 2 osmol mol21 3 0.154 M 3 0.93 5 0.2866 osmol L21. QV is positive and flow is to the right. If PL , PR, then (PL 2 PR) is negative, QV is negative and flow is to the left. Lp is a coefficient characteristic of the membrane, variously called the hydraulic conductivity, hydraulic permeability, or filtration coefficient. When Pfeffer added impermeant solutes to the inner chamber (left chamber), flow was observed into the solution in the absence of any macroscopic hydrostatic pressure differences. That is, when ΔP 5 0 there was a negative flow, and that flow was proportional to the osmotic pressure. Thus it appears that the solute caused a reduction in the pressure on the solution side because the flow is inward and additional pressure on the solution side, the osmotic pressure, is necessary to stop the osmotic flow. The osmotic pressure, which is characteristic of the solution, is equivalent to a reduction in the hydrostatic pressure. If solution is present on both sides of the membrane, we write ½2:7:23 QV 5 A LP ½ðPL 2 πL Þ 2 ðPR 2 πR Þ QV 5 A LP ½ðPL 2 PR Þ 2 ðπL 2 πR Þ QV 5 A LP ðΔP 2 ΔπÞ THE REFLECTION COEFFICIENT CORRECTS VAN’T HOFF’S EQUATION FOR PERMEABLE SOLUTES When membranes are partially permeable to the solute (leaky membranes), the measured osmotic pressure is less than that predicted by van’t Hoff Law. This has led to the definition of another membrane parameter, σ, the reflection coefficient, which is defined as πeff ½2:7:24 σ5 ϕRTC where πeff is the effective or observed osmotic pressure and ϕRTC is the theoretical osmotic pressure that would be observed if the membrane was perfectly semipermeable. The reflection coefficient is different for each pair of solute and membrane. Its range is 0 # σ # 1.0. Rewriting Eqn [2.7.24], we have ½2:7:25 πeff 5 σϕRTC 5 σπ It is important to note that the osmotic pressure is a characteristic of the solution, and it is defined as the pressure necessary to stop osmotic flow when a solution is separated by an ideal, perfectly semipermeable membrane from pure water. Recall that a semipermeable membrane is defined as one that has zero permeability to solute. Every aqueous solution has an osmotic pressure—it is a concentration measure like molarity is—and it can be measured by any of the colligative properties. When a solution is placed in contact with a membrane that is not semipermeable, its osmotic pressure will be reduced by the membrane’s reflection coefficient, which is characteristic not only of the membrane but also of the solute. As a consequence of this, flow across a real membrane that is not semipermeable will be altered. It is governed by the equation ½2:7:26 QV 5 A LP ½ðPL 2 ΣL σi πi Þ 2 ðPR 2 ΣR σi πi Þ QV 5 A LP ðΔP 2 ΣΔσi πi Þ where the summation sign means that all osmotically active solutes in the solution contribute to the total osmotic pressure. We now have three phenomenological coefficients that describe volume and solute flux across a membrane. These are summarized in Table 2.7.3. The question before us now is: what is the physical meaning of LP and σ? 187 188 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Example 2.7.2 Calculate the Net Driving Force and Flow Across a Dialysis Membrane. A semipermeable dialysis membrane has an area of 90.5 cm2 and Lp 5 6.46 3 1026 cm min21 mmHg21. Inside the dialysis membrane was a solution of 5% (w/v) sucrose at a pressure of 10 mmHg. Outside was a solution of 2% sucrose at a pressure of 50 mmHg. The entire apparatus was equilibrated to room temperature, 20 C. What is the net pressure across the membrane? What is the net flow across the membrane? Assume ϕ for sucrose is 1.0. The molecular weight of sucrose is 342 g mol21. First, we calculate the osmotic pressures of the solutions. 5% sucrose means 5 g of sucrose per 100 mL of solution or 50 g L21. We can convert this to molarity by dividing by the molecular weight, 342 g mol21: ½sucrosein 5 50 g L21 =342g mol21 5 0:146 M: The molar concentration of 2% sucrose is similarly calculated as ½sucroseout 5 20 g L21 =342g mol21 5 0:058 M The osmotic pressure of the two solutions is calculated as πout 5 ϕ RT 3 0:058 M 5 62:36 L mmHg mol21 K21 3 293 K 3 0:058 mol L21 5 1060 mmHg The net driving force is calculated as ðPin 2 Pout Þ 2 ðπin 2 πout Þ 5 ð10 mmHg 2 50 mmHgÞ 2 ð2668 mmHg 2 1060 mmHgÞ 5 2 1648 mmHg Positive pressure would force fluid outward; negative net pressure means flow is inward. Note that the systematic insertion of the pressure values is essential to obtaining the correct result. The inward flow is given by QV 5 A LpðΔP 2 ΔπÞ 5 90:5 cm2 3 6:46 3 1026 cm min21 mmHg21 3 2 1648 mmHg 5 2 0:963 cm3 min21 : πin 5 ϕ RT 3 0:146 M 5 62:36 L mmHg mol21 K21 3 293 K 3 0:146 mol L21 5 2668 mmHg Pore TABLE 2.7.3 Phenomenological Membrane Coefficients Coefficient Parameter Calculated as p Permeability ( Js/ΔC)Jv50 Lp Hydraulic conductivity ( Jv/ΔP)Δπ50 ; 2 ( Jv/σΔπ)ΔP50 σ Reflection coefficient 2( Jv/LpΔπ)ΔP50 δ n = N pores/unit area LP FOR A MICROPOROUS MEMBRANE DEPENDS ON THE MICROSCOPIC CHARACTERISTICS OF THE MEMBRANE Here we consider fluid flow across a microporous membrane such as that shown in Figure 2.7.5. We consider that the membrane itself is impermeant to water and solute, but both may go through tiny pores in the membrane. We consider three cases: 1. There is a pressure difference across the membrane but either there is water on both sides of the membrane or the solute particles are very small compared to the size of the pore. Pressuredriven flow will occur. 2. There is no pressure difference across the membrane that separates water from solution, but the solutes on one side of the membrane are FIGURE 2.7.5 Schematic drawing of the hypothetical microporous membrane. We imagine that this membrane is a thin, flat sheet that is pierced by right cylindrical pores of radius a and depth δ, which is equivalent to the thickness of the membrane. There are n 5 N/A pores per unit area of the membrane. The membrane separates two aqueous solutions with a solute concentration CL on the left and CR on the right. We further suppose that the left chamber is subject to the pressure PL and the right side to the pressure PR. too large to fit through the pores. This will produce an osmotic pressure difference and an osmotic flow. 3. There is no pressure difference across the membrane that separates water from solution, but the solutes are small enough to fit through the pore, but not as easily as water. This will produce a smaller osmotic pressure and a smaller osmotic flow than in case 2. Os mos is an d Os motic P res sur e CASE 1: THE SOLUTE IS VERY SMALL COMPARED TO THE PORE The volume flow through the membrane is the number of pores times the volume flow per pore: ½2:7:27 Qv 5 N qv where QV signifies the overall observed macroscopic flow, N is the number of pores, and qv is the flow through a single pore. If we assume laminar flow (see Chapter 1.2), then the flow through each pore will be given by Poiseuille’s Law, which gives the flow through a right cylindrical pipe of radius a and length δ as ½2:7:28 qv 5 πa4 ΔP 8ηδ where η is the viscosity of the fluid. The use of Poiseuille’s law in this situation requires the assumptions of laminar flow, Newtonian fluids, and a pore length that is long compared to the entrance length of the pore. The entrance length is the distance it takes for the fluid to establish a parabolic velocity profile within the pore. The observed macroscopic flux, Jv, is the flow per unit area of membrane per unit time. Substituting Poiseuille’s law into Eqn [2.7.27] gives the volume flux as ½2:7:29 Jv 5 Nqv nπa4 5 nqv 5 ΔP A 8ηδ where n 5 N/A is the density of pores in the membrane 2 the number of pores per unit area. Comparing this to Eqn [2.7.22], JV 5 LP ΔP, we can identify ½2:7:30 Lp 5 nπa4 8ηδ According to this equation, the observed macroscopic flux is linearly related to the pressure difference by a coefficient which includes the thickness of the membrane, δ, the density of pores in the membrane, n, the radius of the pores, a, and the viscosity of the fluid, η. Lp has units of volume per unit time per unit area per unit pressure. CASE 2: THE SOLUTE IS LARGER THAN THE PORE: THE MECHANISM OF OSMOSIS FOR MICROPOROUS MEMBRANES If the solute particles are larger than the pore, then they cannot get through the membrane and the membrane is perfectly semipermeable. Pfeffer has already experimentally determined what happens in this case, and his experimental results permit several conclusions. First, the solute in the water causes the flow because there is no flow when pure water is on both sides of the membrane. Second, the pressure at equilibrium or the flow is directly proportional to the concentration of the solute, if the solutions are sufficiently dilute. Third, the solute causes a reduction in the pressure of the solution because the flow is inward and additional pressure on the solution side is necessary to stop the osmotic flow. Fourth, because the observed osmotic pressure obeys van’t Hoff’s Law only when the membrane is impermeable to solute, the osmotic pressure and the osmotic flow result from the interaction of the solute with the membrane. If the membrane is highly permeable to solute also, no osmosis and no osmotic pressure are observed. We should look to the interaction of the membrane with the solute to explain the mechanism of osmosis. The derivation for one mechanism of osmotic pressure flow and pressure generation is given in Appendix 2.7.A1. Briefly, this derivation recognizes that the total pressure in the bulk solution results from contributions of both solute and solvent molecules. However, solute molecules cannot enter the pores and so they cannot contribute to the pressure within the pore. Thus there is a pressure deficit on the solution side immediately upon entering the pore from that side. Since pressure results from collisions of molecules, resulting in momentum change in the molecules, there is a momentum deficit within the pores due to restricted entry of solute particles into the pores. In Case 2, we have the situation where PL 5 PR, CL 5 CL, and CR 5 0 for a membrane which is impermeant to solute. From Eqn [2.7.A1.9], the pressure immediately within the pore, PL’, is given by ½2:7:31 P 0L 5 PR 2 RTCL assuming the validity of van’t Hoff’s law. A plug of water extending the length of the pore is subjected to a pressure difference, which is given by ½2:7:32 P0L 2 PR 5 2 RTCL because, in this case, PL 5 PR. The negative sign indicates the pressure is higher on the right than on the left. Thus fluid flow should occur from right to left according to Poiseuille’s law: ½2:7:33 qv 5 πa4 ð2πÞ 8ηδ where qv is the flow per pore. In this equation, we see the absurd situation in which one symbol is used to signify two entirely different quantities. The π in the fraction is the geometric ratio, while the π in parenthesis signifies the osmotic pressure. The negative sign indicates flow to the left. The overall volume flux for the membrane is then given as ½2:7:34 Jv 5 N qv nπa4 5 ð2πÞ A 8ηδ It is clear, then, that Jv in the presence of osmotic flow is given by ½2:7:35 Jv 5 2 L p π where Lp is identical to that described earlier for pressure-driven flow (see Eqn [2.7.30]). 189 190 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION PL > PR + RTCL PL = PR + RTCL Membrane CL CL PL Membrane CL RTCL PR RTCL RTCL Membrane PR PR Solute particles qv < 0 Js = 0 CR = 0 x=0 x=δ qv = 0 qv > 0 CR = 0 CR = 0 x=0 x=δ x=0 x=δ FIGURE 2.7.6 The direction of flow in the presence of hydrostatic and osmotic pressure differences across the membrane. Far left: The hydrostatic pressure is the same on both sides of the membrane but the left side contains a solution of concentration CL. In the pore, there is a pressure gradient and flow is toward the solution. Middle: The pressure on the left was increased by RTCL 5 π, the osmotic pressure. This is the equilibrium situation and no flow occurs. Right: Pressure on the left is more than RTCL greater than pressure on the right, and flow is to the right. It is instructive to consider the pressure profiles within the pore which are established due to the osmotic pressure. Poiseuille’s Law is given as πa4 ΔP ½2:7:36 Qv 5 8η ΔX Since @Qv/@x 5 0 at steady state, it follows that ΔP/Δx must be a constant in x and in t: the pressure gradient is linear in x. This allows us to draw the schematic diagrams of pressure and concentration gradients during osmotic flow as shown in Figure 2.7.6. CASE 3: THE REFLECTION COEFFICIENT RESULTS FROM PARTIALLY RESTRICTED ENTRY OF SOLUTES INTO THE PORES When the membrane is freely permeable to the solute (Case 1), there is no osmosis and no osmotic pressure and hydrostatic pressure drives fluid flow. When the membrane is completely impermeable to the solute (Case 2), and there is no hydrostatic pressure, the osmotic pressure is ideally given by van’t Hoff’s Law and drives fluid flow. Here we consider Case 3, a membrane which distinguishes between solute and solvent but which is not completely impermeable to solute. Here we consider pores that are large compared to the solvent particles but they partially exclude solute particles. The partial exclusion is due to the fact that the effective area of the pore is reduced compared to the area available to the solvent. When solute particles enter the pore, it reduces the effective osmotic pressure in direct proportion to the fraction of solute particles that can enter the pore. Consider that the pores have a radius a and that the solute particles are spherical with a radius as. If as , a, then at least some of the solute particles can get through the pore. Suppose that if a solute Solute particle Area available to solvent Membrane surface a a – as as Area available to solute FIGURE 2.7.7 Relative areas of the pore available to solvent water and solute particles. The figure depicts a single pore, looking down along its axis normal or perpendicular to the surface of the membrane. The outer circle is the dimension of the pore that is available to the solvent water. The inner circle represents the dimensions of the pore available to the solute, which is less than that available to solvent. particle hits the rim of the pore prior to entry, then it is reflected back into the bulk solution. The effective pore area will be reduced due to this reflection. The situation is depicted schematically in Figure 2.7.7, looking perpendicular to the membrane along the axis of the pore. The area of the pore which is accessible to solute is given by as 2 As 5 πða2as Þ2 5 πa2 12 ½2:7:37 a The relative area available to solute compared to that available to solvent is as 2 2 πa 12 As a 5 12 as 2 5 ½2:7:38 A πa2 a Os mos is an d Os motic P res sur e The fraction of solute particles which are reflected by the pore is approximated by the ratio of the area in light blue in Figure 2.7.7 to the total area. This is identified with σ, the reflection coefficient: A 2 As As 512 A A 2 as σ 5 1 2 12 a σ5 ½2:7:39 SOLUTIONS MAY BE HYPERTONIC OR HYPOTONIC If ξ 5 as /a, then σ 5 2ξ 2 ξ2 ½2:7:40 We may expect the ratio of the concentration in the pore to the concentration in the bulk solution to be the same as the ratio of the pore area available to solute to the area available to solvent, As/A. This is given by Eqn [2.7.39] to be ½2:7:41 C0L As 512σ 5 CL A In the absence of solvent drag (when Jv 5 0), the concentration profile may be drawn as shown in Figure 2.7.8. In this model, some of the solute molecules can enter the pore and therefore they contribute to the pressure in the pore. We expect the pressure deficit within the pore to be due only to those molecules which are reflected. Thus the osmotic pressure should be σRTΔC 5 σΔπ. The expression for laminar flow in the pore is ½2:7:42 Jv 5 physical meaning only in the range 0 # σ # 1.0. If As 5 A, then the pore is large enough so that the membrane does not discriminate between solute and solvent, and σ 5 0 according to Eqn [2.7.39]. If as $ a, then As 5 0 and σ 5 1.0. When cells are placed in contact with a solution, they may either swell or shrink as shown in Figures 2.7.9 and 2.7.10. These observations introduce the idea of tonicity, which is operationally defined. If we place a cell in a solution and the cell swells, the solution is called hypotonic. If we place a cell in a solution and the cell shrinks, we call that solution hypertonic. If the cell neither swells nor shrinks, the solution is isotonic. OSMOSIS, OSMOTIC PRESSURE, AND TONICITY ARE RELATED BUT DISTINCT CONCEPTS Osmotic pressure is a theoretical concept. It is the pressure necessary to stop osmotic flow if a solution is separated from pure water by a semipermeable membrane. A semipermeable membrane is defined as a membrane that allows the passage of some molecules but not others. It is freely permeable to water but impermeable 1.4 nπa4 ½ΔP 2 σΔπ 8ηδ Solutions that cause cells to swell are hypotonic 1.3 Jv 5 Lp ½ΔP 2 σΔπ Membrane V/V0 (relative volume) 1.2 In the above model, σ is viewed as being due to a hindered entry into the pore. There may be additional hindrance to solute flow through the pore due to interactions with the pore wall, but these are ignored here. According to this view, the reflection coefficient has Solutions that cause cells to shrink are hypertonic 1.1 1.0 0.9 CL Solutions that cause cells to neither shrink nor swell are isotonic σCL 0.8 (1 – σ) CL CR Hypotonic Isotonic 0.7 0.2 (1 – σ) CR x=0 x=δ FIGURE 2.7.8 Concentration profile within a pore that partially excludes solute particles. This profile pertains only when solvent drag is zero. Solvent drag washes the concentration gradient in the direction of volume flow and alters Js. Hypertonic 0.4 0.6 0.8 1.0 1.2 1.4 [NaCl] (%) 1.6 1.8 2.0 2.2 FIGURE 2.7.9 Swelling or shrinking of red blood cells in contact with different concentrations of NaCl. Whole blood was centrifuged to separate the red blood cells from the plasma. Then an aliquot of the plasma was removed and replaced with an equal volume of the indicated concentration of NaCl. When the plasma was replaced with NaCl solution less than about 0.9%, the cells’ volume increased relative to that in normal plasma. These solutions are described as being hypotonic. When the [NaCl] that replaced plasma was greater than 0.9%, the red blood cells shrank. These solutions are called hypertonic. Replacement with 0.9% NaCl caused the cells to neither swell nor shrink: this solution is isotonic. 191 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION to solute. All solutions have an osmotic pressure. In the approximation of the ideal, dilute solution, the magnitude of the osmotic pressure is given by van’t Hoff’s Equation (see Eqn 2.7.7), where C is the molar concentration of osmotically active solutes. This C is the osmolarity. Both osmolarity and osmotic pressure are properties of a solution, but tonicity is not. Tonicity refers to the direction of osmotic flow when a particular solution is placed in contact with a particular cell. Tonicity involves real membranes rather than an ideal, semipermeable membrane. The real membrane has a set of reflection coefficients, one for each solute on each side of the membrane. Whether shrinking or swelling occurs depends on the osmotic concentrations and compositions of the two solutions on the two sides of the cell membrane, and also on how these solutes interact with the real biological membrane. 1.4 Solutions that cause cells to swell are hypotonic 1.3 CELLS BEHAVE LIKE OSMOMETERS 1.2 V/V0 (relative volume) 192 1.1 1.0 Solutions that cause cells to shrink are hypertonic 0.9 Solutions that cause cells to neither shrink nor swell 0.8are isotonic 0.7 Hypotonic Hypertonic Isotonic 0.6 0 2 4 6 8 10 12 14 16 [Glucose] (%) FIGURE 2.7.10 Swelling or shrinking of red blood cells in contact with different concentrations of glucose. Whole blood was centrifuged to separate the red blood cells from the plasma. Then an aliquot of the plasma was removed and replaced with an equal volume of the indicated concentration of glucose. When the plasma was replaced with solutions of glucose less than about 5%, the cells’ volume increased relative to that in normal plasma. These solutions are described as being hypotonic. When the [glucose] that replaced plasma was greater than 5%, the red blood cells shrank. These solutions are called hypertonic. Replacement with 5% glucose caused the cells to neither swell nor shrink: this solution is isotonic. Consider a cell with a volume V0 and a total concentration of osmotically active solutes, ΣCi, resting in an isotonic medium. Now suppose that we rapidly replace the medium with another with a different osmotic pressure which we will symbolize as π. If the new medium is hypotonic, then by definition the cell will swell to a new volume, at which point the cell’s cytoplasmic tonicity will match the medium’s. Similarly, if the new medium is hypertonic, then the cell will shrink to a new volume that will match the tonicity of the medium. We assume here that the volume of the medium is so large that water movement into the cell or out of the cell does not appreciably affect the osmolarity of the medium. We further assume that water movement is fast compared to the movement of solutes or ions across the membrane, and that we can measure the cell’s volume after water movement has occurred but before solute movement. What is the relationship between the medium osmotic pressure and the final cell volume, Vc? In this situation, the total amount of osmotically active solutes in the cell is constant—the intracellular solutes do not move. The total intracellular osmolytes are equal to V0ΣCi, where V0 is the volume under isotonic conditions and Ci is the concentration of solute i in the cytoplasm in the isotonic condition. Now the total amount of osmotically active solutes at any time is Example 2.7.3 Isosmotic Solutions May Not Be Isotonic A. Calculate the osmolarity and osmotic pressure of isotonic saline (NaCl): Isotonic saline is 0.9% NaCl. This is 0.9 g NaCl per dL or 100 mL of solution 5 0.9/0.1 L or 9.0 g L21. The formula weight for NaCl is 58.44 g mol21. The molarity of NaCl is 9 g L21/58.44 g mol21 5 0.154 M. Since NaCl dissociates into two particles per mole, the osmolarity is 2 3 0.154 M 5 0.308 OsM. The osmotic coefficient of NaCl is ϕNaCl 5 0.93, so that the measured osmolarity of this solution would be ϕ 3 C 5 0.93 3 0.308 OsM 5 0.286 OsM. The osmotic pressure at 37 C is 7.27 atm. B. Calculate the osmolarity and osmotic pressure of 1.8% urea in water: The formula weight for urea is 60.0 g mol21 and ϕurea 5 0.95. Its osmolarity is ϕC 5 0.95 3 18 g L21/60 g mol21 5 0.285 OsM. Its osmotic pressure is 7.24 atm. C. What happens when red blood cells are placed in 0.9% NaCl? 1.8% urea? Red blood cells placed in contact with 0.9% NaCl neither shrink nor swell. The solution is isotonic. When placed in 1.8% urea, the cells swell so much that the cells lyse or break open. Although 0.9% NaCl and 1.8% urea are isoosmotic with red blood cell contents, 0.9% NaCl is isotonic and 1.8% urea is not isotonic. Os mos is an d Os motic P res sur e distributed in the volume of the cell, Vc. So the osmolarity at any time is P V0 Ci ½2:7:43 Vc When the osmotic pressure of the medium equilibrates with that of the cell, with the assumption of no solute movement (which is equivalent to assuming σi 5 1.0), we get P V0 Ci ½2:7:44 π 5 RT Vc This can be rewritten as π5 ½2:7:45 V0 Σ RTCi V0 5 πisotonic Vc Vc which we can rearrange to obtain πisotonic Vc 5 V0 ½2:7:46 π According to Eqn [2.7.46], a cell acting as a perfect osmometer would show a cell volume that was inversely proportional to the osmolarity of the external medium, with an intercept of zero. Typically the cell’s volume under isotonic conditions, V0, is the control for their volume under nonisotonic conditions. Thus we rewrite Eqn [2.7.46] as Vc πisotonic 5 V0 π ½2:7:47 Figure 2.7.11 shows the plot of the volume of cardiac cells exposed to various osmolar solutions, with volume normalized to the volume under isotonic conditions. According to Eqn [2.7.47], we expect the data to be linear with an intercept of zero. Actual data from these cardiomyocytes show that the response of relative volume (Vc /V0) is linear within a considerable range, 1.6 but the curve does not extrapolate to zero volume. These real cells are not perfect osmometers. The response of real cells is described by the empirical equation Vc Vb πisotonic Vb 1 ½2:7:48 5 12 V0 V0 π V0 The intercept of this line on the volume axis is Vb/V0, which in the case of cardiomyocytes shown in Figure 2.7.11 is 0.34. This is interpreted to mean that there is a fraction of the cell’s volume that is osmotically inactive. This is partly to be expected. Not all of the volume of the cell is water, and it is the volume of water that dissolves osmotically active solutes and is responsive to changes in the medium osmolarity. Thus the sum of all of the volumes of large particles such as DNA, RNA ribosomes, and membranes contributes to an osmotically inactive volume. Although this is certainly part of the explanation for the osmotically inactive volume, the issue is by no means settled. Other components of the inactive volume might include the volume of a variety of macromolecular assemblies such as the cytoskeleton. CELLS ACTIVELY REGULATE THEIR VOLUME THROUGH RVDs AND RVIs Although the presence of a hypotonic or hypertonic solution initiates swelling or shrinking, respectively, often the volume change is not maintained. A cell that initially swells when placed in a hypotonic medium may eventually lose some of its acquired volume: it undergoes a regulatory volume decrease or RVD. The swelling of the cell activates compensatory mechanisms that cause transport of osmotically active solutes (osmolytes) out of the cell. Similarly, cell shrinking can activate an influx of osmolytes in some cells leading to a compensatory swelling that is called a regulatory volume increase or RVI. These RVDs and RVIs are accomplished by altering the cell’s contents of osmolytes. Relative volume (V/V0) 1.4 1.2 SUMMARY 1.0 Osmosis refers to the movement of fluid across a membrane in response to different concentrations of solutes on the two sides of the membrane. The movement of fluid is toward the more concentrated solution. Osmotic pressure is defined as the pressure that must be applied to the solution side to stop fluid movement when a semipermeable membrane separates a solution from pure water. Here, the semipermeable membrane is permeable to water but not to solute. The osmotic pressure for dilute ideal solutions obeys van’t Hoff’s Law: 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 πisotonic/π FIGURE 2.7.11 Relative volume of isolated cardiac cells exposed to test solutions of different osmolarities. Source: Data from Drewnowska and Baumgarten, American Journal of Physiology 260:C122131, 1991. π 5 RTΣ Cs which can be derived on thermodynamic grounds. Because the solutions are not ideal, the equation is 193 194 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION refined by including an osmotic coefficient, ϕs, characteristic of each solute: πobserved 5 RTΣ ϕs Cs Defined in this way, the osmotic pressure is a pressure deficit caused by dissolving solutes. However, membranes that are somewhat permeable to the solute develop a transient osmotic pressure that is less than that predicted by van’t Hoff’s Law. Thus the actual pressure developed across a membrane that separates a solution from pure water depends on the interaction of the solute with the membrane. The correction for partially permeable membranes requires the reflection coefficient, σ: πobserved 5 RTΣ σs ϕs Cs The magnitude of the pressure tells us nothing of the flow. Osmotic pressure and hydrostatic pressure add to drive fluid flow across the membrane, with a proportionality constant, Lp. The phenomenological equation describing fluid flow is Qv 5 A Lp ðΔP 2 σΔπÞ Osmosis and osmotic pressure is a thermodynamic concept which exists independently of mechanism. In microporous membranes, osmosis is caused by a momentum deficit within the pores due to the reflection of solute molecules by the membrane. This reduces the pressure on the solution side of the pore by π for a semipermeable membrane. Thus there are three characteristic parameters for describing passive material transfer across membranes: the permeability, p, the hydraulic conductivity, Lp, and the reflection coefficient, σ. Osmolarity is a kind of concentration measure, distinct from molarity. It is related to other colligative properties of solutions including freezing point depression, vapor pressure depression, and boiling point elevation. Tonicity is a related concept but involves a real, biological membrane that may not be semipermeable. Tonicity makes reference to a particular cell and its membrane. Solutions may be isoosmotic but not isotonic. 3. What is the reflection coefficient? How does it relate to permeability? 4. The equations in this chapter were derived for a microporous membrane. Would they still hold for a lipid dissolution model of a membrane? 5. How does tonicity differ from osmolarity? Define hypotonic, hypertonic, and isotonic. 6. What is the relationship between volume and osmotic pressure in a perfect osmometer? 7. What is meant by “osmotically inactive” volume? 8. How do cells regulate their volume? APPENDIX 2.7.A1 MECHANISM OF OSMOSIS: FILTRATION VERSUS DIFFUSION DOWN A CONCENTRATION GRADIENT PHYSICAL ORIGIN OF THE OSMOTIC PRESSURE ACROSS A MICROPOROUS MEMBRANE Consider what happens in the vicinity of a single pore, as shown in Figure 2.7.A1.1. Since the pore excludes solute particles, the average concentration must decrease from CL in the bulk solution to zero in moving along the axis of the pore. Thus there is a concentration gradient near the opening of the pore. By Fick’s First Law of Diffusion, we should expect a net diffusion of solute particles into the pore. However, the particles hit the rim of the pore and are reflected back into the bulk solution. On the average, these particles experience a force in the 2x direction. The magnitude of this average force can be obtained from the generalized Fick’s equation (see Eqn [1.6.42]): ½2:7:A1:1 js 5 2 D CL @CðxÞ D 1 F CðxÞ @x RT Membrane Cells respond to swelling or shrinking according to the empirical relation: Pore V=V0 5 ð1 2 Vb =V0 Þ πisotonic =π 1 Vb =V0 where V0 is the volume of the cell under isotonic conditions. Vb is interpreted as the osmotically inactive volume. If Vb 5 0, the cell would behave like an ideal osmometer. Solute particles REVIEW QUESTIONS 1. In which direction does osmotic flow occur? Why? 2. What equation gives the magnitude of the osmotic pressure? What are the limitations of this equation? How would you correct for its errors? CR = 0 x=0 x=δ FIGURE 2.7.A1.1 Concentration gradient for solute particles near a pore. In this case, the solute particles are large compared to the pore and cannot penetrate into the membrane. Os mos is an d Os motic P res sur e We write R in place of k and F in place of f in Eqn [2.7.A1.1] because we are speaking of the force per mole rather than the force per molecule, as was done in Eqn [1.6.42]. Since js 5 0 for an impermeant membrane, Eqn [2.7.A1.1] becomes ½2:7:A1:2 FCðxÞ 5 RT Here we consider the forces acting on an element of fluid with an area A from a point x well within the bulk solution to a point x 1 Δx within the pore near its surface. We consider that the volume element is in mechanical equilibrium. The forces acting on this volume are the forces acting on the solute particles and the forces acting on the surfaces in contact with adjacent fluid. The sum of these forces must be zero to meet the condition of mechanical equilibrium. We identify these forces with the “body” forces, FB, and “contact” forces, FC. Thus we write: 0 5 FB 1 FC for the condition of mechanical equilibrium. The net contact force is the balance of the pressure acting on the surface area to the right and left of the volume element: ½2:7:A1:4 FC 5 A PðxÞ 2 A Pðx 1 ΔxÞ Membrane Solute particle ½2:7:A1:5 FB 5 dV = Adx F CðxÞdV x FB 5 ART ½2:7:A1:6 P(x + Δx) ð x1Δx x @CðxÞ dx @x 5 ART½Cðx 1 ΔxÞ 2 CðxÞ Since C(x 1 Δx) 5 0 because solute particles are not in the pore, Eqn [2.7.A1.6] becomes ½2:7:A1:7 FB 5 2 ART CðxÞ The negative sign in Eqn [2.7.A1.7] indicates that FB is directed to the left. Inserting Eqns [2.7.A1.4] and [2.7.A1.7] into Eqn [2.7.A1.3], we obtain ½2:7:A1:8 A PðxÞ 2 A Pðx 1 ΔxÞ 5 ART CðxÞ or ½2:7:A1:9 PðxÞ 2 Pðx 1 ΔxÞ 5 RTCL This equation says that the net pressure experienced by the volume of fluid immediately adjacent to the pore decreases as one moves from the left compartment into the pore, and that the drop in pressure is RTCL. This analysis suggests that the osmotic pressure develops as a consequence of the interactions between the solute particles and the membrane. The solute particles contribute momentum in the solution. When the particles collide with the membrane, that momentum is transferred to the membrane and not to the fluid within the pore. As a consequence, there is a momentum deficit within the pore (compared to the bulk solution). Since the pressure is the average momentum change per unit area experienced by particles colliding within the fluid, this momentum deficit shows up as a pressure deficit. The resulting pressure difference produces a flow that is given as Pore P(x) ð x1Δx Inserting dV 5 Adx and FC(x) 5 RT @C(x)/@x from Eqn [2.7.A1.2], we obtain @CðxÞ @x The force F is an external force that acts on the solute particles. The total force acting on the solute particles is just Fn(x), where n is the number of moles of particles. Dividing by the volume, V, we have Fn(x)/V 5 FC(x), which is the average force per unit volume. This is the force that the membrane is exerting on the solute bodies, per unit volume, located in the volume of fluid directly in front of the pore. The consequence of this is that there will be a pressure drop at the pore entrance. To see this, we will analyze the forces acting on an element of fluid located directly in front of a pore, as shown in Figure 2.7.A1.2. ½2:7:A1:3 The total body force is the body force per unit volume integrated over the volume. ½2:7:A1:9 Jv 5 nπa4 Δπ 8ηδ where we identify x x + Δx ½2:7:A1:10 x=0 x=δ FIGURE 2.7.A1.2 Forces acting on the plug of fluid immediately in front of a pore in a porous membrane. Lp 5 nπa4 8ηδ DIFFUSIONAL PERMEABILITY OF MICROPOROUS MEMBRANES In the absence of a hydrostatic or osmotic pressure gradient, water will diffuse across a microporous membrane. If we assume the membrane is impermeable 195 196 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION except at the pores, the diffusion through the pores will obey Fick’s Laws of diffusion: jw 5 2 D ½2:7:A1:11 @C @x @C @C2 5D 2 @t @x where jW is the flux through a single pore. At steady state, the flux does not change with time, or with distance, and so the gradient is linear in x and we can write ½2:7:A1:12 πa2 Dw ΔCw jw 5 δ where jw is the flux of water through the pore, a is the radius of the pore, DW is the diffusion coefficient of water in the pore, δ is the length of the pore (equal to the thickness of the membrane), and ΔCW is the concentration difference of water. The overall macroscopic flux for a microporous membrane is given as ½2:7:A1:13 Jw 5 n πa2 Dw ΔCw δ where n is the density of the pores, n 5 N/A, the number of pores per unit area of the membrane. This is analogous to the passive solute flux discussed in Chapter 2.5. We can similarly define a diffusional permeability for water from the relation: ½2:7:A1:14 Jw 5 Pd ΔCw and we can identify a diffusional permeability coefficient for water from Eqns [2.7.A1.13] and [2.7.A1.14] as ½2:7:A1:15 Pd 5 n πa2 Dw δ From the thermodynamic derivation of van’t Hoff’s Law, we write for the chemical potential of water ½2:7:A1:19 μw 5 μ0w 1 RT ln aw 1 V w P insertion of this into Eqn [2.7.A1.18] and at steady-state flow, we obtain for the macroscopic membrane ½2:7:A1:20 Jw 5 nπ2 Dw Cw V w ΔP RTδ Vw is the partial molar volume (in volume per mole, L mol21), and CW is the water concentration, in moles L21, and these cancel: VWCW 5 1.0. The flux given here is in mole per unit area per unit time. To convert to JV we multiply by VW: ½2:7:A1:21 JV 5 V w nπa2 Dw ΔP RTδ This describes pressure-driven flow across the membrane. We recognize that the same terms for Pd are present in this equation. We recognize Eqn [2.7.A1.21] as JV 5 LP ΔP and identify LP as ½2:7:A1:22 LP 5 V w nπa2 Dw V w Pf 5 RTδ RT where Pf is the filtration permeability for water. It has the same expression for Pd given in Eqn [2.7.A1.15] but we obtain it experimentally from LP as ½2:7:A1:23 Pf 5 LP RT n πa2 Dw 5 δ Vw Thus comparing Eqn [2.7.A1.15] to Eqn [2.7.A1.23], Pd and Pf should be the same if the mechanism of osmosis is by diffusion. FILTRATION PERMEABILITY IN THE PRESENCE OF A PRESSURE DIFFERENCE IN A MICROPOROUS MEMBRANE PRESSURE- AND OSMOSIS-DRIVEN FLOW ACROSS A LIPID BILAYER BY DISSOLUTIONDIFFUSION In Chapter 1.6, we found that diffusional flux can be altered by the application of an additional force. In particular, a flux of solute obeyed the relation: Here we consider a membrane that presents all of its area to the solution phase, and water crosses by dissolving in the lipid of the membrane on one side of the membrane, diffusing across the membrane essentially like a vapor, and then coming out of lipid solution back into the aqueous phase on the other side. We imagine that dissolution is rapid (the solution phase and membrane phase are in equilibrium) and that diffusion is comparatively slow. Equilibrium of water in the solution phase with water in the membrane phase is described by equating the chemical potential of water in the two phases: ½2:7:A1:16 Js 5 2 D CF RT where JS is the flux of solute, D is its diffusion coefficient, R is the gas constant, T is the temperature (K), C is the concentration, and F is the total force on the particle per mole. We went on to define the electrochemical potential so that ½2:7:A1:17 F52 dμ dx These equations have general validity and are also true for water. We can write for the flux in the pore that ½2:7:A1:18 Dw dμw jw 5 C RT dx μ0wðsolutionÞ 1 RT ln XwðsolutionÞ 1 V wðsolutionÞ P 5 μ0wðmembranceÞ 1 RT ln XwðmembraneÞ 1 V wðmembraneÞ P ½2:7:A1:24 where Xw(solution) and Xw(membrane) are the mole fractions of water in the solution in equilibrium with the Os mos is an d Os motic P res sur e membrane and in the membrane phase, respectively. The partition coefficient is defined as ½2:7:A1:25 Kw 5 XwðmembraneÞ XwðsolutionÞ ½2:7:A1:23 Consider the case where only osmotic pressure drives water flow and the hydrostatic pressure difference across the membrane is zero. Since water generally partitions poorly into hydrocarbon solvents, we may assume that the mole fraction of water in the membrane phase is low. Thus the water concentration is dilute, and we may replace the mole fraction of water with its concentration: ½2:7:A1:26 CWðmembraneÞ XwðmembraneÞ V lipid If equilibrium is rapid, we can combine Eqn [2.7.A1.25] and Eqn [2.7.A1.26] to get ½2:7:A1:27 CwðmembraneÞ Kw XwðsolutionÞ V lipid For dilute solutions, XW(solution) can be replaced by 1 2 V w CS where CS is the solute concentration: ½2:7:A1:28 CwðmembraneÞ Kw 1 2 V w CS V lipid The concentration of water immediately inside the lefthand side of the membrane is given by Eqn [2.7.A1.28] where CS is the concentration of solute in the left compartment. A similar expression pertains to the water concentration immediately inside on the right side. The net diffusive flux of water across the membrane is given as m ½2:7:A1:29 Jw 5 2 D w Cw;L 2 Cw;R 02δ where Dwm is the diffusion coefficient of water in the membrane phase. Substituting in for the concentrations from Eqn [2.7.A1.28], we get Jw 5 2 Dm K V ðCS;L 2 CS;R Þ w w w V lipid δ 52 In Eqn [2.7.A1.23], we described Pf, the filtration water permeability, as Dm K V ΔCS w w w V lipid δ RT Vw P f 5 LP Applying this result to Eqn [2.7.A1.32], we obtain ½2:7:A1:33 Dm K V w w w Pf 5 V lipid δ This equation was derived for the situation in which there was an osmotic gradient (ΔCS . 0) in the absence of a hydrostatic pressure gradient (ΔP 5 0). The expression for the situation where ΔCS 5 0 and ΔP . 0 can be obtained by returning to Eqn [2.7.24] and setting the mole fractions of water to 1.0 while the pressures differ. The result is that exactly the same LP is derived for pressure-driven flow as for osmotic flow when the mechanism is by rapid dissolution of water followed by slow diffusion of water through the membrane phase. DIFFUSIONAL PERMEABILITY BY THE DISSOLUTIONDIFFUSION MECHANISM: PD The permeability of lipid membranes to a diffusional water flux is expressed by the equation: ½2:7:A1:34 Jw 5 Pd ΔCw where Pd is the diffusional permeability and ΔCw is the difference in water concentration across the membrane. The overall permeation of water takes three steps: dissolution into the membrane phase at the left interface, diffusion across the membrane phase, and reversal of the dissolution at the right interface. If we assume, as we did in the derivation of Pf, that the rate-limiting step is diffusion through the membrane phase, then we may rewrite Eqn [2.7.A1.34] as ½2:7:A1:35 ΔCwðmembraneÞ δ m Jw 5 Dw ½2:7:A1:30 Substituting in for ΔCw(membrane) from Eqn [2.7.A1.27], this is The flux given here is in mole per unit area per unit time. To convert to JV, we multiply by Vw: ½2:7:A1:36 2 JV 5 V w Jw 5 2 Dm K V ΔCS w w w V lipid δ 2 52 Dm K V w w w V lipid RT δ RTΔCS The last term on the far right is the osmotic pressure difference, Δπ. Equation [2.7.A1.31] relates the volume flux to the osmotic pressure when the mechanism of water flow is dissolution and diffusion. We recognize it as a form of the phenomenological equation, JV 5 2LPΔπ and therefore we identify LP as 2 ½2:7:A1:32 LP 5 Dm K V w w w V lipid RT δ Kw ΔXwðsolutionÞ V lipid δ since ΔXw 5 Vw ΔCw, this becomes ½2:7:A1:37 ½2:7:A1:31 m Jw 5 Dw m Jw 5 Dw Kw V w ΔCw V lipid δ and we can identify Pd by comparing Eqns [2.7.A1.37] and [2.7.A1.34] as ½2:7:A1:38 Pd 5 Dm K V w w w V lipid δ Comparing the results for Pf (see Eqn [2.7.A1.33]) and for Pd (see Eqn [2.7.A1.38]), the dissolutiondiffusion mechanism of water transport indicates that Pf/Pd 5 1.0. 197 198 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION If we measure the diffusional permeability of a lipid membrane and LP, and calculate Pf from the LP according to Eqn [2.7.A1.23], we should expect them to be equal provided that the mechanism is dissolution and diffusion. Our earlier results with microporous membranes also concluded that Pf/Pd 5 1.0 if the mechanism of transport was by diffusion. EXPERIMENTS CONFIRM PF/PD 5 1.0 FOR LIPID BILAYERS BUT PF/PD . . 1.0 FOR MOST BIOLOGICAL MEMBRANES The measurement of Pf (from LP) and Pd in planar lipid bilayers confirm that they are the same, indicating that water transfer across pure lipid bilayer membranes is by diffusion. However, in most biological membranes, Pf/Pd is much greater than 1.0, suggesting that the mechanism is not diffusional, but hydrodynamic. The erythrocyte membrane was used in many of these studies, and it was shown that compounds that interact with protein sulfhydryl groups markedly reduced Pf while leaving Pd relatively unchanged, and causing the ratio of Pf/Pd to become 1.0. These early studies pointed to the existence of proteinaceous pores in the erythrocyte membrane, and subsequently these were identified as AQP1, the first in the family of aquaporins. Peter Agre earned the Nobel Prize in 2003 for his discovery of the aquaporins. AQUAPORINS ACCOUNT FOR MOST WATER TRANSPORT IN CELLS Aquaporins are a family of proteins, all about 28 kDa, that are found in a variety of tissues that have high water transport rates such as the intestines and kidneys, salivary glands, pancreas, and more. There are 13 human varieties, labeled AQP0, AQP112. The proteins all have six transmembrane domains and associate to form a functional tetramer, although each part has its own water channel. Although water can permeate through lipid membranes, this pathway is much slower than the AQP-mediated pathway. Thus real biological membranes are a mosaic of lipid pathways for diffusional transport of water and pores for pressure-driven transport of water. Problem Set Membrane Transport 1. The GLUT-1 transporter has a Km for glucose of 1.5 mM. The normal, resting plasma glucose concentration is about 90 mg%. This is a clinical unit that is not part of the ISI but you have to get familiar with it anyway. X mg% means that the solute has that many mg in a deciliter of plasma. One deciliter is 0.1 L 5 100 mL. The molecular weight of glucose is 180 g mol21. From this information, calculate the rate of glucose transport by GLUT-1, as a percent of its maximum, when exposed to normal plasma. Assume saturation kinetics. 2. When CL was 2.3 3 1026 M and CR was zero, the flux across a microporous membrane was found to be 0.234 pmol cm22 s21. The free diffusion coefficient of the material being measured was 0.8 3 1025 cm2 s21. A. What is the permeability of the membrane to the material? B. If the thickness of the membrane is 10 3 1026 m, what is the equivalent relative area available for diffusion of the material? 3. Two membranes, A and B, have permeabilities PA and PB for a given solute. These two membranes are joined together to form a single composite, two-layered membrane, as shown in Figure 2.PS2.1. Examples are the successive filtration/diffusion barriers in the kidney glomerulus, successive permeability barriers in the lung, and successive permeability barriers of unstirred layers adjacent to intestinal epithelial membranes. A B CL CR FIGURE 2.PS2.1 Composite membrane formed from the sandwich of two membranes, A and B, with different characteristics. © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00018-5 2.2 A. How would you define the overall permeability P of the two-layered membrane? Hint: Think about how you would define permeability for any membrane. Think about what you need to know to calculate P. Use CL for the left concentration and CR for the concentration on the right. B. Why is the steady-state solute flux through the composite membrane the same through the membrane A layer and the membrane B layer? Hint: Think of the continuity equation and what it means. C. What is the concentration profile through the composite membrane? Hint: Calculate the concentration at the interface of membranes A and B; call it Cm, in terms of CL and CR. Hint: Equate the flux through the two membranes. D. Find an expression for P in terms of PA and PB. E. Do the permeabilities act like inverse resistances in a series arrangement? 4. When CL was 2.3 3 1026 M and CR was zero, the flux across a microporous membrane was found to be 0.234 pmol cm22 s21. This flux was determined with vigorous stirring, which virtually eliminated any unstirred layers adjacent to the membrane. When the stirrer was turned off, the flux decreased to 0.157 pmol cm22 s21. The free diffusion coefficient of the material being measured was 0.8 3 1025 cm2 s21. A. What is the permeability of the membrane plus unstirred layer? B. What is the permeability of the membrane alone? C. Using the information in Problem #3, what is the thickness of the unstirred layer? (Assume that the diffusion coefficient in the unstirred layer is equal to the free diffusion coefficient.) 5. The surface area of the lungs is about 75 m2, and the thickness of the alveolar diffusion layer is about 0.5 μm. The PO2 in alveolar air is 100 mmHg, while the PO2 of venous blood is 40 mmHg. The diffusion coefficient of O2 in water is about 1.5 3 1025 cm2 s21. The solubility of O2 is given by Henry’s Law as [O2] 5 0.024 PO2 ; here [O2] is expressed in mL O2 at STPD (standard temperature and pressure, dry: 0 C and 1 atm pressure) per mL of water and PO2 is in 199 200 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION atmospheres. Only the dissolved O2 diffuses. What is the initial rate of O2 diffusion from the aggregate alveoli to the blood? (Initial rate means to pretend that the venous PO2 is clamped at 40 mmHg.) Give the answer in mL O2 per minute and in mol O2 per minute using the Ideal Gas Law. Assume 37 C, R 5 0.082 L atm mol21 K21. Remember that 1 atm 5 760 mmHg. 6. Heart cells contain a NaCa exchanger with a stoichiometry of 3Na:1Ca. The following questions pertain to this transporter. A. The free energy of transport of Ca21 across the sarcolemma of the heart cell can be calculated from the following conditions during the rest phase of the heart beat: 10. ½Ca21 o 5 1:2 3 1023 M ½Ca21 i 5 0:1 3 1028 M Em 5 20:085 V Calculate the free energy for the reaction Caout-Cain. Recall that R58.314 J mol21 K21, T5310 K, `59.6493104 C mol21. Remember that Ca21 has two electrical charges per atom. B. Calculate the free energy for the reaction Nain-Naout for the following conditions during the rest phase of the heart: 11. ½Na1 o 5 145 3 1023 M ½Na1 i 5 14 3 1023 M Em 5 20:085 V C. Which way does the NaCa exchange proceed at rest? 7. Ischemia refers to the condition of no blood flow. When the artery perfusing an area of tissue is blocked, oxygen can no longer be delivered to support energy metabolism. Under these conditions, the ATP concentration falls and ADP and Pi concentrations rise. What will this do to the free energy of ATP hydrolysis (also called the affinity of ATP hydrolysis)? What do you think will be the consequence of ischemia on the effectiveness of the ion pumps? (Kammermeir, Schmidt, and Jungling, Free energy change of ATP hydrolysis: a causal factor of early hypoxic failure of the myocardium? J. Mol. Cell. Cardiol. 14:267277, 1982). 8. The membrane of the sarcoplasmic reticulum (an internal membrane in muscle cells) has multiple ion channels, so the membrane potential across this membrane is believed to be zero. This membrane has a Ca-ATPase pump that links two Ca21 atoms to the hydrolysis of ATP. Using a free energy of ATP hydrolysis of 257.1 kJ mol21, what is the thermodynamic limit of Ca21 accumulation if the free [Ca21] on the cytosolic face is 1 3 1027 M? (Hint: The thermodynamic limit is when the free energy change for the transport reaction is zero.) 9. Each of the reactions shown in Eqn [2.5.15] has a forward and reverse rate constant. Show that 12. 13. for a passive transport process the product of all the forward rate constants is equal to the product of all the reverse rate constants. (Hint: Use a principle called detailed balance, which states that at equilibrium all steps in a reaction sequence must also be at equilibrium.) Consider that a membrane separates two compartments, each containing a solution of 10 mL. The left compartment has an initial concentration CL. The membrane has permeability p to the solute and area A. A. Derive an expression for CL and CR as a function of time. B. Suppose that you obtained experimental values for CL(t) or CR(t). What plot of the data could you make to determine p? C. What would happen if we doubled both the volume and the surface area to the time course of equilibration of CL and CR. The surface area of the lungs is about 75 m2, and the thickness of the alveolar diffusion layer is about 0.5 μm. The PO2 in alveolar air is 100 mmHg, while the PO2 of venous blood is 40 mmHg. The diffusion coefficient of O2 in water is about 1.5 3 1025 cm2 s21. The solubility of O2 is given by Henry’s Law as [O2] 5 0.024 PO2 ; here [O2] is expressed in mL O2 per mL of water and PO2 is in atmospheres. Only the dissolved O2 diffuses. The volume of blood in the lung is 70 mL, and the volume of air at the end of normal inspiration volume is about 2.8 L (at body temperature of 37 C). A. Derive an equation for the time course of oxygen equilibration between blood and air in the lungs. B. Using the values given here, estimate the half-time of equilibration. C. Most of the oxygen in the blood is not free but is bound to hemoglobin within the red blood cells. Do you think this would accelerate or decelerate the rate of equilibration of blood and air oxygen pressures? Vesicles of the sarcoplasmic reticulum have embedded in their membrane an active CaATPase pump. Several different isoforms of this primary active pump are expressed in different tissues. When exposed to ATP, Mg, and Ca, Ca21 ions are accumulated and eventually reach a steady-state uptake. If the pump is quickly quenched by adding extravesicular EGTA, which complexes activator Ca21 and thereby stops the pump, the accumulated Ca21 will leak back out. Monitoring the intravesicular Ca21 with time allows one to estimate the permeability of the vesicles. Derive an expression for the amount of Ca21 remaining in the vesicle as a function of time and suggest a plot to determine the permeability. What other information might you need to know to determine the permeability? The osmotic coefficient for CaCl2 under physiological conditions is 0.85. Calculate the osmotic pressure of a solution of 10 mM CaCl2 at 37 C. Problem Set Recall here that R 5 0.082 L atm mol21 K21. Give the answer in both atm and mmHg (1 atm 5 760 mmHg). 14. The kidney filters plasma to produce an ultrafiltrate, which is the first step in the formation of urine. This filtrate is called an ultrafiltrate because the kidney can retain even small particles like plasma proteins. The force behind ultrafiltration is the blood pressure. The kidney ultrafiltration occurs at a structure called the glomerulus, which is a group of small blood vessels (capillaries) that are closely joined to another structure, Bowman’s capsule, that forms a double-walled cup for the collection of the ultrafiltrate. The filtration barrier is a combination of the capillary walls and structure in Bowman’s capsule. A. Calculate the filtration coefficient (Lp) for the basement membrane of kidney glomeruli using the following approximations: 1. Pore radius 5 35 Å 2. Pore length 5 600 A 3. Fractional pore area 5 5% 4. Blood plasma viscosity 5 0.02 poise (dyne s cm22) The “fractional pore” is the total area of the pores divided by the total area of the membrane. A dyne is a g cm s22. B. Calculate the glomerular filtration rate (GFR) assuming a total area for both kidneys of 1.5 m2 and a driving force ΔP 5 20 mmHg. The GFR is the total volume of ultrafiltrate produced per minute. Its units should be in cm3 min21. Make it so. 15. The observed osmotic pressure of solutions of plasma proteins increases more rapidly than concentration. Empirical fits to the concentration dependence of osmotic pressure are given by Landis and Pappenheimer (Handbook of Physiology, vol 2, section 2, pp. 9611034, 1963): πalbumin 5 2:8C 1 0:18C2 1 0:012C3 where π is in units of mmHg and C is in units of g% (i.e., g of protein per deciliter of plasma). A. According to this equation, as C becomes more dilute the relation approaches van’t Hoff’s Equation. Keeping in mind the units of the variables, what is the molecular weight of albumin? Assume that the temperature is 37 C. (By the way, osmotic measurements were the first measurements of protein molecular weights.) B. What is the contribution of albumin to the osmotic pressure of plasma when it contains 4.0 g% of albumin? C. The osmotic pressure of plasma proteins and associated ions is called the oncotic pressure. If the plasma oncotic pressure is 25 mmHg, how much of the oncotic pressure is contributed by globulins, fibrinogen, and other components? 16. Assume that serum albumin is a sphere of diameter 31 Å. Assume that the glomerular membrane is pierced by pores of equivalent diameter of 35 Å. A. Give an estimate of σ for albumin for the glomerular membrane. B. Calculate the concentration of albumin in the ultrafiltrate. C. If the GFR is 120 mL min21, calculate the daily filtered load of albumin (how much is filtered every day). How does this compare with the recommended dietary intake of 0.8 g protein per kg body weight per day? 17. The value of Lp for the red blood cell is about 1.8 3 10211 cm3 dyne21 s21. Its surface area is about 1.35 3 1026 cm2 (Solomon, Methods in Enzymology, pp. 192222, 1989). A. What is the initial osmotic flow if the osmolarity inside is initially 300 mOs M and the osmolarity outside is 275 mOs M? (Assume σ 5 1.0 for all solutes.) B. If the volume of the cell is 100 3 10212 cm3, how long would it take to double its volume provided that the osmotic pressure and area of the membrane and Lp did not change? C. In the case described, how much water would be required to enter the cell to equilibrate the osmotic pressure between inside and outside? Assume that the outside bath is essentially infinite so that its osmotic pressure is kept constant. 18. Osmotic pressure is one of a class of properties of solutions that are called colligative properties. The others in this class include vapor pressure depression, boiling point elevation, and freezing point depression. These properties are different expressions of the same phenomenon: the lowering of the activity of water by dissolution of solutes. Various osmometers have been made using one or another or these properties. Table 2.PS2.1 shows several solutions of sucrose and glucose, their water concentrations, TABLE 2.PS2.1 Solute Concentration, Water Concentration, and Freezing Point Depression in Sucrose and Glucose Solutions Sucrose Solutions Glucose Solutions [Sucrose] (M) [Water] (M) Δ ( C) [Glucose] (M) [Water] (M) Δ ( C) 0 55.45 0 0 55.46 0 0.029 55.12 0.06 0.028 55.28 0.05 0.059 54.77 0.11 0.056 55.11 0.10 0.081 54.42 0.16 0.084 54.94 0.16 0.118 54.07 0.23 0.112 54.76 0.21 0.179 53.72 0.35 0.140 54.59 0.27 Source: Data from Handbook of Chemistry and Physics, CRC Co, Cleveland, OH, 1965. 201 202 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION and their freezing point depression. Because dissolving solutes invariably dilutes the solvent, water, you can see from the table that the water concentration also decreases with increasing solute concentration. Sucrose, however, is twice as large as glucose, so we might expect that dissolution of sucrose would dilute the water further. Therefore, the colligative properties of solutions cannot be proportional to both solute and solvent concentration. Plot separately the freezing point depression against the water concentration and against the solute concentration. Which relationship shows the same dependence? From this result, do you expect osmosis to be dependent on solvent water concentration or on solute concentration? 19. From the data in Table 2.PS2.1, calculate the coefficient relating freezing point depression to molarity in the equation: ½2:PS2:1 PL 2 PR 5 π 5 RT apure water ln asolution Vw Here the subscripts L and R refer to the left and right sides of the semipermeable membrane, respectively, where pure water is on the right side. If the solution was ideal, we can use the mole fraction for the activity and this equation becomes ½2:PS2:3 π52 RT ln Xw Vw In the further approximation of a dilute solution, this equation was transformed to ½2:PS2:4 [Sucrose] Observed RT Cs (M) π (atm) (atm) 0.098 2.47 2.40 2RT/Vw ln RT/Vw ln xw (atm) (P0/Ps) (atm) 2.48 2.50 0.824 27.2 20.4 24.6 26.6 1.399 58.4 35.1 48.8 57.3 1.823 95.2 45.5 72.6 93.2 2.146 139.0 55.7 96.0 135.6 2.55 187.3 64.5 118.9 186.5 Source: Data from Glasstone, Textbook of Physical Chemistry, van Nostrand, 1946. Tf 2 T 5 Δ 5 kf C where Tf is the freezing point of the pure liquid, T is the freezing point of the solution, kf is the coefficient, and C is the concentration. 20. Chapter 2.7 gives us several expressions for the osmotic pressure. The most complete is derived from Eqn [2.7.6] and is ½2:PS2:2 TABLE 2.PS2.2 Values for the Observed Osmotic Pressure and the Osmotic Pressure Calculated from the van’t Hoff Equation, from the Mole Fraction of Water, and from the Measured Vapor Pressure of Solutions π 5 RT Cs The activity of water in Eqn [2.PS2.2] is measured by the vapor pressure. Table 2.PS2.2 tabulates the calculations of the osmotic pressure from the vapor pressure measurements, the mole fraction of water and from the concentration of solute. Plot the ratio of each of the observed osmotic pressures to the calculated osmotic pressure, in columns 3, 4, and 5, against the sucrose concentration. The values you calculate for the ratio of the observed π to the calculated π in column 3 is ϕ(5π/RT ln Cs), the osmotic coefficient. The ratio of the observed π to column 4 (i.e., (π/[ 2 RT/V w ln xw]) defines the rational osmotic coefficient, g. A. Why is the equation using the mole fraction of water a better predictor of the osmotic pressure than the van’t Hoff Equation? TABLE 2.PS2.3 Molecular Weight and Concentration of Plasma Proteins. Protein Average Molecular Concentration in Weight Plasma (g dL21) Albumin 69,000 4.2 Fibrinogen 330,000 0.3 Immunoglobin G 150,000 0.8 Immunoglobin M 750,000 0.3 α2 Globulins 100,000 0.7 α1 Globulins 50,000 0.5 β Globulins 100,000 0.8 B. Why is the equation using the vapor pressure a better predictor of the osmotic pressure than the equation using the mole fraction of water? C. Are these methods good predictors of the osmotic pressure in dilute solutions? 21. KrebsHenseleit buffer has the following composition: 119 mM NaCl, 25 mM NaHCO3, 3.2 mM KCl, 1.2 mM MgSO4, 1.2 mM KH2PO4, 11 mM glucose, and 1.4 mM CaCl2. Calculate its osmolarity and osmotic pressure, at 37 C, assuming ϕ 5 1 for all solutes. Would you expect this solution to be hypertonic, hypotonic, or isotonic to mammalian cells? 22. Plasma contains a variety of proteins that exert osmotic pressure. A list of these, with their molecular weights and approximate protein concentrations, is given in Table 2.PS2.3. From these data, calculate the approximate osmotic pressure in plasma that is due just to plasma proteins. 23. A. The device shown in Figure 2.PS2.2 was used to determine the steady-state flow and pressure across a dialysis membrane. The data Problem Set Syringe drives fluid into the compartment Syringe pump O ring retainers Motor Dialysis membrane 1 L beaker Pressure transducer measures pressure Water Pressure increases and forces fluid out to match inflow Recorder shows pressure increase until steady - state is reached FIGURE 2.PS2.2 Device used to determine LP. Water was injected into the inner chamber at a known rate using a syringe pump. This increases the pressure within the chamber and forces fluid out through the membrane. The pressure increases as more water is injected and eventually a steady state is reached in which the rate of injection matches the rate of filtration through the membrane. The steady-state pressure is measured continuously by a pressure transducer and recorded. TABLE 2.PS2.4 Pressure at Steady-State Flow Across a Dialysis Membrane Flow Rate (cm3 min21) 0 Syringe for addition or removal of solutions Pressure (mmHg) 0 0.0097 180 0.0194 360 0.0388 680 that were obtained are given in Table 2.PS2.4. The area of the membrane that was exposed to flow was 90.5 cm2. Determine LP for the membrane. B. The membrane in part A was used to separate pure water on the outside from a 0.75 M sucrose solution on the inside. The flow across the membrane was measured using the device shown in Figure 2.PS2.3. 24. When vesicles of the cardiac sarcoplasmic reticulum (CSR) are incubated with ATP, Mg21 and Ca21, they take up Ca21 and reach a pseudo steady state. This is a steady state that changes, but slowly. The uptake of Ca21 is mediated by the SERCA2a Ca-ATPase. The uptake reaction can be quenched by adding EGTA to the external solution, which binds the Ca21 outside of the vesicles, or by adding glucose plus hexokinase, that converts the ATP to ADP and glucose-6 phosphate. When the uptake reaction is stopped, Ca21 that was already taken up by the vesicles leaks out passively. A. The amount of Ca21 taken up by the vesicles is generally normalized to the amount Graduated tube O ring retainers 1 L beaker Dialysis membrane Water Magnetic stir bar Magnetic stirrer FIGURE 2.PS2.3 Device for measuring osmotic flow at constant ΔP 5 0. The inner compartment could be drained and then filled with various experimental solutions. The outer compartment contained pure water. As fluid enters the compartment across the dialysis membrane as a result of the osmotic pressure difference, it forces fluid down the horizontal tube without any increase in the hydrostatic pressure. The rate of fluid flow can be estimated from the rate of fluid movement down the tube. of CSR protein in mg rather than being expressed as a concentration. A typical steady-state Ca21 uptake is 40 nmol mg21. In separate experiments, the enclosed 203 204 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION volume of the CSR vesicles was determined to be 5 μL mg21. What is the approximate concentration of Ca21 inside the vesicles at steady state? B. The average vesicle size determined by electron microscopy is about 150 nm. What is the volume and surface area of a vesicle this size, assuming it is a sphere? C. Given that the enclosed volume of the aggregate vesicles is 5 μL mg21, how many vesicles are there per mg of CSR protein? How much surface area is there per mg of CSR protein? D. The initial passive efflux at a load of 40 nmol mg21 when the pump is stopped is 16 nmol min21 mg21. Convert this to a flux in units of nmol cm22 s21 by dividing by the surface area per mg of CSR protein and converting min to s. E. From the information in part D, What is the passive permeability to Ca21 in units of cm s21? 25. Oral Rehydration Solution used for Oral Rehydration Therapy, as recommended by WHO/UNICEF, has the following composition: 2.6 g NaCl 2.9 g Na3C6H5O7∙2 H2O (trisodium citrate dihydrate) 1.5 g KCl 13.5 g glucose per L of solution. Calculate the osmolarity of this solution. Trisodium citrate will dissociate fully at neutral pH to 3 Na1 ions and one citrate23 ion. Assume complete dissociation of NaCl and KCl. Cell Signaling Learning Objectives G G G G G G G G Distinguish among autocrine, paracrine, endocrine, neural, and neuroendocrine signaling Explain why Ca21 is a special ion with respect to signaling Describe in general terms what ligand-gated ion channels do List the major steps in turning on and off of heterotrimeric G-protein-coupled receptors Distinguish among Gs, Gi, Gq, and G12/13 mechanisms Describe the steps in a Gs mechanism’s activation of PKA and how they differ among tissues Describe the steps in a Gq mechanism’s activation of CAM kinase List the four types of catalytic receptors Describe the steps in a JAKSTAT pathway Describe in general terms lipophilic signaling molecules' effects on gene transcription 2.8 CELL-TO-CELL COMMUNICATION CAN ALSO USE DIRECT MECHANICAL, ELECTRICAL, OR CHEMICAL SIGNALS Mechanical signals can originate in the external environment, as in the case of sensory transduction, or they can be the signals from another cell. Mechanical signaling requires close contact of cells and generally occurs through cell junctions as discussed in Chapter 2.2. Mechanical force originates on filaments within cells that eventually connect to the extracellular matrix through cytoskeletal elements. Transmission of these forces occurs through the extracellular matrix, but it is also sensed by neighboring cells. All of the pressure sensors in the body are really stretch receptors, in which mechanical stretch is transduced into electrical signals or chemical signals. SIGNALING TRANSDUCES ONE EVENT INTO ANOTHER Electrical signals are usually used within the cell, as part of a signaling pathway to communicate intracellularly, and most often to move the signal rapidly from one place in the cell to another. Less frequently, direct electrical coupling occurs between cells. Such electrical coupling uses gap junctions, whose structure was discussed in Chapter 2.1. This mechanism is vitally important in coordinating some smooth muscle contraction and cardiac contraction. In its broadest context, cell signaling involves the transduction of some event into another event. In sensory transduction, a sensory cell is exposed to some external signal that is transduced to produce a nervous signal, the action potential. As we will see later in Chapter 3.2, this action potential can move along cell membranes to rapidly convey the signal, the action potential, to remote parts of the sensory neuron. The action potential is then transduced to release neurotransmitter at the synapse—the gap between one neuron and another. The neurotransmitter is then transduced to form the response of the postsynaptic cell, the one on the other side of the synapse. In the case of cutaneous (skin) senses, the original sensory signal is mechanical—a push or a pull on the nerves in the skin. The mechanical signal is transduced to an electrical signal, and the electrical signal is then transduced to a chemical signal. This simple series of events illustrates the use of mechanical, electrical, and chemical signals in the body (see Figure 2.8.1). Chemical signals do not require close contact and can be classified according to the distances involved, the mechanism of transmission, and the target of the chemical signals. These various classes of inter- and intracellular communication are shown in Figure 2.8.2. In some cases the signaling molecule remains bound to the cell and so transmission of this signal requires contact between the signaling cell and its target. This contactdependent signaling is important in development and in immune responses. In other cases, the signaling cell releases a chemical that either acts locally (a paracrine or autocrine signal) or travels through the blood to act on remote targets (an endocrine signal). Autocrine signals have receptors on the signaling cell itself or others like it. Paracrine signals affect other types of cells located in the neighborhood of the signaling cell. Neurons also release signaling molecules, usually at the end of a long extension of the cell, the axon. When the neural chemical signal enters the blood and acts on distant targets, it is called a neuroendocrine signal. When it acts in the vicinity of its release site, it is a neurotransmitter. G G 205 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00019-7 206 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION The sensory cell transduces a mechanical stimulus to an electrical signal, the action potential At a terminal, the action potential is transduced to a chemical signal, a neurotransmitter Sensory neuron cell body Interneuron Mechanical stimulus Action potential The final response can be a mechanical, chemical, or electrical response The action potential moves along neuron cell processes to convey the signal to distant parts of the cell The postsynaptic cell transduces the chemical signal to another electrical signal FIGURE 2.8.1 Transduction of signals. Some kinds of sensory cells can transduce mechanical stimuli to electrical signals which can be conveyed along their surface for rapid spatial relay of the signal. At the end of the cell, the electrical signal is transduced to a chemical signal to convey the signal across the gap between the cells. The postsynaptic cell transduces this chemical signal back to an electrical signal. SIGNALS ELICIT A VARIETY OF CLASSES OF CELLULAR RESPONSES Intra- and intercellular signals begin a cascade of events that eventually changes cell behavior. The response of cells to signaling events includes altered: G G G G G G ion transport; metabolism; gene expression or differentiation; shape, movement, or force production; cell growth or cell division; apoptosis or programmed cell death. ELECTRICAL SIGNALS AND NEUROTRANSMITTERS ARE FASTEST; ENDOCRINE SIGNALS ARE SLOWEST The speed of response to an initial stimulus depends on the mode of delivery of the signal and the mechanism of the response in the target cells. Electrical signals are the fastest way to transmit a signal from one place in the body to another, in milliseconds, but the overall response depends on what happens in the target cell. If the response involves changes in activity of proteins already present in the target cell, the response can be rapid. If the response involves altered gene expression that requires synthesis of new protein, response can take hours. If it involves altered cell growth, it can take days to complete. Neurotransmitter signaling is the fastest response, followed by changes in cell shape or the development of force. Endocrine signals are slowest but last longer. VOLTAGE-GATED ION CHANNELS CONVEY ELECTRICAL SIGNALS Ion channels allow ions to cross biological membranes that they otherwise cannot penetrate. Because they are electrically charged, movement of ions makes a current, and the current either charges or discharges the membrane—the current alters the membrane potential. Thus the voltage-gated ion channels are largely responsible for alterations in the membrane potential that is rapidly conveyed through the cell. VOLTAGE-GATED Ca21 CHANNELS TRANSDUCE AN ELECTRICAL SIGNAL TO AN INTRACELLULAR Ca21 SIGNAL Cells maintain a very low intracellular [Ca21] (,100 nM) to avoid Ca21 precipitation with phosphate and organic phosphates (ATP, etc.) present in high concentrations in the cytoplasm. The low cytoplasmic [Ca21] allows increases in cytoplasmic [Ca21] to be used as a signal. Multiple types of voltage-gated Ca21 channels (voltagedependent calcium channel, VDCCs) reside on the surface membrane of many cells. Depolarization of the cell membrane opens these channels, causing Ca21 to move from the ECF, with 1.2 mM [Ca21], to the intracellular compartment. The cytoplasm contains a number of proteins that bind Ca21 with high affinity and that change shape or activity upon Ca21 binding. The effects of increasing cytoplasmic [Ca21] include the following: 1. Stimulussecretion coupling: The increased [Ca21] binds to Ca21 sensors on vesicles, causing the fusion of the secretory vesicles with the plasma membrane and release of secreted products into the ECF. 2. Excitationcontraction coupling: The increased [Ca21] binds to Ca-sensitive elements on contractile filaments or cytoskeletal elements, causing either force development or shortening by muscle cells. 3. Calmodulin-dependent activation of enzymes: Calmodulin is a small cytosolic protein that binds four Ca21 molecules and then activates many enzymes such as myosin light chain kinase in smooth muscle. C ell Signal ing Mechanical signals from the cytoskeleton are transferred to neighboring cells through the extracellular matrix that links to the cytoskeleton Mechanical signals Electrical signals Electrical signals can be directly transferred to a neighboring cell through gap junctions Chemical signals Receptor Signaling cell Contact dependent Target cells Signaling cell Response Signaling cell Target cell ECF ECF Response Response Chemical signal (local hormone) Paracrine Autocrine Signaling cell Target cell Endocrine ECF Bloodstream ECF Neuron Neuroendocrine Rapid electrical signal Response ECF Bloodstream Target cell ECF Response Axon Neurotransmitter Neuron Target cell Neuron Response Diffuse Response Discrete FIGURE 2.8.2 Main classes of signaling. Mechanical signals can pass from cell to cell through filaments in the extracellular matrix attached to membranebound proteins in the surfaces of cells, particularly at cell junctions such as desmosomes. Electrical signals can also pass directly from one cell to another through gap junctions. Some signaling molecules remain bound to the surface and so the signal affects the target cell only by direct contact of the signaling cell with the target cell. Cells can release chemical signals that act locally. When they affect the signaling cell, or others like it, they are autocrine signals. When they affect other nearby cells, they are paracrine signals. Signals that are released into the bloodstream to affect distant target cells are endocrine signals. If they are released from long processes by neurons, they are neuroendocrine signals. Nerve cells release a variety of chemical signals at terminals near target cells. These are neurotransmitters. If they are released very close to clustered receptors on target cells, they are discrete neurotransmitters. If they are released into a general area to affect multiple cells, they are diffuse neurotransmitters. Electrical signals are most often used intracellularly to rapidly convey the action potential from one part of the cell to another. This is the fastest movement of a signal in the body. 4. Direct activation of enzymes: Ca21 can directly bind to some enzymes, such as PKC, and activate them. Figure 2.8.3 illustrates these aspects of Ca21 signaling in cells. LIGAND-GATED ION CHANNELS OPEN UPON BINDING WITH CHEMICAL SIGNALS Fast release of chemical signals by electrical signals in nerve terminals, followed by an electrical signal in the target cell, is the fastest mechanism of signaling used in the body. This is the classic neurotransmitter mechanism: the electrical signal (the action potential) on the presynaptic cell propagates down the axon to the nerve terminal where it causes a local intracellular Ca21 signal that releases chemical neurotransmitter into the ECF immediately adjacent to the postsynaptic cell. The chemical signal then binds to a receptor which is also an ion channel, causing a flux of ions across the postsynaptic cell membrane and a resulting electrical signal in the postsynaptic cell. Three major classes of surface membrane, ligand-gated ion channels (LGICs) have been identified, as shown in Figure 2.8.4, and many of these have multiple subtypes. The major types are distinguished by their structures. Receptors for acetylcholine, serotonin (55 hydroxy tryptamine), gamma amino butyric acid (GABA), and glycine all consist of five subunits. Each family, such as nicotinic ACh receptor, has 207 208 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 1 An action potential or depolarization of the surface membrane opens voltagedependent Ca channels on the surface 2 Ca rushes in from the high ECF [Ca2+] to the low cytoplasmic [Ca2+] 3 Increased cytoplasmic [Ca2+] binds to Ca sites on target proteins Ca2+ VDCC Synaptotagmin 4D ... directly activating some enzymes Ca2+ 4A ... causing stimulus-secretion coupling in neurons or most endocrine cells RyR1 PKC Ca2+ CSQ Ca2+ Ca2+ Calmodulin Ca2+ Ca2+ TnC MLCK Ca2+ 4B ... activating contraction in muscle cells Ca2+ Ca2+ 2+ Ca Inactive Active 4C... binding to calmodulin, activating a variety of enzymes FIGURE 2.8.3 Electrically coupled calcium signaling. Depolarization of the cell membrane opens a calcium channel that lets Ca21 into the cell. At rest, the cytoplasmic [Ca21] is very low. Upon stimulation, influx of Ca21 raises the [Ca21] enough for Ca21 to bind to Ca21-binding sites on specific proteins. Ca21 binding to synaptotagmin causes fusion of secretory vesicles with the plasma membrane. Binding to troponin C (TnC) activates force in skeletal or cardiac muscle. Ca21 binding to calmodulin activates a number of enzymes such as MLCK (myosin light chain kinase) involved in smooth muscle contraction. In other cases, Ca21 directly activates some enzymes; PKC is shown. RyR1 is the ryanodine receptor on the endoplasmic reticulum membrane; CSQ is calsequestrin, a calcium-binding protein in the lumen of some ER membranes. Tetrameric channels Pentameric channels Acetylcholine Serotonin Na+,K+ Nicotinic ACh GABA Glycine Na+ Cl– 5-HT3 GABAA Trimeric channels Glutamate Cl– Gly Ca2+ Na+,K+ NMDA AMPA GluN1–3 GluA1–4 ATP Na+,K+ Kainate GluK1–5 Na+,K+, Ca2+ P2x P2x1–7 FIGURE 2.8.4 Ligand-gated ion channels. These channels reside in the plasma membrane and respond to specific ligands by allowing specific ions to cross the membrane. The channels are classified according to their structure and agonist or chemical signal that opens the channel. The names of the channels are at the bottom of the figure, and alternate naming conventions have been proposed. Each family of channels has multiple isoforms that depend on the subunit make-up of the channels. C ell Signal ing 4 1 Chemical signal binds to its receptor Chemical signal 2 α Subunit hydrolyzes GTP and dissociates from effector 3 α Subunit dissociates from βγ GTP exchanges for subunit and both may bind to and GDP on the α subunit change effector behavior Activated state Basal state 5 α Subunit reassociates with βγ subunit and with the receptor to return to the basal state Effector 1 Effector 2 Receptor GDP Gα Subunit βγ Subunits Pi GTP GDP FIGURE 2.8.5 General scheme for heterotrimeric GPCRs. The receptors are membrane-bound proteins that bind chemical signals. The heterotrimeric Gprotein consists of an α subunit that binds and hydrolyzes GTP and a βγ subunit that does not dissociate. Binding of the ligand to its receptor triggers the exchange of GTP for GDP and subsequent dissociation of the α subunit and βγ subunit and both are then able to alter the behavior of effector targets in the cell. The α subunit spontaneously hydrolyzes its GTP, and the α subunit reassociates with the βγ subunit to return to the basal, unstimulated state. multiple subtypes consisting of different subunits, but each member of the subtype responds to one chemical signal, acetylcholine in this case. Receptors for glutamate each have four subunits, and this family of LGIC has further subtypes distinguished by artificial agonists (stimulators of the receptor), NMDA (N-methyl D-aspartic acid), AMPA (α-amino-3-hydroxy-5-methyl-4 isoxazole propionic acid), and kainate. HETEROTRIMERIC G-PROTEINCOUPLED RECEPTORS (GPCRS) ARE VERSATILE G-protein-coupled signaling pathways are versatile because of their modular structure: they consist of receptors, heterotrimeric G-proteins, and effectors. Receptors are membrane-bound proteins that bind signaling molecules on the external surface of cells. Binding alters the conformation of the receptor, and this change is transferred to a heterotrimeric G-protein, consisting of α, β, and γ subunits. In the unstimulated state, the Gα subunit binds GDP. Upon ligand binding to its receptor, the receptor causes the Gα GDP to exchange GTP for the bound GDP, and the Gα GTP dissociates from the βγ subunits. The dissociated subunits can then bind to effector molecules, exerting some change in their behavior. The Gα GTP has inherent GTPase activity, so it reverts back to Gα GDP and reassociates with the βγ subunit. The overall plan of the G-protein signaling pathway is shown in Figure 2.8.5. THERE ARE FOUR CLASSES OF G-PROTEINS: GαS, GαI/GαO, GαQ, AND Gα12/Gα13 β ADRENERGIC STIMULATION IS AN EXAMPLE OF A GαS MECHANISM A number of signaling molecules can bind to a receptor that is coupled to a Gαs-protein. All Gαs-protein’s α subunit binds to a membrane-bound enzyme, adenylyl cyclase, which converts ATP into 30 ,50 cyclic AMP, or cAMP. Increasing the concentration of cAMP activates protein kinase A (PKA), which phosphorylates (adds a phosphate to) specific target proteins. Phosphorylation of these target proteins alters cell behavior, the final consequence of exposure to the chemical signal. The basic pathway for this signaling mechanism is shown in Figure 2.8.6. This signal is turned on by activation of adenylyl cyclase and then PK and then phosphorylation of proteins. It is turned off by the simultaneous inactivation of all three of these. The α subunit of the G-protein spontaneously hydrolyzes its GTP and reassociates with the βγ subunit. This inactivates adenylyl cyclase and stops its synthesis of cAMP. The cAMP is also broken down by phosphodiesterase (PDE) so that the increased [cAMP] is removed. The cAMP dissociates from the regulatory subunits of the PKA, causing it to inactivate so that target proteins are no longer phosphorylated. However, this does not reverse phosphorylation of proteins. The phosphorylated proteins are dephosphorylated by specific enzymes called protein phosphatases. There are four classes of serine/threonine phosphoprotein phosphatases: PP1, PP2a, PP2b, and PP2c, which dephosphorylate proteins phosphorylated at serine and threonine residues. Fine control of this system is provided by the regulation of both adenylyl cyclase, as described, and by the regulation of both PDE and the protein phosphatases. α2 ADRENERGIC STIMULATION IS AN EXAMPLE OF A GαI/GαO MECHANISM Other molecules bind to their G-protein-coupled receptor (GPCR) and release a Gα subunit that inhibits adenylyl cyclase. These are referred to as Gi mechanisms. An example is epinephrine binding to α2 receptors on neurons and is illustrated in Figure 2.8.7. Other members of this class achieve the same end—reduction in cAMP levels—by activating PDE. Retinal photoreceptor 209 210 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 1 Epinephrine binds to 2 β1 or β2 receptor Epinephrine 3 GTP exchanges for GDP on the α subunit Activated state α Subunit dissociates from βγ subunit, binds to adenylyl cyclase and activates it 4 Adenylyl cyclase converts ATP to cAMP Basal state Adenylyl cyclase β receptor GDP Gα Subunit βγ Subunits Phosphodiesterase hydrolyzes cAMP Pi GTP GDP 5 cAMP binds to regulatory ATP subunit of PKA, releasing the active catalytic subunit cAMP Phosphodiesterase AMP Regulatory subunit Inactive PKA Catalytic subunit PKA catalytic subunit phosphorylates target proteins, changing their behavior Active PKA ATP ADP P P 6 Protein phosphatase FIGURE 2.8.6 Mechanism of Gs-coupled receptors. Many different kinds of ligands bind to Gs-coupled receptors. Epinephrine is shown, which binds to β1 and β2 receptors that are coupled to Gs proteins. Activation follows the general scheme shown in Figure 2.8.5. Here the Gαs subunit binds to adenylyl cyclase, activating it and increasing the cytoplasmic concentration of cAMP. This activates PKA that phosphorylates a variety of target proteins. The signal is turned off by (1) hydrolysis of GTP by the Gα subunit and dissociation and removal of activation of adenylyl cyclase; (2) hydrolysis of cAMP by PDE and removal of activation of PKA; (3) dephosphorylation of target proteins by protein phosphatases. 1 Epinephrine binds to 2 α2 receptor GTP exchanges for Epinephrine GDP on the α subunit 3 α Subunit dissociates from βγ subunit, binds to adenylyl cyclase and inhibits it Activated state Adenylyl cyclase Basal state α2 Receptor Gα Subunit GDP – αγ Subunits GTP GDP ATP cAMP FIGURE 2.8.7 Mechanism of Gi-coupled receptors. Here epinephrine binds to α2 receptors, which is followed by the inhibition of adenylyl cyclase and a reduction in cytoplasmic [cAMP]. cells, for example, activate cGMP PDE. This mechanism is illustrated in Figure 2.8.8. Still other ligands, such as acetylcholine, bind to M2 receptors that cause inhibition of adenylyl cyclase, and the βγ subunit activates a K1 channel (see Figure 2.8.9). Thus this class of GPCR exerts a variety of effects including direct inhibition of adenylyl cyclase, activation of PDE, and direct activation of K1 channels. C ell Signal ing Light α Subunit dissociates from βγ subunit binds to cGMP phosphodiesterase and activates it Metarhodopsin II cGMP PDE βγ Subunits Activated state Basal state + Transducin GDP GMP Gα subunit GDP GTP Na+ Guanylyl cyclase + cGMP FIGURE 2.8.8 GPCR involved in retinal signal transduction. Through a series of steps, light converts rhodopsin to metarhodopsin II, which binds to a heterotrimeric G-protein called transducin. The Gα subunit exchanges GTP for GDP, dissociates from the βγ subunit, and activates cGMP PDE. Guanylyl cyclase in these cells makes cGMP continuously, and cGMP opens a Na1 channel. Degrading the cGMP reduces [cGMP] and therefore regulates the open state of the channel, producing an electrical signal. α Subunit binds to adenylyl cyclase and inhibits it Acetylcholine βγ Subunits K+ Activated state Adenylyl cyclase Basal state M2 receptor GDP Gα Subunit GTP βγ Subunit binds to K channel and activates it GDP ATP cAMP FIGURE 2.8.9 Gi mechanism involved in M2 GPCR. Acetylcholine binds to a variety of receptors. The M2 receptor, present in heart, couples binding of acetylcholine to a Gαi subunit that inhibits adenylyl cyclase. The βγ subunit released by acetylcholine binding directly activates a K1 channel. GαQ/Gα11 GPCR ACTIVATES PHOSPHOLIPASE C AND RELEASES CA FROM INTRACELLULAR STORES modulatory proteins, called GAPs, for GTPase Activating Proteins, facilitate the inactivation of these small monomeric GTPases. The overall plan is shown in Figure 2.8.11. A third class of GPCR activates phospholipase C on the surface membrane, which cleaves phosphatidyl inositol bisphosphate to produce diacylglycerol (DAG) and inositol triphosphate (IP3). The IP3 releases Ca21 from the endoplasmic reticulum, while the DAG activates protein kinase C (PKC) (see Figure 2.8.10). The superfamily of small monomeric GTPases is divided into several major subfamilies, including: Gα12/Gα13-COUPLED RECEPTORS ACTIVATE SMALL MONOMERIC GTPASES The Gα12 is the last of the four major families of heterotrimeric G-proteins (Gs, Gi, Gq, and G12) that we will discuss. Gα12 and Gα13 are linked to GTP exchange factors (GEFs) that activate small monomeric G-proteins by exchanging their bound GDP with GTP. A second set of Ras, Rho, Rab, Ran, and Arf These are all small (2040 kDa) proteins that are membrane bound due to covalent attachment of lipids (such as N-myristoylation on Arf proteins). The Rho GTPases regulate the cytoskeleton and play a role in the regulation of smooth muscle contraction. One of its effectors is Rho Kinase, a serinethreonine protein kinase that phosphorylates myosin light chain phosphatase and inactivates it (see Chapter 3.8). Rho is also involved in cell cycle progression and gene expression. The Rab family of monomeric GTPases regulates vesicular traffic as well as modulation of actin 211 212 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION α Subunit dissociates from βγ subunit, activates PLC Epinephrine Activated state DAG activates PKC Diacylglycerol Phospholipase C PIP2 Basal state α1 Receptor PKC Gα subunit GDP βγ Subunits P GTP GDP IP3 ATP Ca2+ Ca2+ Ca2+ IP3 binds to IP3 receptor on ER and releases stored Ca2+ to the cytoplasm ADP + Pi FIGURE 2.8.10 Mechanism of Gαq signaling. Binding of ligand to its GPCR results in release of the Gαq and βγ subunits. The Gαq subunit activates phospholipase C, which cleaves phosphatidyl inositol bisphosphate in the surface membrane, liberating DAG and IP3. IP3 releases Ca21 stored in the ER and DAG activates PKC that phosphorylates sets of target proteins. Pi GTPase activating proteins (GAPs) Active Inactive GTP GDP GTP GDP GTP exchange factors (GEFs) FIGURE 2.8.11 Heterotrimeric G-proteins with Gα12-activated GEFs that promote the exchange of GTP for GDP on small monomeric G-proteins. This converts the small monomeric G-proteins into an active form. Reversion to the inactive form is catalyzed by GAPs, GTPase activating proteins. The GEF binds to the Gα12 subunit. dynamics. Members of the Ran family regulate transport of materials between the cytoplasm and nucleus. Arf stands for “ADP-ribosylation factor.” All six mammalian Arfs are located in the Golgi and regulate vesicular transport there. THE RESPONSE OF A CELL TO A CHEMICAL SIGNAL DEPENDS ON THE RECEPTOR AND ITS EFFECTORS According to Figures 2.8.6, 2.8.7, and 2.8.10, epinephrine can bind to β GPCRs, with the final activation of adenylyl cyclase, increased [cAMP], and activation of PKA; or it can bind to α2 receptors with the inhibition of adenylyl cyclase, decreased [cAMP], and no activation of PKA; or it can bind to α1 receptors with subsequent activation of PLC, release of Ca21 from intracellular stores, and activation of PKC. These different effects typically occur in separate cells. Thus the response of the cell depends on the receptor for the chemical signal and not on the chemical signal alone. Further, some cells respond differently than others because they express entirely different sets of target proteins. In the liver, for example, the primary target for PKA from the stimulation of β adrenergic receptors is phosphorylase kinase, which then phosphorylates phosphorylase, activating it, and glycogen synthetase inactivating it. In the heart, the primary targets of β adrenergic stimulation are the voltagedependent Ca21 channels in the surface membrane, phospholamban on the sarcoplasmic reticulum, ryanodine receptors (RyR2) on the sarcoplasmic reticulum, and troponin I (TnI). These are illustrated in Figure 2.8.12. Thus the effects of epinephrine using the same type of receptor are glycogenolysis in the liver and increased contractile strength in the heart. These differences indicate that the final effect is a function of (1) the chemical signal, (2) its receptor, (3) the effector, and (4) the targets within the cell. CHEMICAL SIGNALS CAN BIND TO AND DIRECTLY ACTIVATE MEMBRANEBOUND ENZYMES A variety of extracellular chemical signals can bind to receptors on the surface membrane; these are amplifying enzymes or directly activate an amplifying enzyme. They are of four main types: receptor guanylyl cyclase, receptor serine/threonine kinase, receptor tyrosine kinase, and receptor-associated tyrosine kinase. These are illustrated in Figure 2.8.13. C ell Signal ing Ca2+ Epinephrine βγ Subunits Activated state Adenylyl cyclase Basal state Inactive PKA β receptor Ca2+ GDP Gβ subunit cAMP TnI ATP GTP GDP Regulatory subunit Catalytic subunit Myosin filament Actin filament Active PKA ATP Phosphorylase kinase (inactive) Ca2+ ADP P ATP Phosphorylase kinaase (active) ADP Phosphorylase (active) Phosphorylase (inactive) ADP + Pi ATP SERCA2a Phospholamban Glucose -1-Pi Ryanodine receptor Glycogen Glycogen synthase P (inactive) Glycogen synthase (active) ADP ATP Liver Heart FIGURE 2.8.12 Different effects of beta adrenergic stimulation of liver cells versus heart cells. In the liver, the primary response of beta adrenergic stimulation is activation of glycogenolysis through activation of glycogen phosphorylase through a cascade of protein phosphorylation reactions. In the heart, the primary response of beta adrenergic stimulation is faster activation and relaxation of the muscle through control of cytoplasmic [Ca21] by phosphorylation of voltage-dependent Ca21 channels on the surface of the cell, phosphorylation of the ryanodine receptor on the surface of the sarcoplasmic reticulum, phosphorylation of TnI on the contractile filaments, and phosphorylation of phospholamban to relieve inhibition of the SERCA2a Ca21 pump on the surface of the SR. MANY SIGNALS ALTER GENE EXPRESSION So far we have discussed signaling molecules that cannot penetrate the cell membrane. These bind to receptors on the surface of the cell, and the binding is transduced into a second messenger such as Ca21, cAMP, cGMP, IP3, and DAG, one of a number of small monomeric GTPases, or causes phosphorylation of intracellular proteins. A number of lipophilic signaling molecules penetrate the cell membrane and bind to receptors either in the cytoplasm or in the nucleus and alter the expression of specific genes in the cell. Hormones that alter gene expression this way include the sex hormones testosterone, estrogen, and progesterone; corticosteroids, including glucocorticoids, produced by the adrenal cortex; mineralocorticoids, produced by the adrenal gland that regulate electrolyte and water balance; vitamin D, which regulates calcium and phosphate balance, among other effects; and thyroid hormone and retinoic acid, which have effects in almost all cells. NUCLEAR RECEPTORS ALTER GENE TRANSCRIPTION The nuclear receptors constitute a superfamily of proteins that are structurally related and perform similar functions, but they exhibit specificity for binding ligands and specificity of action. These nuclear receptors include the following: estrogen receptors α and β (ERα and ERβ); androgen receptor (AR); progesterone receptor (PR); glucocorticoid receptor (GR); mineralocorticoid receptor (MR); vitamin D receptor (VDR); thyroid receptors α and β (TRα and TRβ); retinoic acid receptor types α, β, and γ (RARα, RARβ, and RARγ); and 9-cis-retinoic acid receptors (RXRα, RXRβ, and RXRγ). These nuclear receptors are restricted to the nucleus with the exception of the mineralocorticoid receptor (MR) and the GR which reside in the cytoplasm. Upon binding with its ligand, MR and GR move into the nucleus where they bind to the regulatory region of a modulated gene. Binding of the lipophilic ligand with its receptor changes the conformation of the receptorligand complex, allowing the receptor to bind to specific nucleotide sequences on the DNA, called response elements. Binding of these receptors begins a cascade of events that eventually activates the transcription of specific genes. The newly transcribed mRNA is transferred to the cytoplasm where it is translated into a protein. The set of proteins regulated by the hormones confers specific capabilities on the cells in which they are expressed. Thus cell function is regulated by controlling cell concentration of specific active proteins. 213 214 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Receptor guanylyl Receptor serine/threonine cyclase kinase Receptor tyrosine kinase TGFβ Insulin Receptor-associated ANP tyrosine kinase Insulin Type II receptor Growth hormone receptor Type I receptor GHR + P P PI-3K ATP GTP P cGMP PKB smad2/3 P P P IRS-2 IRS-1 P P JAK2 Grb-2 + P ATP Ras P STAT5 smad4 P P STAT or GR P DNA RNA polymerase II Nucleus FIGURE 2.8.13 Catalytic receptors. Chemical signals binding to the extracellular surface of the cell are coupled to enzymes that alter intracellular components. Receptor guanylate cyclases respond to atrial natriuretic peptide or brain natriuretic peptide, and binding increases the concentration of cGMP in the cell, which increases activity of cGMP-dependent protein kinase. Receptor serine/threonine kinases: dimers of a variety of growth factors including transforming growth factor beta (TGF-β), myostatin, and bone morphogenic protein (BMP) bind to an activin type II receptor that recruits an activin type I receptor and phosphorylates it. This active serine/threonine kinase then phosphorylates one of a family of proteins called smads (smad2 or smad3 is shown), which binds to smad4. The complex enters the nucleus and regulates gene expression. Receptor tyrosine kinases: Insulin or insulin-like growth factor (e.g., IGF-1 and IGF-2) binds to the insulin receptor. The activated receptor phosphorylates several intracellular substrates. Insulin receptor substrate-1 (IRS-1) is shown. The phosphorylated proteins can activate two main pathways: the PI-3K (phosphatidyl inositol 3 kinase) pathway activates PKB and PKC downstream. The Ras pathway is activated by Grb-2 that binds to IRS and then activates a Ras GTP exchange protein leading eventually to transcription factors. Receptor-associated tyrosine kinases are used by growth hormone (shown), prolactin and erythropoietin, and most interleukins. Their binding induces close proximity of two receptors that bind members of the Janus family of tyrosine kinases (JAKs: JAK1-3 and TYK2). These transphosphorylate themselves and their receptor and phosphorylate proteins called STAT (for signal transduction and activation of transcription). These form dimers with other transcription factors and regulate gene transcription. NUCLEAR RECEPTORS RECRUIT HISTONE ACETYLASE TO UNWRAP THE DNA FROM THE HISTONES Heterochromatin is highly condensed DNA that cannot be transcribed. Euchromatin is more easily accessible for the assembly of transcriptional subunits, and DNA in this configuration has a higher rate of transcription. The configuration of chromatin is regulated by the acetylation of histones. Histones are a family of proteins, described in Chapter 2.2, that form a complex with DNA called a nucleosome that is stabilized by the attraction of the negatively charged DNA to the positively charged histones. In this form, the DNA cannot be transcribed. Acetylation reduces the association of the DNA with the histones by reducing the positive charge on the histones. Deacetylation promotes condensation of the DNA into heterochromatin. The enzyme histone acetyl transferase (HAT) sticks acetyl groups on the histones, and histone deacetylase removes them. Both of these enzymes associate with coactivators and repressors of transcription. A variety of proteins associated with nuclear receptors possess HAT activity. NUCLEAR RECEPTORS RECRUIT TRANSCRIPTION FACTORS Once acetylation of the histones has allowed the reorganization of chromatin, several other complexes of proteins bind to the DNA to initiate transcription. The transcription factor TFIID (for transcription factor polymerase II) is a complex of proteins that binds to a TATAA sequence on the DNA some 2530 nucleotides upstream of the initiation site (see Chapter 1.3). TFIID contains a TATAA-binding protein (TBP), which binds directly to the TATAA sequence, and a series of other factors called TBP-associated factors (TAFs). TBP then binds a second basal transcription factor, TFIIB. This allows the binding of RNA polymerase II to the complex, which is then fully activated by the binding of assorted other transcription factors, TFIIF, TFIIE, and TFIIH. The nuclear receptor influences transcription through specific proteins that interact both with the nuclear receptor on its recognition site and with the RNA polymerase complex on its initiation site. This interaction is possible because the DNA can form loops that closely appose the nuclear receptors and the C ell Signal ing Nucleosome H2A H2B H3 H4 RNA polymerase II 6 TFIIE 4 TFIIF TFIIH TATAA H1 TFIIB TBP TAF HAT 3 SRC-1 TFIID 5 1,25 (OH)2D3 1 DRIP Ligand binding site 2 VDR RXR ACTTGG ACTGGG 3′ 3nt VDRE RXR VDR 5′ AF-1 Dimerization site DNA binding site FIGURE 2.8.14 Simplified model of activation of transcription of DNA by vitamin D3. Not all steps in this process are established, and the figure is meant to convey some of the players and their postulated roles. It is not to be taken too seriously. The active form of vitamin D3, 1,25(OH)2D3, binds to its receptor, VDR (vitamin D receptor), which is nonspecifically bound to DNA. Binding of the ligand (1) results in dimerization of the VDR with RXR (9-cis-retinoic acid receptor) and binding of the dimer to specific vitamin D-responsive elements on the DNA (VDRE) (2). These are similar repeat motifs on the DNA as indicated by the sequences ACTTGG and ACTGGG. Transactivation begins with the recruitment of coactivators with HAT activity (3). One of these is SRC-1 for the steroid receptor coactivator. Acetylation of histones causes chromatin remodeling that facilitates transcription. Actual initiation of transcription requires the binding of TBP to the TATAA box, along with several TAFs (4). This complex is TFIID, for transcription factor for RNA polymerase II. TFIID then binds TFIIB, which forms a bridge to RNA polymerase II. Several proteins, called vitamin D-receptor interacting proteins, or DRIPs, form a bridge to RNA polymerase II, stabilizing the preinitiation complex (5). Transcription is then initiated (6). preinitiation complex. This process is illustrated for VDR in Figure 2.8.14. In other cases, nuclear receptors can also regulate gene expression by suppressing transcription. For example, glucocorticoids suppress the effects of transcription factor nuclear factor κB (NF-κB), which stimulates genes as part of the inflammatory response. Glucocorticoids reduce inflammation by this effect (see Chapter 9.5). cAMP Phosphodiesterase AMP Cytoplasm Regulatory subunit Inactive PKA Catalytic subunit OTHER SIGNALING PATHWAYS ALSO REGULATE GENE EXPRESSION The binding of signaling molecules to nuclear receptors is only one of many routes for the regulation of gene expression. As noted earlier, receptor serine/ threonine kinases can alter gene expression through phosphorylation of smads; receptor tyrosine kinases alter gene expression through Ras, and receptorassociated tyrosine kinase alters gene expression through phosphorylation of STATs. Signals that affect 30 ,50 cyclic AMP levels also regulate gene expression. Specific genes possess a regulatory sequence called the cAMP response element or CRE. PKA phosphorylates CREB (CRE-binding protein), a transcription factor that binds to the CRE. CREB can also be phosphorylated by CaM kinase II and CaM kinase IV. Transcriptional activation by CREB requires coactivators including CREB binding protein (CBP) with a molecular weight of 300 kDa. CREB forms homodimers to activate transcription, but it can also form heterodimers with CREM (CRE modifier) that either activate or inhibit transcription. A schematic of CREB’s involvement in the regulation of transcription is shown in Figure 2.8.15. Active PKA ATP ADP CREB CREB Pi CBP Nucleus CREB Pi CREB CREB Pi Pi DNA CRE FIGURE 2.8.15 Modulation of gene transcription by cyclic AMP-dependent PKA. Activation of adenylyl cyclase occurs on the surface membrane of the cell through a Gs-coupled receptor for a hormone. The increased cAMP activates PKA by dissociating the regulatory subunit from the catalytic subunit. The catalytic subunit translocates to the nucleus where it phosphorylates CREB, a protein that binds to cyclic AMP responsive elements (CRE) on the DNA strand. CREB binds to CBP and may form homodimers or heterodimers and may activate or inhibit transcription. A variety of cofactors are recruited to complete the mechanism. 215 216 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Small molecular weight, llipophilic signal Polypeptide hormone 3 2 1 4 Trimeric G-proteins Amplifying enzyme DAG PLC PKC AC + NMDA Ca2+ Ca2+ VDCC Secretory protein Ca2+ IP3 P ATP cAMP JAK PP CREB PP P STAT5 CREBP mRNA + Growth hormone ATP PKA Ca2+ 5A P STAT or GR + Proteins RNA Polymerase II - Bind to response element Nuclear receptors Nucleus PKB P Insulin + Ribosome PI-3K 5B IRS-1 + P Ras FIGURE 2.8.16 Summary of major signaling pathways. (1) Voltage-dependent channels open to convey electrical responses. (2) LGICs convert chemical signals into electrical signals. (3) Four different classes of heterotrimeric GPCRs convert extracellular chemical signals into intracellular chemical signals. The GPCR coupled to phospholipase C (Gq) and to adenylyl cyclase (Gs) is shown, but other responses also occur. (4) Small lipophilic signaling molecules enter the cell and affect gene transcription and other processes. (5) Extracellular chemical signals, larger proteins, activate enzymes that produce intracellular signals. Growth hormone and insulin receptor types are shown. SUMMARY OF SIGNALING MECHANISMS A synopsis of the signaling mechanisms discussed in this chapter is shown in Figure 2.8.16. The main classes of signaling include the following: (1) voltage-gated ion channels, including the fast Na1 channel and K1 channels involved in action potential origination and propagation. These channels maintain electrical signaling that is rapidly conveyed over the surface of the cell. (2) LGICs, the mainstay of synaptic transmission between neurons and between nerve and muscle. (3) Heterotrimeric GPCR, including four main subtypes: those that excite adenylyl cyclase (Gs mechanisms); those that inhibit adenylyl cyclase (GI mechanisms); those that activate phospholipase C, releasing IP3 and DAG (Gq mechanisms); and those that stimulate GEFs to activate one of a family of small GTPase proteins. (4) Extracellular signals, which directly activate enzymes such as guanylyl cyclase, receptor serine/threonine kinase, receptor tyrosine kinase, and receptor-associated tyrosine kinase. (5) Signaling molecules that bind cytoplasmic or nuclear receptors. Other signaling mechanisms, such as those involving sphingosine phosphate, are not shown here. SUMMARY Mechanical, electrical, or chemical signals are used by cells to communicate. Mechanical signaling and some chemical signaling require close contact. Electrical signaling is the fastest way to move a signal from one part of a cell to another and to adjacent cells. Chemical signals can be used by a cell to regulate itself (autocrine), its near neighbors (paracrine) or distant cells through the medium of the blood (endocrine). Nerve cells use electrical signals over long cell processes to cause release of chemical signals near their target cells. Neuroendocrine signals are chemical signals released by neurons into the blood. Voltage-gated ion channels make electrical signals possible. The influx of Ca21 ions carries an electrical and a chemical signal because Ca21 binds to specific receptors inside the cell to initiate secretion, to activate enzymes indirectly through CAM kinase or directly through activation of other enzymes, or to activate contraction. LGICs convert chemical into electrical signals. This is used in neurotransmission: an action potential on one cell is converted to an electrical signal on another. Many chemical signals have multiple types of receptors, so that the effect in the postsynaptic cell depends on the chemical released by the presynaptic cell and the receptor expressed by the postsynaptic cell. Many chemical signals bind to receptors on the surface membrane that are linked to heterotrimeric G-proteins (GPCRs). These dissociate upon ligand binding and generally the α subunit activates an amplifying enzyme such as adenylyl cyclase (for Gs mechanisms) or phospholipase C (for Gq mechanisms), which increase the C ell Signal ing formation of 30 ,50 cyclic AMP or IP3, respectively. Other GPCRs inhibit adenylyl cyclase (Gi) or recruit a number of small monomeric G-proteins such as Ras and Rho (G11/12). The βγ subunit can also affect intracellular targets. Increased cytosolic cAMP activates PKA that phosphorylates specific target proteins. IP3 released by Gq-coupled receptors causes Ca21 release from ER stores and activation of CaM kinase. Other chemical signals bind to surface receptors that are catalytic. The four classes are receptor guanylyl cyclase, receptor serine/threonine kinase, receptor tyrosine kinase, and receptor-associated tyrosine kinase. Examples of these signals include insulin and growth hormone. Insulin binding activates an intrinsic tyrosine kinase that phosphorylates insulin receptor substrates that bind phosphatidyl inositol 3 kinase, PI-3K. This forms PIP3, which activates a phosphoinositidedependent protein kinase (PDK). Growth hormone activates a receptor that in turn activates a member of the Janus Kinase family of proteins, which then phosphorylates STAT5 (signal transduction and activation of transcription). The phosphorylated STAT molecule turns on specific genes. Small lipophilic chemical signals such as thyroxine, vitamin D, and the steroid hormones penetrate the cell membrane and bind to receptors either in the cytosol or in the nucleus. These receptors bind to specific regions of DNA called response elements. The receptors recruit a large number of accessory proteins that unravel the DNA and direct the synthesis of mRNA that codes for specific proteins. The phosphorylation state of a set of regulated proteins depends not only on the activity of the protein kinase but also on the activity of the protein phosphatases. Both the kinases and the phosphatases may be regulated to alter the phosphorylation state of cellular proteins. REVIEW QUESTIONS 1. What is an autocrine hormone? Paracrine hormone? Endocrine hormone? 2. What is the fastest way to convey a signal from one part of the body to another? 3. How does an electrical signal on the surface of a cell become a Ca21 signal in its interior? List four distinct ways that Ca21 can affect cell function. 4. What is a ligand-gated ion channel? What is the source of the extracellular ligand? Is the ligand a chemical, electrical, or mechanical signal? 5. What is meant by “G-protein-coupled receptor”? A Gαs mechanism couples to what amplifying enzyme? What is the product of this enzyme? What does this product do? What is a Gαi mechanism? 6. What amplifying enzyme is activated by a Gαq mechanism? What are its products? What do these products do? 7. Name the four classes of catalytic receptors. 8. Why do peptide hormones have receptors facing the extracellular space? 9. How do small lipophilic signals affect cell function? What is meant by “nuclear receptor”? What do these receptors do? What is meant by the term “response element”? 10. How does cAMP alter gene expression? 217 2.9 ATP Production I: Glycolysis Learning Objectives G G G G G G G G G G G G Be able to draw a diagram showing the relationship among glycolysis, tricarboxylic acid cycle, and electron transport chain Explain what is meant in describing ATP as the “energy currency” of the cell Write the empirical formula for glucose List three sources of glucose in the body Define glycogenolysis Define gluconeogenesis Describe how glycogenolysis is regulated in the liver by epinephrine Explain why muscle tissue does not contribute to plasma glucose directly Describe how glucose gets into cells Describe what is meant by substrate-level phosphorylation Explain the function of lactate dehydrogenase during rapid glycolysis Describe how the rate of gluconeogenesis can be increased TAKE A GLOBAL VIEW OF METABOLISM Intermediary metabolism comprises all of the transformations of biological chemicals that allow the cell to produce energy and synthesize materials that make it up. It is a bewildering array of chemicals and their interconnected pathways. Within this, there are processes that are the composite of many of the individual processes. Glycolysis, for example, occupies a special place in the metabolic scheme. We ought to have some appreciation of its place without having to recall all of the transformations that occur within it. The same is true of the citric acid cycle, also known as the Kreb’s cycle or the tricarboxylic acid cycle. This set of metabolic transformations is central to energy production in cells. We ought to understand the role of the metabolic pathways without necessarily knowing all of the individual transformations that occur within them. ENERGY PRODUCTION OCCURS IN THREE STAGES: BREAKDOWN INTO UNITS, FORMATION OF ACETYL COA, AND COMPLETE OXIDATION OF ACETYL COA Figure 2.9.1 shows the overall plan of energy-producing reactions in cells. These occur in three stages. In Stage 1, foodstuffs consisting of proteins, lipids, and carbohydrates are broken down into their constituent subunits. These are the amino acids, simple sugars like glucose, and fatty acids and glycerol. In the second stage, these simple subunits are broken down to form acetyl coenzyme A. Coenzyme A is a chemical that acts as a carrier for the two-carbon acetyl group, but it is not a carrier in the sense of being transported across a membrane. It is being carried forward in a biochemical reaction. The formation of acetyl CoA is accompanied by the incorporation of some of the energy of the food into ATP, and some limited formation of another compound, NADH. NADH is nicotinamide adenine dinucleotide. It acts as a carrier for reducing equivalents. We will learn more about NADH later on in this chapter. These reducing equivalents are later used to produce ATP in the mitochondria. The third stage of energy production takes place in the mitochondria and involves the complete oxidation of acetyl CoA to water and CO2 and produces the major proportion of reducing equivalents. The energy stored in NADH produced in this stage is converted to energy stored in ATP via the electron transport chain which is coupled indirectly to the ATP synthetase in the inner mitochondrial membrane. ATP IS THE ENERGY CURRENCY OF THE CELL Electricity is a very versatile form of energy that has come to dominate modern society. We use it to operate heavy machinery, melt metal for casting or extrusion, power drills, pumps, and saws; run television, toasters, ovens, and computers—we use it for almost everything. We generate this electric power by burning coal, natural gas, and even public refuse, but we can also “burn” nuclear material. These methods 218 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00020-3 ATP P roduction I: G lycol ysis Food Proteins Amino acids Fats Polysaccharides Stage 1: Breakdown of foodstuffs to simple units Fatty acids, glycerol Simple sugars ADP + Pi Glycolysis ATP NADH Stage 2: Breakdown of simple units to acetyl CoA with limited production of ATP and NADH Pyruvate NADH CO2 Acetyl CoA Citric acid cycle CO2 Stage 3: Complete oxidation of acetyl CoA with the production of lots of NADH and ATP Reducing power: NADH ADP + Pi Electron transport chain (ETC) ATP synthetase ATP O2 H2O NH3, urea CO2 Waste products FIGURE 2.9.1 Overall scheme of intermediary metabolism. In the first stage, macronutrients found in food are broken down into their constituent subunits. In the second stage, these are converted to acetyl CoA in a process that produces only a little ATP and NADH. In the third stage, the acetyl CoA is completely oxidized, accompanied by the production of lots of ATP and NADH. generate power by boiling water to turn a turbine connected to a dynamo. We can also generate electric power by turning a dynamo by moving water or wind. We can also use solar radiation to generate useful electrical power. Our cells have an analogue of the power plant: the mitochondrion. It does not make electric power, but it does generate chemical energy in the form of ATP. Just like electric power, ATP can be generated from multiple kinds of fuel. Carbohydrates, fats, or amino acids can all be “burned” to produce energy that is stored in the terminal phosphate bond of ATP. Just like societal production of electrical energy, ATP formation has a final common pathway in the mitochondria, the “power house of the cell.” Analogous to electricity, ATP can also be produced outside of the mitochondria. Like electricity, this form of chemical energy is very versatile. ATP fuels chemical work such as the synthesis of materials. It fuels mechanical work such as muscle contraction and movement of the cytoskeleton. It fuels electrical work in moving ions across membranes. The extra energy not directly captured by these processes is used to heat the body. All of these activities require ATP to be split into ADP and inorganic phosphate, Pi. The human body continuously splits ATP, and the steady state requires that this continuous splitting is matched to a continuous reformation of ATP from ADP and Pi. This idea is shown in Figure 2.9.2. FUEL RESERVES ARE STORED IN THE BODY PRIMARILY IN FAT DEPOTS AND GLYCOGEN Energy that the body uses for movement, biochemical synthesis, and transport all ultimately derives from chemical energy stored in food. However, the body stores some of this energy in its own materials. These include the fat deposits in adipose tissue and glycogen granules stored in the muscles and liver. Energy is not stored as protein deposits, but body proteins can be 219 220 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Food Proteins Polysaccharides Amino acids Simple sugars Glycolysis Fats Fatty acids, glycerol ADP + Pi ADP + Pi ATP NADH ATP Chemical work Electrical work Mechanical work Heat Pyruvate NADH CO2 Acetyl CoA Citric acid cycle CO2 Reducing power: NADH ADP + Pi Electron transport chain (ETC) ATP synthetase ATP O2 Chemical work Electrical work Mechanical work Heat H2O NH3, urea CO2 Waste products FIGURE 2.9.2 ATP as the energy currency of the cell. ATP is continuously being used for a variety of purposes that include chemical synthesis, production of mechanical force and transport of materials, and movement of ions that constitutes an electrical current. ATP hydrolysis also generates heat. This continuous use of ATP varies with the state of activation. ATP hydrolysis is coupled to resynthesis of ATP in order to maintain a constant supply of energy so that activation that is coupled to increased rates of ATP hydrolysis is simultaneously linked to increased rates of ATP synthesis. and are continuously used as energy sources. We begin our discussion of energy metabolism with glucose. H HOH2C6 HO 1 5 GLUCOSE IS A READILY AVAILABLE SOURCE OF ENERGY Many cells of the body, particularly those in the central nervous system, depend crucially on glucose as an energy source. Glucose is a six-carbon compound with the empirical formula of C6H12O6. It is called a carbohydrate because its chemical formula is close to Cn(H2O)n, indicating a 1:1 ratio between carbon and water. Thus its empirical formula is equivalent to a hydrated carbon atom. The chemical structure of glucose is shown in Figure 2.9.3. The blood plasma O H 4 3 H OH 2 H OH OH α-D glucopyranose FIGURE 2.9.3 Chemical structure of α-D-glucopyranose. Glucose can exist in several configurations, one of which is shown here. The glucose atoms within the molecule are numbered 1 through 6 as shown in the figure. The pyranose ring forms a six-membered structure that approximates a plane. The hydroxyl side groups project from the plane either up or down. At C-1, C-2, and C-4 it is down and at C-3 it is up. When the hydroxyl group is down it is designated as α; when it is up it is designated as β. ATP P roduction I: G lycol ysis typically contains glucose at levels between 80 and 120 mg%. Recall that mg% is mg of glucose per 100 mL of plasma (51 dL). Glucose enters the circulation from several sources. The first source is directly from foodstuffs. Plant starches in the food we eat are broken down to glucose which is absorbed from the intestine into the portal blood (blood that flows from intestine to liver) and then into the general circulation. Another source of glucose is from glycogen stored in the liver and in muscle. Glycogen is a polymer of glucose in which the glucose subunits are stuck together end to end. There are two ways of doing this, called an α-1,4 glucosidic bond and an α-1,6 glucosidic bond. This nomenclature merely names the numbers of the carbon atoms that are attached to one another and the α signifies the stereochemistry of how the bond is formed. The chemical structure of glycogen is shown in Figure 2.9.4. Glucose is stored as glycogen in many cells, but in large quantities in the muscles and liver. The glucose in glycogen cannot release its chemical energy while it is bound in the glycogen. It must first be broken down to the constituent subunits, the glucose molecules, by a process called glycogenolysis. The root word “lysis” means “break down,” so glycogenolysis means “glycogen break down.” Liver glycogen can contribute to blood glucose, whereas muscle glycogen is converted to glucose in the muscle fiber and used only for muscle activities. A third source of glucose is from amino acids. Some amino acids can be used to produce glucose through a process called gluconeogenesis. Literally, this means “new glucose formation.” HO HO HO 4 6 5 1 HO O HO 4 HO HO HO α–1,4-bond O 3 2 O HO HO O HO O 1 OH O HO HO O OH O HO 4 5 α–1,6-bond O 3 2 1 OH O 6 O OH O HO HO O OH O FIGURE 2.9.4 Structure of glycogen. Note that glycogen is a branched polymer of glucose. The α-1,4 glucosidic bond connects linear chains of glucose molecules. The α-1,6 glucosidic bond causes the chain to branch. GLUCOSE RELEASE BY THE LIVER IS CONTROLLED BY HORMONES THROUGH A SECOND MESSENGER SYSTEM Glycogenolysis in the liver is controlled partly by hormones. A hormone is a material which is released from secretory cells in the body that travels through the body via the blood, and has an effect on target cells located some distance away (see Chapter 2.8). One of the important hormones regulating glycogenolysis in the liver is epinephrine. Epinephrine does not enter the liver cell. It binds to a receptor on the hepatocyte (liver cell) surface and a “second messenger” is produced within the cell. The receptor for epinephrine is a G-protein-coupled receptor (GPCR), as discussed in Chapter 2.8. The receptor is coupled to a heterotrimeric G-protein, a class of protein that binds GTP, guanosine triphosphate. In the case of the epinephrine receptor, the G protein is a Gαs, meaning that the α subunit of the heterotrimeric G-protein stimulates adenylyl cyclase to increase the cytosolic concentration of cyclic AMP (30 ,50 cyclic adenosine monophosphate). The cAMP is the “second messenger” within the hepatocyte. The cAMP then activates an enzyme, protein kinase A (PKA), in the liver cell. PKA begins a cascade of phosphorylation reactions that shuts down glycogen synthesis and activates glycogen breakdown according to the scheme shown in Figure 2.9.5. After activation by cAMP, the system returns to its inactivated state in two ways. First, the cAMP produced by adenylyl cyclase is degraded to AMP (adenosine monophosphate) by another enzyme, cAMP phosphodiesterase. This turns off the second messenger signal. Second, protein phosphatases dephosphorylate the proteins that were phosphorylated during activation of the cascade. There are four classes of serine/threonine phosphoprotein phosphatases: PP1, PP2a, PP2b, and PP2c. PP1 dephosphorylates many of the proteins phosphorylated by PKA. The balance between phosphorylated and dephosphorylated proteins is set by the competing activity of the kinases and phosphatases. THE LIVER EXPORTS GLUCOSE INTO THE BLOOD BECAUSE IT CAN DEPHOSPHORYLATE GLUCOSE-6-P In the liver, glycogenolysis ends at glucose-1-P. This is converted to glucose-6-P by another enzyme, phosphoglucomutase. The glucose-6-P is then converted to glucose by glucose-6-phosphatase. This enzyme is extremely important because only the liver, kidney, and intestine have it, allowing them to release glucose into the blood from glucose-6-P; neither glucose-1-P nor glucose-6-P can exit the cell. Muscle cells have glycogen stores that can be broken down to provide energy, but only for the muscle cell because they lack 221 222 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION cAMP + Protein kinase A ATP ADP Phosphorylase kinase (inactive) Phosphorylase kinase-P (active) Phosphorylase -P (active) Phosphorylase (inactive) FIGURE 2.9.5 Cascade of activation events to shut down glycogen synthesis and activate glycogenolysis upon stimulation of the liver with epinephrine. Epinephrine binds to a G-Protein-Coupled Receptor on the surface of the hepatocytes which stimulates adenylyl cyclase to increase formation of 30 ,50 cyclic adenosine monophosphate (cAMP). The increased cAMP stimulates protein kinase A, which then phosphorylates the enzyme phosphorylase kinase, so-named because it phosphorylates another enzyme, phosphorylase. Phosphorylase has its name because it phosphorylates glycogen during glycogenolysis to produce glucose-1phosphate. PKA also phosphorylates glycogen synthase, converting it from its active form to an inactive form. glucose-6-phosphatase. Muscles cannot contribute glucose to the blood. A SPECIFIC GLUCOSE CARRIER TAKES GLUCOSE UP INTO CELLS In muscle cells, glucose can be taken up from the blood by a glucose transporter, GLUT, of which there are multiple isoforms. The one in muscle and fat is GLUT4. The number of these receptors is regulated hormonally, and they exist in a latent form in vesicles stored within the cell. The GLUT4 transporters are particularly sensitive to the hormone insulin. Brain, liver, and red blood cells have GLUT transporters that are not regulated by insulin, and therefore these tissues are insensitive to insulin. Muscle cells can also derive glucose from glycogenolysis within the cell. The fate of glucose, whether derived from blood or glycogen, is conversion to pyruvate through the process of glycolysis. GLYCOLYSIS IS A SERIES OF BIOCHEMICAL TRANSFORMATIONS LEADING FROM GLUCOSE TO PYRUVATE Figures 2.9.6 and 2.9.7 show the reactions of glycolysis that produce pyruvate from glucose. These reactions occur in the cytoplasm. The pyruvate then enters the mitochondria where it is completely oxidized and produces a number of ATP molecules per molecule of pyruvate. Glycogen Glycogen synthase-P (inactive) Glucose-1-P Glycogen synthase (active) ATP ADP Protein kinase A + cAMP GLYCOLYSIS GENERATES ATP QUICKLY IN THE ABSENCE OF OXYGEN Glycolysis can generate ATP in the absence of oxygen. This is described as anaerobic metabolism. It results from substrate-level phosphorylation. This is distinct from oxidative phosphorylation that occurs in the mitochondria. Substrate-level phosphorylation refers to the formation of ATP from ADP and a phosphorylated intermediate, rather than from ADP and inorganic phosphate, Pi, as is done in oxidative phosphorylation. The amount of ATP that is generated by glycolysis is relatively low. Two ATP molecules are required to start glycolysis (from glucose), and four are generated by substrate-level phosphorylation. An additional two NADH molecules are generated, which can be used to generate another three to five ATP molecules through the electron transport chain in the mitochondria. So a net gain of 57 moles of ATP can be generated from the conversion of 1 mole of glucose to 2 moles of pyruvate. The total energy in the oxidation of glucose is 2867 kJ mol21. The energy in 7 moles of ATP is about 7 3 57.1 kJ mol21 5 399.7 kJ mol21. This represents capture of only some 14% of the total energy available from glucose oxidation. GLYCOLYSIS REQUIRES NAD 1 Glycolysis occurs in the cytoplasm and it generates some NADH from NAD1. The NAD1 is an obligatory substrate for the reaction of glyceraldehyde-3phosphate to 1,3-diphosphoglycerate. If NAD1 is not ATP P roduction I: G lycol ysis CH2OH OH O Glucose OH OH OH CH2OH OH O ATP Glucokinase (liver) Hexokinase (muscle) ADP CH2OPO3 Phosphoglucomutase O OH Glucose-6-phosphate Glycogen OH OH OPO3 OH OH OH Glucose-1-phosphate Phosphoglucose isomerase CH2OPO3 OH O OH Fructose-6-phosphate OH Pi Fructose-1,6-diphosphatase H2O CH2OH ATP Phosphofructokinase ADP CH2OPO3 OH O OH OH Fructose-1,6-diphosphate CH2OPO3 Glyceraldehyde-3-phosphate Aldolase Dihydroxyacetone phosphate CH2OPO3 HC O C HC OH O CH2OH Triose phosphate isomerase CH2OPO3 FIGURE 2.9.6 First part of glycolysis, leading from glucose to two three-carbon intermediates that are readily interconvertible. Chemical structures and names of the intermediates are shown in black. The enzymes that participate in the interconversions are shown in blue. Glycolysis begins by phosphorylation of glucose in two successive steps, forming glucose-6-phosphate and then forming fructose-1,6-diphosphate. These steps in glycolysis require ATP to “prime” the process. regenerated, glycolysis will halt. In the presence of oxygen, NADH is oxidized in the mitochondria to regenerate NAD1, but NADH itself cannot cross the mitochondrial membrane. Two shuttles transfer the “reducing equivalents” across the mitochondrial membrane. These are the glycerolphosphate shuttle and the malate/aspartate shuttle (see Chapter 2.10). Figure 2.9.8 illustrates the requirement of glycolysis for NAD1. If the glycolytic generation of NADH exceeds the mitochondrial oxidation of cytoplasmic NADH, then cytoplasmic NAD1 will become depleted and its absence will limit the metabolic flux through glycolysis. Under these conditions, the cell must regenerate NAD1 from NADH in order to allow glycolysis to continue. This is achieved by making lactic acid from pyruvate through the enzyme lactate dehydrogenase, LDH. Lactic acid production occurs all the time, but increases when glycolysis is going faster than the mitochondria can accommodate the metabolic flux of cytoplasmic NADH, regardless of the state of oxygenation of the tissue. A good example of this occurring physiologically is in muscle during brief strenuous exercise such as a 200-m sprint. In this case, nearly all of the energy will be supplied by glycolysis. In order for glycolysis to continue, the muscle will produce lactic acid, which will leave the muscle and travel to the liver. The oxygen necessary to oxidize the accumulated lactic acid constitutes part of the “oxygen debt” that must be repaid when oxygen is available. Provided the liver is adequately oxygenated, the liver will reoxidize the lactic acid to pyruvate, which can then 223 224 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION HC O HC OH Glyceraldehyde-3-phosphate is converted to muscle glucose is called the Cori cycle (see Chapter 3.7). CH2OPO3 NAD+ Pi Glyceraldehyde-3-phosphate dehydrogenase NADH + H+ OPO3 C O 1,3-Diphosphoglycerate OH HC Substrate-level phosphorylation CH2OPO3 ADP Phosphoglycerate kinase ATP – C O O HC 3-Phosphoglycerate OH CH2OPO3 Phosphoglycerate mutase – O C O OPO3 HC 2-Phosphoglycerate CH2OH Enolase O Substrate-level phosphorylation – C O C OPO3 Phosphoenolpyruvate CH2 ADP Pyruvate kinase ATP O – C O C O Pyruvate CH3 FIGURE 2.9.7 Second part of glycolysis leading from glyceraldehyde-3-P to pyruvate. Chemical structures and names of the intermediates are shown in black. The enzymes that participate in the interconversions are shown in blue. ATP is formed twice in this sequence, once in the conversion of 1,3-diphosphoglycerate to 3-phosphoglycerate and for a second time in the conversion of phosphoenolpyruvate to pyruvate. The formation of ATP directly from phosphorylated intermediary metabolites is called substrate-level phosphorylation. Two molecules of glyceraldehyde-3-phosphate are formed from every molecule of glucose. Thus glycolysis produces 4 ATP per molecule of glucose. be converted to glucose by gluconeogenesis. The glucose so formed can be released by the liver into the blood for use again by the muscle. The overall process by which muscle glucose becomes blood lactic acid which GLUCONEOGENESIS REQUIRES REVERSAL OF GLYCOLYSIS Energy transduction in cells involves glycolysis, as we have reviewed it, plus the complete oxidation of pyruvate in the mitochondria, plus the oxidation of other fuels such as fats and proteins. Some tissues (liver, intestine, kidney) export glucose into the blood for the muscles to use during exercise. As mentioned earlier, the liver can mobilize glycogen stores for this purpose, but it can also make new glucose from the amino acids derived from proteins. The process of making new glucose from proteins is called gluconeogenesis. It involves chemically transforming the hydrocarbon parts of amino acids into intermediates of carbohydrate metabolism, and then running glycolysis backwards to form glucose. How this is accomplished is illustrated in Figure 2.9.9 for the effect of glucagon on liver cells. Briefly, glucagon activates glycogenolysis through means similar to what we have described earlier for epinephrine. This produces glucose-1-phosphate. In the liver, phosphoglucomutase converts glucose-1-phosphate to glucose-6-phosphate. Glucose-6-phosphatase removes the phosphate from glucose-6-phosphate to produce glucose, which is then released into the blood stream. Activated PKA also phosphorylates CREB, the cyclic AMP responsive element binding protein. This activates its binding to the CRE, cAMP responsive element. Activation of CRE increases the transcription of another transcriptional activator that then turns on the synthesis of PEPCK, phosphoenolpyruvate carboxy kinase. This enzyme converts oxaloacetate to phosphoenolpyruvate. The oxaloacetate is a common carbohydrate intermediate formed from the glucogenic amino acids. These are amino acids that form glucose (see Chapter 2.11). PKA also indirectly regulates a key controlling enzyme in glycolysis: phosphofructokinase 1/fructose biphosphatase 1(PFK1/FBPase1). PFK converts fructose-6-phosphate to fructose-1,6-biphosphate; FBPase converts fructose-1,6-biphosphate to fructose-6phosphate. The FBPase1 activity and PFK1 activities are regulated by cytosolic levels of fructose-2,6-biphosphate (FBP). Fructose-2,6-biphosphate stimulates PFK activity and it inhibits FBPase activity. Fructose-2,6biphosphate levels are determined by the activity of phosphofructose kinase 2 and fructose-2,6-biphosphatase (FBPase2) which convert fructose-6phosphate to fructose-2,6-biphosphate. The activities of PFK2/FBPase 2 reside on a single polypeptide chain. PKA phosphorylates PFK2/FBPase2, stimulating the FBPase2 activity and inhibiting the PFK2 activity. This reduces the level of fructose-2,6-biphosphate, which subsequently removes activation of PFK1 and removes inhibition of FBPase1. The net result is an inhibition of PFK1, which thereby slows glycolysis, and activation of FBPase1, which increases gluconeogenesis. ATP P roduction I: G lycol ysis Glucose ATP Glucose-6-P ATP Fructose-6-P Fructose-1,6-diP 2 NAD+ 2 NADH 2 Glyceraldehyde-3-P 2 1,3-Diphosphoglycerate 2 3-Phosphoglycerate 2 ATP 2 2-Phosphoglycerate 2 NAD+ 2 NADH 2 Lactate 2 Phosphoenolpyruvate 2 Pyruvate 2 ATP Lactate dehydrogenase 1 FIGURE 2.9.8 Necessity for regenerating NAD during rapid glycolysis. When NADH oxidation by the mitochondria cannot keep pace with glycolysis, [NAD1] falls and [NADH] rises. The oxidation of NADH by lactate dehydrogenase, converting pyruvate to lactate, occurs to regenerate NAD1 so that glycolysis can continue to generate some ATP. SUMMARY Cells use chemical energy to power their synthetic, mechanical, and transport work. The chemical energy stored in the terminal phosphate bond of ATP is used as a common energy source for all of these processes. Cells produce ATP by linking the energy of oxidation of foodstuffs to the chemical synthesis of ATP. Oxidation of carbohydrates, fats, and proteins all give rise to ATP as a common energy currency for the cell. The overall process of energy production occurs in three stages: (1) breakdown of foodstuffs into component units (amino acids for the proteins; fatty acids and glycerol for fats; glucose and fructose for carbohydrates); (2) formation of acetyl CoA with limited formation of ATP and NADH; (3) complete oxidation of acetyl CoA with the production of lots of NADH and ATP through the electron transport chain in the mitochondria. Carbohydrates provide the most rapid source of ATP. Glucose in the blood can be taken up by tissues through specific glucose transporters in their cell membranes (GLUT1, GLUT4) to provide a ready source of energy. Liver and muscle cells store carbohydrates in a readily usable form called glycogen. Liver can convert glycogen stores to blood glucose but muscle uses its glycogen stores inside the muscle by converting it to glucose-1-phosphate. Glycogen utilization begins with its breakdown into component glucose molecules, a process called glycogenolysis. In liver cells this is regulated by hormones. One important hormone, epinephrine, helps raise blood glucose by mobilizing liver glycogen. It achieves this task by binding to a receptor on the outside surface of hepatocytes. This receptor is coupled to a G-protein, so-named because it binds and then hydrolyzes guanosine triphosphate, GTP. This G-protein consists of three subunits: α, β, and γ. Upon binding of hormone, the α subunit dissociates and activates adenylyl cyclase, which converts intracellular ATP to 30 ,50 -cyclic AMP (cAMP). The cAMP then activates protein kinase A (PKA), which phosphorylates a number of target proteins involved in glycogen metabolism. PKA phosphorylates phosphorylase kinase, which then phosphorylates phosphorylase, the enzyme that breaks down glycogen. It also phosphorylates glycogen synthase, inactivating it. In this way, increasing cAMP turns on glycogenolysis and inhibits glycogen synthesis. The end product of glycogenolysis is glucose-1phosphate. This is converted to glucose-6-phosphate by phosphoglucomutase. Glucose-6-phosphate can then enter glycolysis, the conversion of glucose to pyruvic acid that occurs in the cytoplasm. Liver cells can convert 225 226 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Active PKA Active PKA + + Glycolysis Glucose-6-phosphatase Glucose-1-Pi Glucose Glucose-6-Pi Plasma glucose Glycogen Hexokinase + Active PKA FBPase2 Fructose-6-P ATP PFK2 Fructose-2,6 P2 PFK2 CREB mRNA − − Fructose-2,6 P2 –Pi RNA Polymerase II + FBPase2 + PFK1 FBPase1 − Fructose-1,6-diP + -Pi -Pi Active PKA Active PKA Bind to CRE + Nucleus PEPCK Gluconeogenesis 2 Phosphoenolpyruvate Oxaloacetate Ribosome 2 Pyruvate Malate Amino acids FIGURE 2.9.9 Mechanism of action of glucagon on liver cells to put glucose into the blood. Glucagon increases key processes, indicated by the circled 1 signs, through increasing cAMP in the hepatocytes. Gluconeogenesis is the synthesis of new glucose from amino acids, indicated by the pathway highlighted in blue. Gluconeogenesis requires the conversion of fructose-1,6 diphosphate to fructose-6-phosphate, the reverse of the reaction that occurs during glycolysis. This is accomplished by inhibiting PFK1 and activating FBPase. Inhibition of PFK1 and activation of FBPase are brought about by decreasing fructose-2,6 diphosphate levels by stimulating FBPase2 and inhibiting PFK2 through phosphorylation mediated by PKA. glucose-6-phosphate to glucose, which it then exports into the blood. Muscle cells lack glucose-6-phosphatase and so cannot export glucose into the blood. is produced to regenerate NAD1 so that glycolysis can continue. The first stage in glucose oxidation is glycolysis, in which one molecule of glucose is converted to two molecules of pyruvic acid. Glycolysis generates some ATP by substrate-level phosphorylation (occurring at the level of the phosphorylated intermediates of glycolysis as opposed to synthesis from ADP and Pi). It also requires another compound, nicotinamide adenine dinucleotide (NAD1). This compound is converted to NADH during glycolysis, and under aerobic conditions is regenerated from NADH by the mitochondria. When glycolysis outstrips the ability of mitochondria to regenerate NAD1, NADH can be converted to NAD1 by linking this conversion to the production of lactic acid from pyruvate through the enzyme lactic dehydrogenase. Thus in strenuous activity lactic acid REVIEW QUESTIONS 1. What is glucose? Name three processes that supply body cells with glucose. 2. What is glycogen? What is glycolysis? Why can liver convert glycogen to blood glucose but muscle cannot? 3. Why is ATP first consumed in glycolysis instead of being produced? What is substrate-level phosphorylation? 4. Why do muscle cells produce lactic acid during bursts of activity? 5. What is lactate dehydrogenase? 6. What is the Cori cycle? 7. What is gluconeogenesis? ATP Production II: The TCA Cycle and Oxidative Phosphorylation Learning Objectives G G G G G G G G G G G G G G G G G Describe the reaction catalyzed by pyruvate dehydrogenase Describe in general terms the function of the water-soluble vitamins Describe what it means to say that NADH is a carrier for reducing equivalents Define reduction potential and describe how it can be measured Be able to calculate the energy released in a reduction oxidation reaction List the number of NADH molecules generated per turn of the TCA cycle List the number of FADH2 molecules generated per turn of the TCA cycle List the number of ATP molecules (or equivalent) produced by substrate-level phosphorylation per turn of the TCA cycle Indicate where CO2 is released during glucose oxidation Describe what is meant by the “electron transport chain” Tell the approximate magnitude and sign of the membrane potential across the inner mitochondrial membrane Give the stoichiometry of ATP formation from NADH; from FADH2 Be able to calculate the electrochemical potential difference for H1 ions across the inner mitochondrial membrane Describe in words how the ATP synthase makes ATP Describe the chemiosmotic hypothesis for oxidative phosphorylation Describe in general terms how cytoplasmic NADH enters the mitochondria Describe how ADP and Pi get into the mitochondrion and how ATP leaves it OXIDATION OF PYRUVATE OCCURS IN THE MITOCHONDRIA VIA THE TCA CYCLE Pyruvate is the end product of glycolysis. Its metabolism continues in the mitochondria via the “TCA cycle,” the tricarboxylic acid cycle, so named because many of the intermediates have three carboxyl groups. It is also 2.10 referred to as the Krebs cycle in honor of Sir Hans Krebs, who did much of the pioneering work in describing it, and it is also referred to as the citric acid cycle because citric acid is formed in it. This series of metabolic transformations occurs in the inner mitochondria of cells. Its fuel source is pyruvic acid derived from glycolysis in the cytosol. The TCA cycle can also be initiated within the mitochondria by the oxidation of fatty acids to form acetyl CoA (see Chapter 2.11). PYRUVATE ENTERS THE MITOCHONDRIA AND IS CONVERTED TO ACETYL CoA The mitochondria have two membranes, an outer membrane and an inner membrane. The outer membrane is relatively permeable, whereas the inner membrane is highly impermeable to most materials. Pyruvate produced in the cytosol by glycolysis crosses the inner mitochondrial membrane by facilitated diffusion on its own pyruvate carrier. Inside the matrix of the mitochondria, pyruvate is converted to acetyl coenzyme A. This conversion of pyruvate to acetyl CoA requires three different enzymes and five different coenzymes, which are organized into a multienzyme complex called pyruvate dehydrogenase. Three of the coenzymes required here are vitamins: thiamine, riboflavin, and niacin. The water-soluble vitamins all find their use in mammals as part of enzymatic reactions, and most of the B vitamins are involved in carbohydrate metabolism. The overall reaction is shown in Figure 2.10.1 along with the structure of coenzyme A. PYRUVATE DEHYDROGENASE RELEASES CO2 AND MAKES NADH The production of acetyl CoA from pyruvate is noteworthy because here is the first production of CO2 from glucose. This gas forms a major waste product that must be eliminated, largely through the lungs. Second, the reaction produces NADH from NAD1. NADH is a carrier for reducing equivalents in the cell. The structures of NAD1 and NADH are shown in Figure 2.10.2. The conversion of NAD1 to NADH is a reduction reaction. Oxidation and reduction are two halves of the same process, dealing with the exchange of electrons 227 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00021-5 CH3 Pyruvate dehydrogenase NAD+ C NADH + H COO– S C CH2 O Acetyl group O CH2 CH3 Pyruvate CoASH CO2 Thiamine PP Lipoic acid Riboflavin NH C O HC H3C OH C CH3 CH2 O –O P O Coenzyme A O NH2 –O P O N N O CH2 O –O N O OH P O N OH FIGURE 2.10.1 Conversion of pyruvate to acetyl CoA by pyruvate dehydrogenase, indicated in blue. Note that the reaction releases CO2. Coenzyme A is a complex of ATP with pantothenic acid. It carries the acetyl group on through biochemical reactions in the cell. The reaction requires five different coenzymes (coenzyme A, NAD1, thiamine pyrophosphate, lipoic acid, and riboflavin). Three of these coenzymes are vitamins (niacin, thiamine, and riboflavin). Reducing equivalents Niacin H O C O O NH2 O P O– + OH OH O– N P O O CH2 N O NH2 N N O NH2 C CH2 N+ O O H H+ + 2e– O CH2 N O P O– O OH OH O– N P O O CH2 N O OH OH OH OH NAD+ Nicotinamide adenine dinucleotide NADH Oxidized Reduced FIGURE 2.10.2 Formation of NADH from NAD1. NAD1 is the oxidized form; NADH is the reduced form. NH2 N N AT P P roduc ti on II : T h e T CA Cy c le a nd O xidat ive P hos phor yla tion between chemicals. A mnemonic device for oxidation/ reduction reactions is LEO GER which stands for “Loss of Electrons, Oxidation; Gain of Electrons, Reduction.” When NAD1 gains two electrons in the form of H2, it is reduced. At the same time, the chemical from which the H2 is extracted is oxidized. In oxidation/reduction reactions, one chemical is reduced while the other is oxidized. Now the electrons in the reduction reaction had to come from someplace, and that someplace is another chemical. So in a reductionoxidation reaction there are always two redox pairs, one being reduced and the other being oxidized in the process. THE AFFINITY OF A CHEMICAL FOR ELECTRONS IS MEASURED BY ITS STANDARD REDUCTION POTENTIAL If compound A binds electrons more tightly than compound B, we expect that A will take electrons from B in the reaction A 1 B-A2 1 B1. In this reaction, A is reduced and B is oxidized. This can be written as the sum of two half-reactions: A 1 e2 -A2 and B-B1 1 e2 The relative tendency for a compound to be reduced is called its reduction potential. The standard redox potential is measured against a standard half-reaction, arbitrarily assigned the reduction potential of zero. This is the reduction/oxidation of hydrogen: 2H1 ðaqÞ 1 2e2 -H2 ðgÞ E0 5 0:00 Here E0 is referred to as the standard reduction potential of hydrogen. Because oxidation is the reverse reaction of reduction, the standard oxidation potential is the negative of the standard reduction potential. Standard reduction potentials are measured as shown in Figure 2.10.3 against a Standard Hydrogen Electrode. It is measured under standard conditions of 1 M concentration of all reactants and 1 atm pressure of H2 gas. The standard reduction potential is given the symbol E0. Recall from Chapter 1.3 that the potential is the work done in bringing a positive unit charge from infinite separation to the point at which the potential is to be defined. The work done in moving a charge through a potential is just the charge times the potential. This was expressed in Eqn [1.3.13]: ½2:10:1 ΔEnergy 5 qΔU where U is the potential and q is the total charge. In electrochemistry we generally use E for the potential, as we have done above, in units of volts, and q is in coulombs. This gives the energy change in volt-coulombs or joules. This energy change is the change in free energy, and so Eqn [2.10.1] can be rewritten as ½2:10:2 ΔG 5 z` ΔE00 Voltmeter E0 > 0 E0 = 0 – H2 gas (1 atm) in Salt bridge Pt black electrode 1M Reactant Test half cell 1M H+ Standard hydrogen electrode FIGURE 2.10.3 Measurement of the standard electrode potential. One half-cell containing standard concentrations of reactant (1 M) is connected to the standard hydrogen electrode (SHE) with a finely divided Pt black electrode bubbled with 1 atm of H2 gas and 1 M H1 in solution. The voltage between the two half-cells is the reduction potential. If electrons flow to the reactant, then the reactant is being reduced and has a positive standard reduction potential—it has a higher affinity for electrons than hydrogen. Note that current is defined as positive charge flow, which is opposite to electron flow. Voltages are typically measured with a potentiometer that finds the voltage necessary to stop current flow, so that the measurement occurs at equilibrium when no current flows. where ΔE00 is the difference in reduction potential between the two half-cells, z is the valence of the carrier, and ` is the Faraday 5 98,500 coulombs mol21. The Faraday converts the charges to coulombs and normalizes the free energy change to the free energy change per mole. Since the charge carrier is the electron, z 5 21, this equation becomes ½2:10:3 ΔG 5 2 `ΔE00 This is the free energy change per mole of electrons. If there are n electrons involved per reaction, the free energy change per mole of reaction is ½2:10:4 ΔG 5 2 n`ΔE00 THE REDUCTION POTENTIAL DEPENDS ON THE CONCENTRATION OF OXIDIZED AND REDUCED FORMS, AND THE TEMPERATURE The standard reduction potential is defined at unit concentrations of all reactants. When the concentrations are not 1 M, the measured reduction potential changes. Consider the reduction of A in contact with a Standard Hydrogen Electrode, as described earlier. We can write the two half-cell reactions as 2 A 1 n H1 test 1 n e -AHn n=2 H2 -n HSHE 1 n e2 1 A 1 n=2 H2 1 n H1 test -AHn 1 n HSHE 229 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION where SHE denotes Standard Hydrogen Electrode. The free energy per mole for the overall reaction is calculated as 0.4 Δμ 5 μ0AHn 1 RT ln½AHn 1 n μ0HSHE 1 n RT ln½H1 SHE n n 2μ0A 2 RT ln½A 2 μ0H2 2 RT ln fH2 2 2 2n μ0Htest 2 n RT ln½H1 test ½2:10:5 where fH2 is the fugacity of hydrogen, analogous to the activity in aqueous solutions. Since the pressure of H2 gas at the Standard Hydrogen Electrode is 1 atm, fH2 is 1.0. Collecting terms in Eqn [2.10.5], we get ½2:10:6 The top line of Eqn [2.10.6] is the free energy per mole for the standard reduction potential for A. This is given by Eqn [2.10.4]. Remembering that [H1] for the Standard Hydrogen Electrode 5 1 M, its term drops out, because ln [H1] 5 ln 1 5 0, and we write ½AHn 2 n RT ln H1 test ½A Δμ 5 2 n`ΔE0 1 RT ln ½2:10:7 The observed potential is related to the free energy change of the reaction through Eqn [2.10.4] and so we have ½AHn 2n`ΔE 5 2n`ΔE0 1 n RT ln H1 test 1 RT ln ½A ½2:10:8 1 In most situations, the [H ] in the test half-cell can be kept nearly constant by the use of chemical buffers. In this case, its contribution to the free energy will also be constant, and we can define a practical reduction potential (ΔE0’) that incorporates the standard reduction potential and the pH term. Doing this, plus dividing both sides by 2 n`, we come to ½2:10:9 ΔE 5 ΔE00 1 RT ½A ln n` ½AHn The argument of the logarithm is inverted from Eqn [2.10.8] because of multiplying through by 21 to convert the minus sign in 2 n` to positive values. What this equation means is that the actual reduction potential depends on the relative concentrations of the oxidized form ([A]) and reduced form ([AHn]). The reduction potential as a function of the oxidation state of a redox reaction is shown in Figure 2.10.4 for NADH, Ubiquinone, and Cytochrome C (more on these later). Note that when [A] 5 [AH], which occurs Cyt Cox + e– CoQred CoQox + 2H+ + 2e– 0.0 –0.2 –0.4 n Δμ 5 μ0AHn 1 n μ0HSHE 2 μ0A2 μ0H2 2 n μ0Htest 1 RT ln½AHn 2 1 2 RT ln½A 1 n RT ln½H1 SHE 2 n RT ln½H test Cyt Cred 0.2 Reduction potential 230 –0.6 0.0 NADH 0.2 NAD+ + H+ + 2e– 0.4 0.6 Fractionof oxidation 0.8 1.0 FIGURE 2.10.4 Reduction potential for various redox reactions found in the cell as a function of their oxidation state. The reduced form is AHn in Eqn [2.10.9] and the oxidized form is A. As AHn is oxidized by addition of a strong oxidant, it is converted to A and the reduction potential changes according to Eqn [2.10.9]. When the reaction is 50% complete, [A] 5 [AHn] and the argument of the logarithm becomes 1.0, and ln 1.0 5 0. At this point, the measured reduction potential is the practical standard reduction potential: ΔE 5 ΔE0'. For NADH this is 20.32 volts; for Coenzyme Q (CoQ) this is 10.030 volts; for cytochrome C this is 10.23 volts. The higher reduction potential means that oxidized CoQ will oxidize NADH by taking electrons from it, and oxidized Cyt C will oxidize CoQ by taking electrons from it. when the reactant is 50% oxidized, the measured ΔE is equal to ΔE00 . THE TCA CYCLE IS A CATALYTIC CYCLE The biochemical transformations that constitute the TCA cycle are shown in Figure 2.10.5. This is a catalytic cycle in that the intermediates themselves are not altered by the cycle. It starts with oxaloacetate, a 4-carbon dicarboxylic acid, condensing with acetyl CoA to produce citrate. As the cycle continues, NADH is generated in each of three separate reactions (at isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and malate dehydrogenase) and FADH2 is generated at succinate dehydrogenase. FADH2 is the oxidized form of FAD, flavin adenine dinucleotide. It is another chemical carrier of reducing equivalents whose structure is shown in Figure 2.10.6. GTP is generated at succinyl CoA synthetase, and CO2 is generated twice, at isocitrate dehydrogenase and α-ketoglutarate dehydrogenase. The GTP generated in the cycle by substrate-level phosphorylation is formally equivalent to ATP as the two high-energy compounds are readily interconverted. The overall TCA cycle is Acetyl CoA 1 2H2 O 1 GDP 1 Pi 1 FAD 1 3NAD1 CoASH 1 2 CO2 1 GTP 1 3 NADH 1 3 H1 1 FADH2 AT P P roduc ti on II : T h e T CA Cy c le a nd O xidat ive P hos phor yla tion Acetyl CoA O H3C Malate HC HO O C COO– H2C OH H2C COO– C COO– H2C COO– Aconitase Citrate synthase CoASH NADH + H COO– Citrate SCoA COO– Oxaloacetate Malate dehydrogenase C H2C COO– HC COO– H C COO– NAD+ HO H2C – COO Isocitrate NAD+ Isocitrate dehydrogenase Fumarase TCA Cycle HC COO– HC COO– CO2 NADH + H Fumarate H2C FADH2 H2C NAD+ O FAD C COO– Alpha ketoglutarate COO– NADH + H Succinate dehydrogenase CoASH H2C COO– – H2C COO H2C COO– CH2 Succinate O GTP C CoASH Alpha ketoglutarate dehydrogenase SCoA CO2 GDP Pi Succinyl CoA Succinyl CoA synthetase Pyruvate dehydrogenase FIGURE 2.10.5 Metabolic interconversions in the TCA cycle. Note that the two-carbon fragment, acetic acid, is carried by CoA to combine with oxaloacetate to form citrate. The oxaloacetate is regenerated by the cycle. Thus the cycle’s intermediates are catalysts in that they are not consumed. The two carbons in acetate are converted to CO2 by the cycle. The result is that the two carbons in acetate are converted to CO2 and a bunch of reducing equivalents (8 per 2-carbon acetate). Note that O2 is not explicitly required for this process, but it is required for the continued operation of this cycle. If we add in the formation of acetyl CoA from pyruvate, the overall reaction is Pyruvate 1 3 H2 O 1 GDP 1 Pi 1 FAD 1 4 NAD1 3 CO2 1 GTP 1 4 NADH 1 4 H1 1 FADH2 The NADH and FADH2 produced during this overall series of reactions must be returned to NAD1 and FAD, or the process will stop. The reduced NADH and FADH2 are oxidized by a special system called the electron transport chain (ETC). In the process of being oxidized, the energy stored in these compounds enables the synthesis of ATP through a process called oxidative phosphorylation. THE ETC LINKS CHEMICAL ENERGY TO H 1 PUMPING OUT OF THE MITOCHONDRIA The ETC consists of an array of proteins inserted in the inner mitochondrial membrane. The overall plan is this: NADH delivers two electrons to a series of chemicals that differ in their chemical affinity for these electrons (see Figure 2.10.7). This is expressed in their reduction potential (see above) which is related to 231 232 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION O O O Ribose O– CH2 HO O P O O– P O N O Adenine CH2 N CH CH2 Flavin NH2 HO CH HO CH N H3C N N CH HO CH HO CH P O – N CH2 H H3C N N H3C N NH NH2 N CH2 N + 2H+ + 2e– O O– P O O OH OH CH2 H3C O HO O O N N N OH OH O NH O Reducing equivalents H O Flavin adenine dinucleotide Oxidized Reduced FIGURE 2.10.6 Structure of oxidized and reduced FAD. FAD stands for “flavin adenine dinucleotide.” It consists of a flavin part, a ribose part, and an adenine part. The ribose 1 flavin is better known as riboflavin or vitamin B2. The binding of reducing equivalents is associated with gain of energy, which can be released on oxidation. Electron transport through complexes I, III, and IV is linked to H+ pumping from the matrix to the intermembrane space Outer mitochondrial membrane H+ H+ H+ Complex I 2e– Cyt C CoQ 2e– Inner mitochondrial membrane CoQ 2e– NADH Complex IV Complex III Complex II NAD+ H+ Intermembrane space 2e– FADH2 FAD H+ H+ Mitochondrial matrix NADH feeds into the beginning of the ETC— three complexes pump H+ 1/2 O2 + 2H+ FADH2 feeds into the ETC after complex I; only two complexes pump H+ H2O The final electron acceptor is oxygen; without it, the whole ETC backs up FIGURE 2.10.7 The electron transport chain (ETC). NADH feeds in reducing equivalents at the beginning of the ETC, which hands them on to proteins with progressively higher affinity until at the end of the chain the electrons are combined with oxygen. Complexes I, III, and IV use the chemical energy of oxidation to pump H1 ions from the mitochondrial matrix to the intermembrane space. This makes an electrical current that separates charge and produces a potential difference across the mitochondrial membrane. their free energy. The energy is released gradually, in steps, and the ETC complexes use the decrease in free energy to pump hydrogen ions from the matrix space to the intermembrane space between the inner and outer mitochondrial membranes. This pumping of hydrogen ions produces an electrochemical gradient for hydrogen ions and the energy in this gradient is used to generate ATP from ADP and Pi. AT P P roduc ti on II : T h e T CA Cy c le a nd O xidat ive P hos phor yla tion OXYGEN ACCEPTS ELECTRONS AT THE END OF THE ETC Molecular oxygen oxidizes the last step in the ETC. This is the point at which oxygen is consumed by the mitochondria, producing water. Without oxygen to finally oxidize the ETC, the chain itself will remain reduced and no further reducing agents can be fed through it. That is, NADH cannot be converted to NAD1 (at the far left of the chain) in the absence of oxygen. FADH2 also cannot be converted to FAD in the absence of oxygen. The ΔE0’ for the redox reaction 1/2O2 1 2H1 1 2e2H2O is 10.82 volts. The cascade of electrons down the ETC is shown diagrammatically in Figure 2.10.8. Note that the electrons travel down the cascade towards ever more positive reduction potentials, as these signify increasing affinity for electrons. PROTON PUMPING AND ELECTRON TRANSPORT ARE TIGHTLY COUPLED If the proton gradient is at equilibrium with the free energy of electron transport, then the electrons cannot be transported through the ETC. The energy stored in the electrochemical gradient of protons across the inner mitochondrial membrane must also be drained in some way for the ETC to continue operating. The energy in the proton electrochemical gradient is used to make ATP. The coupling of the electrochemical gradient of H1 across the inner mitochondrial membrane with ATP synthesis is called chemiosmotic coupling (because –0.5 NADH –0.4 –0.3 THE ATP SYNTHASE COUPLES INWARD H 1 FLUX TO ATP SYNTHESIS The inner mitochondrial membrane contains many copies of a protein called the F0F1ATPase. This is also called ATP synthase. It consists of two parts: the F0 component spans the membrane and provides a channel for protons to move into the matrix from the intermembrane space. The F1 component is a complex of five proteins with the composition α3β3γδε, with a molecular weight of about 360,000. The F0 Part of the complex consists of an integral membrane, a subunit, a b dimer, and 815 small c-subunits. The structure of the ATP synthase is shown in Figure 2.10.9. This remarkable complex couples the movement of H1 to the synthesis of ATP through mechanical intermediates. Hydrogen ions from the inner matrix access the c subunit via a channel in the a subunit, causing a rotation of the c subunit turbine. This rotates the γ subunit, which has a cam-like protrusion that deforms the α and β subunits. Each time the cam passes one of the three αβ complexes, ATP is formed from ADP and Pi bound to the αβ subunits. Because there are three of these αβ subunits, each turn of the c-protein turbine produces 3 ATP molecules. NAD+ 4H+ –0.320 volts NADH dehydrogenase complex –0.2 Reduction potential (E0', volts) there is a concentration difference across the membrane and an electric potential). It was first proposed by Peter Mitchell in 1961, who was awarded the Nobel Prize for the work in 1978. 2H+ –0.1 0.0 0.030 volts CoQ Cytochrome C reductase complex 0.1 0.2 0.230 volts Cyt C 4H+ 0.3 0.4 0.5 Cytochrome C oxidase complex 0.6 0.7 0.8 0.820 volts 0.9 2H+ + 1/2 O2 H2O FIGURE 2.10.8 Cascade of electrons in the electron transport chain. It begins with the production of NADH by reactions in glycolysis or TCA cycle. The reduction potential of NADH/NAD1 is 20.320 volts. Electrons are passed to the NADH dehydrogenase complex that pumps 4H1 per 2 e2 out of the mitochondrial matrix to the intermembrane space. Electrons are then passed to Coenzyme Q, with a reduction potential of 0.030 volts. Coenzyme Q passes the electrons to the cytochrome C reductase complex that pumps 2H1 per 2 e2. Electrons then are transferred to cytochrome C with a reduction potential of 0.230 volts. The electrons that are taken by the Cytochrome C oxidase complex that pumps 4H1 per 2 e2. Cytochrome C oxidase is finally oxidized by molecular oxygen, whose reduction potential is 0.820 volts. 233 234 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Binding of H+ to the c unit causes rotation of the c complex, which is connected to the γ subunit H+ c c c a c c + H+ H+ H ε H+ H+ Inner mitochondrial membrane H+ The shaft has a cam-like projection that interacts with the αβ subunits that form ATP from ADP and Pi γ b dimer ATP ADP + Pi β α δ ATP β So if there are 10 C units then 10 H+ cause a complete rotation of the C rotor and one rotation of the γ shaft that makes 3 ATP because there are three αβ subunits on each ATP synthase α α ADP + Pi β ADP + Pi ATP FIGURE 2.10.9 The F0F1ATPase or the ATP synthase of mitochondria. It consists of a complex of proteins that make up a tiny H1-driven turbine. H1 ions enter from the intermembrane space through protein a. They bind to a c unit in the rotator, which cause a rotation of the c-complex. After a nearly complete rotation, the H1 ion is removed to the mitochondrial matrix. The c-complex binds the γ subunit and rotates it. The αβ subunits of the head are kept steady by the stator components, δ and the b dimer. The γ subunit has a projection that interacts with each of the αβ subunits, and this mechanical interaction is used to synthesize ATP from ADP and Pi. Complete rotation of the c-complex requires as many H1 as c-subunits. For each complete rotation, the ATP synthase makes 3 ATP molecules. THE PROTON ELECTROCHEMICAL GRADIENT PROVIDES THE ENERGY FOR ATP SYNTHESIS The ETC pumps H1 ions out of the matrix into the intermembrane space. The stoichiometry is about 10H1 ions per 2e2 when they originate from NADH, and about 6H1 when the 2e2 originate from FADH2 (see Figure 2.10.7). Because the H1 ions move without counter ions, this movement is an outward current that separates charge, and therefore there is a potential developed across the inner mitochondrial membrane. This potential varies depending on the state of mitochondrial activity, but a typical value is about 160 mV, negative inside. In addition to the potential, there is a concentration difference in H1 established across the membrane. The pH of the intermembrane space is about 7.0, whereas the pH of the matrix is about 8.0. Recall that pH 5 2log [H1], so that pH 5 7.0 implies that [H1] 5 1027 M and pH 5 8 means [H1] 5 1028 M. Thus there is a 10-fold difference in the [H1] established by the ETC. When H1 ions travel from the intermembrane space to the matrix, they release the free energy stored in the electrochemical gradient for H1, enabling the F1 subunit to synthesize ATP from ADP and Pi. This proton electrochemical gradient is sometimes called the proton motive force. The free energy for H1 transfer from the intermembrane space to the mitochondrial matrix is calculated as Δμout-in 5 μin 2 μout 5 μ0 1 RT ln½H1 in 1 `ψin 2 μ0 ½2:10:10 2 RT ln½H1 out 2 `ψout 5 RT ln ½H1 in 1 `ðψin 2 ψout Þ ½H1 out The free energy change for ATP synthesis under the conditions of the cell varies from cell to cell and from place to place within the cell because the local concentrations of ADP, Pi, ATP, and ions that bind to them (H1, Ca21, and Mg21) also vary from place to place. Nevertheless, we have already calculated an approximate free energy change for ATP hydrolysis under conditions of the cell to be 257.1 kJ mol21. The free energy of ATP synthesis should be the opposite of this, 157.1 kJ mol21. According to the result in Example 2.10.1, there is not enough energy in one H1 transport to synthesize ATP. If we assume integral stoichiometry, we need at least three of them. The free energy for the reaction ½2:10:11 1 ADP 1 Pi 1 3H1 out -ATP 1 3Hin AT P P roduc ti on II : T h e T CA Cy c le a nd O xidat ive P hos phor yla tion EXAMPLE 2.10.1 Calculate the Free Energy in the Mitochondrial H1 Electrochemical Gradient The free energy per mole of H1 is given by Eqn [2.10.1] as 1 ½H in ΔμHout -Hin 5 RT ln 1 `ðψin ψout Þ ½H1 out 21 21 T 5 310 K, Inserting values of R 5 8.314 J mol K , [H1]in 5 1028 M, [H1]out 51027 M, we calculate the chemical part as ! ! ½H1 in 1028 M 21 21 K 3 310 K 3 ln RT ln 5 8:314 J mol ½H1 out 1027 M Using ` 5 9.649 3 104 C mol21 and ψin 5 20.16 V (with ψout 5 0), the electrical part of the free energy change is `ðψin ψout Þ 5 9:649 3 104 C mol21 3 ð2 0:16 V 0Þ 5 2 15:44 kJ mol21 Thus the total free energy change for H1 transfer from the intermembrane space to the matrix, for this condition given here, is 2 21.37 kJ mol21. 5 2 5:93 kJ mol21 is the sum of the free energy of two processes: ½2:10:12 ADP 1 Pi-ATP 1 3H1 out -3Hin We add the two to get Δμ 5 ΔμADP1Pi-ATP 1 3ΔμH1out -H1in ½2:10:13 5 57:1 kJ mol21 1 3ð2 21:37 kJ mol21 Þ 5 27:01 kJ mol21 The negative free energy change for this coupled reaction indicates that this process will proceed spontaneously. That is, there is enough energy in the electrochemical gradient of H1 across the inner mitochondrial membrane to synthesize 1 ATP for every 3H1 ions transported. As it turns out, the stoichiometry is not integral. NADH FORMS ABOUT 2.5 ATP MOLECULES; FADH2 FORMS ABOUT 1.5 ATP MOLECULES The amount of ATP formed from oxidative phosphorylation has been controversial but a consensus seems to have been reached. Measurements show that electron transport beginning with NADH results in 10H1 ions being transported from matrix to intermembrane space, and with FADH2 the number is 6, because the first complex is bypassed. What happens to these H1 ions? Most are used to drive the ATP synthase as described in Figure 2.10.9, but some are used to bring the phosphate into the mitochondria from the cytosolic compartment (see Figure 2.10.10) and some are used for other transport processes. Mitochondria from the heart of cows has 8 c-subunits in their ATP synthase, suggesting that 8H1 ions are needed for one complete rotation of the rotor and synthesis of 3 ATP molecules. This gives a nonintegral stoichiometry: each ATP requires 8/3 5 2.67H1 ions! Our calculation above indicates that the minimum number is 57.1 kJ mol21/21.87 kJ mol21 5 2.61. Because 1H1 is required to import Pi, the number of ATP produced by NADH becomes 10/(2.67 1 1) 5 2.7 ATP/NADH. If the proton motive force is used to drive other processes, the ATP yield will be lower. Recent studies suggest a relatively constant H1/ATP ratio of 4.0, including transport processes. In this case ATP production from NADH would be 10H1 per NADH/4H1 per ATP 5 2.5 ATP per NADH. For FADH2, the ratio would be 6H1 per FADH2/4H1 per ATP 5 1.5 ATP per FADH2. These numbers are approximate and tentative. The ATPase from different sources has different c-ring sizes that may cause differences in the H1/ATP ratio and therefore the ATP/NADH ratio. ATP CAN BE PRODUCED FROM CYTOSOLIC NADH The NADH produced in the cytosol by glycolysis cannot enter the mitochondrial matrix, yet it must be oxidized back to NAD1 to allow glycolysis to proceed. This can be accomplished by lactate dehydrogenase, as described in Chapter 2.9, but this does not extract the energy of combustion remaining in the lactic acid. Two types of shuttle mechanisms have the effect of bringing cytosolic reducing equivalents into the matrix, without NADH itself actually entering the matrix. These shuttles are the glycerol phosphate shuttle and the malate/aspartate shuttle. In the glycerol phosphate shuttle, NADH is oxidized to NAD1 by the cytosolic glycerol-3-phosphate dehydrogenase, while dihydroxyacetone phosphate is simultaneously reduced to glycerol-3-phosphate. Glycerol-3-phosphate then penetrates the mitochondrial outer membrane and reduces FAD to FADH2 by the mitochondrial glycerol-3-phosphate dehydrogenase to form dihydroxyacetone phosphate. In this way, we start with cytoplasmic NADH and dihydroxyacetone phosphate and we end up with mitochondrial FADH2 and cytoplasmic NAD1 and dihydroxyacetone phosphate. So the reducing power of NADH is transferred to FADH2, which then enters the ETC to generate 1.5 ATP molecules. Because complex I is bypassed, only 1.5 ATP are made per molecule of NADH passed on to the mitochondria by the glycerol P shuttle. Figure 2.10.11 illustrates the glycerol phosphate shuttle. 235 236 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 3 Some additional H+ is used to import Pi and other substrates, such as pyruvate, for the mitochondria. The overall requirement for ATP synthesis, including transport, is about 4H+ per ATP 1 The ETC uses the energy in electron binding to NADH and FADH2 to create a concentration difference in [H+] and a membrane potential across the inner mitochondrial membrane. Three complexes of the ETC pump a total of 10 H+ per NADH or 6H+ per FADH2 4H+ 2H+ 4H+ Outer mitochondrial membrane – 10H+/NADH ADP3ATP4- 1H+/ATP Pyruvate –3 10 –7 + Intermembrane space 0.3H+/ATP ETC –160 mV OH– M = pH 7.0 OH– 2.7H+/ATP PO4 ++++++++ F0 Inner mitochondrial membrane ------------ NADH 10–8 M = pH 8.0 F1 NAD+ 1.3 OH–/ATP ATP Pi + ADP H2O 1.3 H+/ATP Mitochondrial matrix 2.7H+/ATP 2 The ATP synthase uses the energy in the H+ gradient and membrane potential to synthesize ATP from ADP and Pi. About 2.7H+ ions are required for each ATP synthesized FIGURE 2.10.10 Overall coupling of the ETC to the ATP synthase. The ETC pumps electrons from the mitochondrial matrix to the intermembrane space, creating a [H1] concentration gradient and an electrical potential difference. The ATP synthase uses the energy in this gradient to link ATP synthesis to H1 ions going down their electrochemical gradient. Because of its mechanism (see Figure 2.10.9) in heart mitochondria 8H1 ions make 3 ATP molecules, for a stoichiometry of 2.7H1/ATP. However, the H1 gradient is used for mitochondrial transport as well. For each Pi that enters the mitochondria, 1.0H1 ion is used—the exit of OH2 is equivalent to the entry of H1. Small amounts of H1 flow is also used for other transport processes. The total H1 required for ATP synthesis, including transport, is about 4.0. Because each NADH causes ETC to pump 10H1 ions, this means that about 2.5 ATP molecules are formed per NADH. ATP exit and ADP entry do not require energy, as shown. In the malate/aspartate shuttle, cytoplasmic NADH is used to convert cytoplasmic oxaloacetate to malate, which can be carried across the mitochondrial inner membrane by a dicarboxylate carrier by facilitated diffusion with no metabolic energy expenditure. Inside the matrix, the malate is converted back to oxaloacetate, generating the NADH back, which then transfers electrons to the ETC. To complete the cycle, oxaloacetate must get back outside. This is accomplished by converting oxaloacetate to aspartate (using glutamate as a substrate). The aspartate is transported out of the mitochondria where it is converted back to oxaloacetate and glutamate. The cycle is completed when glutamate goes back into the mitochondria. In this shuttle, 2.5 ATP molecules are produced for each cytosolic NADH because it is effectively transferred into the matrix as NADH. Figure 2.10.12 illustrates the malate/aspartate shuttle. MOST OF THE ATP PRODUCED DURING COMPLETE GLUCOSE OXIDATION COMES FROM OXIDATIVE PHOSPHORYLATION Figure 2.10.13 shows the production of ATP throughout glycolysis and the TCA cycle. Glycolysis utilizes 2 moles of ATP per mole of glucose and then produces 4 moles ATP per mole by substrate-level phosphorylation. The 2 moles of NADH produced by glyceraldehyde-3-P dehydrogenase can be converted to either 5 moles of NAD+ NADH + H+ Glycerol-3-P dehydrogenase Cytosolic compartment Dihydroxyacetone P Glycerol-3-P Outer mitochondrial membrane Dihydroxyacetone P Glycerol-3-P H+ Complex I H+ E-FADH2 E-FAD 2e– Inner mitochondrial membrane Complex IV Complex III NAD+ Complex II H+ Intermembrane space 2e–Cyt C CoQ 2e– CoQ NADH H+ 2e– FADH2 FAD H+ H+ Mitochondrial matrix 1/2 O2 + 2H+ H2O FIGURE 2.10.11 The glycerol phosphate shuttle for transferring cytosolic reducing equivalents to the mitochondria. Cytosolic NADH 1 H1 is converted to NAD1 by glycerol P dehydrogenase. Glycerol-3-P crosses the outer mitochondrial membrane and is converted back to dihydroxyacetone P by a mitochondrial glycerol-3-P dehydrogenase in the inner mitochondrial membrane. The dihydroxyacetone P goes back into the cytosol. The reduced mitochondrial glycerol-3-P dehydrogenase reduces ubiquinone in the inner membrane, which passes the reducing equivalents on to complex III of the ETC. NAD+ NADH + H+ Cytosolic compartment Cytosolic malate dehydrogenase Malate Oxaloacetate Aspartate Aspartate aminotransferase α-ketoglutarate Glutamate Outer mitochondrial membrane Malate α-ketoglutarate Glutamate Aspartate Intermembrane space Inner mitochondrial membrane α-ketoglutarate Malate Glutamate Oxaloacetate Aspartate Aspartate aminotransferase Mitochondrial malate dehydrogenase NAD+ NADH + H+ Mitochondrial matrix FIGURE 2.10.12 The malate shuttle. Two cycles running in opposite directions have the net effect of transferring NADH from cytosol to mitochondrial matrix. In one cycle, oxaloacetate is converted to malate while NADH is converted to NAD1. Malate crosses the inner mitochondrial membrane where it is converted back to oxaloacetate and NADH. The oxaloacetate is then converted to aspartate, which leaves the mitochondria and passes back into the cytosol where the aspartate is converted back to oxaloacetate. A second cycle runs in the opposite direction. Malate entry into the mitochondrial matrix is accompanied by α-ketoglutarate exit. In the cytosol, the latter is converted to glutamate by aspartate amino transferase, which links the reaction α-ketoglutarate-glutamate to the reaction aspartate-oxaloacetate. The glutamate exchanges with aspartate across the inner mitochondrial membrane. 238 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Glucose ATP Glucose 6-P Fructose 6-P ATP Fructose 1,6 diP ATP synthase 2 NADH ETC 3-5 ATP 2 Glyceraldehyde 3-P 2 NAD+ 2 1,3 Diphosphoglycerate 2 ATP 2 3-Phosphoglycerate 2 2-Phosphoglycerate 2 Phosphoenol pyruvate ATP synthase 5 ATP ETC 2 ATP 2 Pyruvate 2 NAD+ 2 NADH 2 Acetate 5 ATP 2 NADH Oxaloacetate ATP synthase 2 NAD+ ETC Malate Citrate 2 NAD+ Isocitrate Fumarate 2 NADH Alpha ketoglutarate ATP synthase 2 3 ATP ETC 2 FADH2 Succinate NAD+ 5 ATP ETC ATP synthase Succinyl CoA 2 NADH 2 FAD 2 GTP ATP synthase ETC 5 ATP 2 GDP + 2 Pi FIGURE 2.10.13 Overall ATP production from glycolysis, TCA cycle, and ETC. Glycolysis produces 4 ATP, but consumes 2 ATP molecules, per mole of glucose. It also produces 2 moles NADH per mole of glucose. These reducing equivalents are cytosolic and can result in 35 moles of ATP depending on how the NADH enters the mitochondria. In the mitochondria, a total of 8 moles of NADH are produced per mole of glucose, including 2 in the pyruvate dehydrogenase step and 6 in the TCA cycle. These each produce 2.5 moles of ATP per mole of NADH, so the NADH produces 8 moles NADH/mol glucose 3 2.5 moles ATP/mol NADH 5 20 moles ATP/mol glucose. The 2 FADH2 produced at succinate dehydrogenase produce 1.5 moles of ATP per mole of FADH2 for a total of 3 moles ATP/mol glucose. Succinyl CoA synthetase produces 2 moles GTP/mol glucose, which is energetically equivalent to 2 moles ATP/mol glucose. Total ATP production is thus 57 moles in glycolysis 125 moles in the mitochondria, for a total of 3032 moles of ATP per mole of glucose. ATP through the malate/aspartate shuttle or 3 moles of ATP through the glycerol phosphate shuttle. MITOCHONDRIA HAVE SPECIFIC TRANSPORT MECHANISMS In order for oxidative phosphorylation to work, the inner mitochondrial membrane must be impermeable to H1 ions. H1 ions in solution are usually bound to water as H3O1, the hydronium ion, which is very small. Therefore, the inner mitochondrial membrane must be relatively impermeable to most ions in order for the membrane to establish the potential difference and concentration difference in [H1]. At the same time, things have to be able to get in and out. We have already discussed two such carriers—those that operate the shuttles that allow cytosolic reducing equivalents to enter the matrix. Several other carriers are shown schematically in Figure 2.10.14. ATP produced in the matrix must leave the matrix to power cellular activities. This occurs through facilitated diffusion by the ATPADP translocase. Since ATP has AT P P roduc ti on II : T h e T CA Cy c le a nd O xidat ive P hos phor yla tion –160 mV – Outer mitochondrial membrane + ATP/ADP Pi transporter Pyruvate translocase carrier Intermembrane space H+ ADP3– 10–7 M = pH 7.0 PO4–3 Pyruvate +++++++++ ETC Inner mitochondrial membrane ----------- NADH NAD+ 10–8 M = pH 8.0 ATP4– H+ OH– OH– Mitochondrial matrix –160 mV – Outer mitochondrial membrane + Intermembrane space H+ ETC Na/H Na/Ca Electrophoretic Exchanger Exchanger uniport 10–7 M = pH 7.0 Na+ Na+ Ca2+ + + + + + + + + + Inner mitochondrial membrane – – – – – – – – – NADH Na+ NAD+ H+ 10–8 M = pH 8.0 Ca2+ Ca2+ Mitochondrial matrix FIGURE 2.10.14 Selected transport mechanisms in mitochondria. See text for details. more negative charges, and its concentration is higher in the mitochondria where it is produced, the transport is “downhill” in both directions. This translocase is poisoned by atractyloside. Phosphate must also enter the matrix in order to be incorporated into ATP. Since phosphate is highly charged, movement across the membrane against a strong electric force is energetically unfavorable. Phosphate is carried across in exchange for OH2 ions. The outward flow of OH2 is equivalent to an inward flow of H1. Therefore, some of the energy of the electrochemical H1 gradient is used to transport phosphate into the mitochondria. Mitochondria also take up Ca21 because of the large negative potential inside. This uptake occurs through a channel called the electrophoretic uniport. This name signifies that the electrical gradient is the driving force for Ca21 movement and that only one ion moves (uniport). Ca21 taken up this way must be able to exit. This is accomplished by a Na/Ca exchanger that couples Ca21 exit with Na1 entry. This is another example of secondary active transport, in which “uphill” transport of Ca21 is linked to “downhill” movement of Na1. Of course, the Na1 taken up by this process must also have an exit. Na1 efflux from the mitochondria is through a Na/H exchanger. The Na/H exchanger is also an example of secondary active transport, in which uphill movement of Na1 is coupled to downhill movement of H1. In the mitochondria neither Na1 efflux nor H1 entry is linked to ATP hydrolysis. They are powered by the ETC pumping H1 out of the mitochondrial matrix into the intermembrane space, thereby creating a large potential and a concentration gradient for H1 ions. SUMMARY Glycolysis converts glucose into pyruvate, which enters the mitochondria by facilitated diffusion on its own carrier. The mitochondria couples the oxidation of pyruvate to the formation of the high-energy chemical bond in ATP, a process called oxidative phosphorylation. Inside the mitochondrial matrix, pyruvate enters into a series of reactions, beginning with pyruvate dehydrogenase. This series of reactions converts the three-carbon pyruvate molecule into three molecules of CO2. In the first step, pyruvate dehydrogenase converts pyruvate into acetyl coenzyme A, producing a molecule of CO2 and a molecule of NADH, nicotinamide adenine dinucleotide. Some of the chemical energy of oxidation of pyruvate is captured in the energy of electrons binding to NADH. 239 240 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION This energy is converted to ATP later on by the ATP synthase that relies on the electrochemical gradient of H1 that is established by the ETC. The second reaction combines acetyl CoA with oxaloacetate to form citric acid. This is the first step in the citric acid cycle, or the tricarboxylic acid cycle (TCA cycle), also called the Krebs cycle for Sir Hans Krebs. This cycle is a catalytic cycle in that it regenerates oxaloacetate and all parts of the acetate part of acetyl CoA are converted to CO2 or H2O. The cycle captures energy of oxidation by forming NADH or FADH2, flavin adenine dinucleotide, or by forming GTP directly in the conversion of succinyl CoA to succinate by succinyl CoA synthetase. Each turn of the cycle produces 3 NADH molecules, 1 FADH2 molecule, 1 molecule of GTP, and 2 molecules of CO2. The mitochondria make additional ATP molecules from NADH and FADH2. These transfer their electrons to the ETC, which uses the energy of binding of the electrons, expressed as the reduction potential, to pump hydrogen ions from the mitochondrial matrix to the intermembrane space. This active pumping produces a concentration difference of H1 ions and a large electrical potential. The ATP synthase in the inner mitochondrial membrane uses the energy in the electrochemical potential difference for H1 to make ATP by a process that converts electrochemical energy to mechanical energy and back again to ATP. NADH enters the ETC early and its energy causes H1 pumping at complexes I, III, and IV of the ETC. A total of 10H1 are pumped out per NADH molecule. The ATP synthase requires 4H1 ions to make ATP (about 3 for the direct use of the ATP synthase and 1 for transport of reactants into the mitochondria). As a consequence, each NADH produces about 2.5 ATP molecules. FADH2 enters the ETC after complex I, and so it makes only 1.5 ATP molecules. Thus complete oxidation of pyruvate to CO2 and H2O forms 4 NADH, 1 FADH2, and 1 GTP for a total equivalent of 4 3 2.5 1 1 3 1.5 1 1 5 12.5 ATP molecules. Since there are two pyruvate molecules formed from glucose, the TCA cycle accounts for 25 ATP molecules per glucose molecule. NADH is also generated by glycolysis in the cytoplasm. This NADH is oxidized in the mitochondria, but indirectly because NADH itself cannot cross the inner mitochondrial membrane. Instead, two shuttle systems have the effect of transferring cytosolic NADH to mitochondrial matrix NADH. The glycerol phosphate shuttle converts cytosolic NADH to mitochondrial FADH2, whereas the malate shuttle converts it to mitochondrial NADH. Glycolysis produces a net gain of 2 ATP and 2 NADH per glucose molecule. REVIEW QUESTIONS 1. How does pyruvate get into the mitochondria? How does cytosolic NADH get into the mitochondria? How do ADP and Pi get into the mitochondria? How do ATP get out of the mitochondria? 2. Is NADH reduced or oxidized? Is FAD reduced or oxidized? Why does NADH make more ATP than FADH2? 3. What is the TCA cycle? How many CO2 molecules are released per pyruvate? Per acetate? In general, where is CO2 released? 4. How many ATP molecules are produced during glycolysis per mole of glucose? Does this ATP production require oxygen? 5. What determines the direction of electron flow in the ETC? How do you calculate the energy of a reductionoxidation reaction? 6. How many ATP molecules are produced in mitochondria during oxidative phosphorylation? Does this ATP production require oxygen? 7. Why is there a membrane potential across the inner mitochondrial membrane? If the ATP synthetase lets in H1, why does not this current depolarize the mitochondria? 8. What is the chemiosmotic hypothesis? Some materials are proton ionophores. What effect would these have on oxidative phosphorylation? ATP Production III: Fatty Acid Oxidation and Amino Acid Oxidation Learning Objectives G G G G G G G G G G G G G Describe the chemical structure of a triglyceride Describe how adipocyte lipolysis is regulated by catecholamines and insulin Describe the main route of glycerol oxidation Describe the role of carnitine in the import of fatty acids into mitochondria List the number of NADH, FADH2, and acetyl CoA produced for each turn of the beta oxidation cycle Account for the number of beta oxidation cycles for palmitic acid List the number of NADH, FADH2, and GTP produced for oxidation of acetyl CoA Account for the total numbers of ATP molecules produced by oxidation of palmitic acid List three chemicals that comprise the ketone bodies Distinguish between the terms glucogenic and ketogenic for amino acids List the amino acids that are exclusively ketogenic Recognize the amino acids that are exclusively glucogenic Name the amino acid that is required for feed into the urea cycle 2.11 DEPOT FAT IS STORED AS TRIGLYCERIDES AND BROKEN DOWN TO GLYCEROL AND FATTY ACIDS FOR ENERGY Most fatty acids in the body are stored as triglycerides (triacylglycerol, TAG; see Figure 2.11.1), the acyl esters of three fatty acids with a glycerol molecule. Blood carries triglycerides in special structures called lipoproteins, in which the water-insoluble lipids are coated with special proteins. These lipoproteins are made solely in the intestines and the liver. All tissues can also store triglycerides in lipid droplets in the cytoplasm, Tripalmitin H H H H 1C 2C O Glycerol C 3 O O C O C O CH2 CH2 CH2 H2C H2C H2C CH2 CH2 CH O C 2 H2C FATS AND PROTEINS CONTRIBUTE 50% OF THE ENERGY CONTENT OF MANY DIETS In the previous chapters, we saw how carbohydrates are metabolized through glycolysis to form pyruvate, producing some energy in the process. Pyruvate is then converted to acetyl CoA, which enters the TCA cycle to produce reducing equivalents. These reduced compounds, NADH and FADH2, are then oxidized by the respiratory chain of the mitochondria to produce a proton electrochemical gradient that is then used to produce ATP. The typical American diet contains only about 49% of its calories as carbohydrates, with another 35% coming from fat and 16% from protein. Thus the fats and proteins must also be used to generate cellular energy. How are fats and proteins used to make ATP? H2C H2C CH2 CH2 H2C H2C H2C CH2 CH2 CH2 CH2 H2C H2C H2C H2C H2C CH2 CH2 CH2 CH2 CH2 H2C H2C CH2 H2C H2C Palmitic acid CH2 CH2 H2C H2C H2C CH3 CH3 CH3 CH2 FIGURE 2.11.1 Structure of tripalmitin as an example of a triglyceride. These are stored in adipose tissue and released into the circulation. Lipase breaks down the triglyceride by hydrolyzing the ester bonds between the fatty acids carboxyl group and the glycerol hydroxyl group. 241 © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00022-7 242 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Epinephrine - + TAG DAG ATGL Glycerol Natriuretic peptide Insulin + + MAG HSL MGL Glycerol Fatty acids Triglyceride Diacylglycerol Free fatty acid Monoacylglycerol Free fatty acid Free fatty acid FIGURE 2.11.2 Adipocyte lipolysis. Adipose triglyceride lipase (ATGL) begins lipolysis by converting triglycerides (TAG) in the lipid droplet to diacylglycerol (DAG) and a free fatty acid. Hormone-sensitive lipase (HSL) continues the process by hydrolyzing a second fatty acid, producing monoacylglycerol (MAG). Monoacylglycerol lipase (MGL) then hydrolyzes the last fatty acid ester bond to produce glycerol and a third free fatty acid. Hormones and nerves control lipolysis by activating or inhibiting ATGL and HSL. Epinephrine can activate ATGL and HSL through β receptors and this action is blocked by insulin. Epinephrine can inhibit ATGL and HSL through α2 receptors. Natriuretic peptides also can activate lipolysis through independent mechanisms. but adipocytes contain very large lipid droplets that may occupy most of the cell. The adipocytes are specialized for storage and release of lipids for energy. The circulating triglycerides can be hydrolyzed to form glycerol and three fatty acids through the action of lipoprotein lipase that both hydrolyzes the lipid and acts as a bridge in lipoprotein uptake. Lipid droplets can also be hydrolyzed for energy within the cell, but the adipocyte alone can export the glycerol and free fatty acids derived from the lipid store. Lipid droplets are surrounded by a phospholipid monolayer to which is absorbed the protein perilipin, which has five members (PLIN1-5) that are distributed in different tissues. In adipose tissue, the major form is PLIN1. PLIN1 stabilizes the lipid droplet against lipolysis. Lipolysis begins by ATGL, adipose triglyceride lipase, that converts tricacylglycerides (TAG) into diacylglycerol (DAG) and a free fatty acid (FA). Hormone-sensitive lipase (HSL) then converts the DAG into monoacylglycerol (MAG) and a free fatty acid. The MAG is further converted to glycerol and an FA by the enzyme monoglycerol lipase. Control of this process is hormonal and neural, as shown in Figures 2.11.2 and 2.11.3. The catecholamines, derived either from circulating epinephrine or from sympathetic nervous stimulation, can either activate or inhibit lipolysis. Insulin inhibits lipolysis, and natriuretic peptide stimulates it. In Chapter 2.9, we learned that epinephrine mobilizes liver glycogen stores by activating glycogenolysis through a Gs mechanism. Epinephrine can also mobilize lipid stores by activating lipolysis through a Gs mechanism, as shown in Figure 2.11.3. Binding of catecholamines to β1 or β2 receptors on the adipocyte is followed by the activation of adenylyl cyclase and an increase in 3’,5’-cyclic AMP in the adipocyte cytosol. This activates protein kinase A (PKA) that phosphorylates a number of targets including perilipin, PLIN1, and HSL (hormone-sensitive lipase). PLIN1 binds ABHD5 (α/β hydrolase domain containing 5) that activates ATGL. Phosphorylation of PLIN1 releases the ABHD5 to activate ATGL. Simultaneously, HSL that was previously located in the cytosol binds to the lipid droplet where it participates in lipolysis. Adipocytes also have α2 receptors that are coupled to a Gi mechanism that inhibits adenylyl cyclase activity. Occupancy of these receptors has the opposite effect, an inhibition of lipolysis, as occupancy of the β receptors. Insulin inhibits lipolysis by activating cAMP phosphodiesterase, type 3B, by phosphorylation by PKB. This activated PDE-3B decreases the cAMP concentration, leading to inactivation of lipolysis. Natriuretic peptide increases lipolysis by activation of HSL through phosphorylation by PKG, a cGMPactivated protein kinase. GLYCEROL IS CONVERTED TO AN INTERMEDIATE OF GLYCOLYSIS Plasma glycerol is taken up by tissues and converted to α-glycerophosphate by an enzyme called glycerol kinase, by phosphorylating the glycerol with the terminal phosphate group of ATP. The glycerophosphate is converted to dihydroxyacetone phosphate ATP P roduction III: F att y Acid O xidat ion and A mino Aci d Oxi da tion Epinephrine βγ subunits Adenylyl cyclase Epinephrine α2 receptor βγ subunits - β1 or β2 receptor + Gα subunit PDE-3B P cAMP GDP ATP + PKB Gα subunit Inactive PKA IRS-1 + P + AMP Active PKA HSL Insulin PI-3K + MGL ATGL ABHD5 P P PLIN1 P ABHD5 Insulin receptor P IRS-2 TAG PLIN1 ATGL DAG DAG MAG MGL TAG Basal state P HSL P P Glycerol + 3 FA Stimulated state Lipid droplet FIGURE 2.11.3 Control of lipolysis by hormones. Binding of epinephrine to β1 or β2 receptors activates adenylyl cyclase by the Gαs subunit. This increases formation of cAMP and therefore increases the concentration of cAMP in the cell. Increased cAMP activates protein kinase A that phosphorylates PLIN1 and HSL. Phosphorylation of PLIN1 releases ABHD5 that then binds to ATGL and activates it. PKA phosphorylation of HSL results in its translocation from the cytosol to the lipid droplet surface. In this position, the enzymes completely hydrolyze TAG to glycerol and 3 fatty acids, which are exported out of the adipocyte. Lipolysis is inhibited by binding of epinephrine to α2 receptors that inhibit adenylyl cyclase. Insulin binding to its receptor results in a cascade that activates PKB that phosphorylates cAMP phosphodiesterase. This lowers the cAMP concentration and inhibits lipolysis. by another enzyme, α-glycerophosphate dehydrogenase, requiring NAD1, and thus the glycerol can enter into glycolysis and the TCA cycle to be fully oxidized to CO2 and H2O to provide energy as ATP to the cell (see Figure 2.11.4). FATTY ACIDS ARE METABOLIZED IN THE MITOCHONDRIA AND PEROXISOMES Free fatty acids are formed in the cytoplasm by the action of lipase on stored triglycerides, but the fatty acids themselves are degraded and oxidized only in the mitochondria and peroxisomes. The fatty acids have surface activity (they lower the surface tension) and can impair membrane integrity. Therefore, the fatty acids are carried in solution by fatty acid binding proteins. These are low-molecular-weight proteins (about 14,000 Da) that probably have a dual function of decreasing the concentration of the free fatty acids and of enhancing the diffusion through the cytoplasm by carrying the fatty acids. The first step in the metabolism of free fatty acids is their import into the inner mitochondrial matrix by combining with a carrier substance, carnitine (see Figure 2.11.5). First, the fatty acid is combined with coenzyme A by the enzyme thiokinase, which hydrolyzes ATP to AMP and PPi. This fatty acyl CoA is then transferred to carnitine by the enzyme carnitine fatty acyl transferase. Once the fatty acyl carnitine is transferred to the mitochondrial matrix, it is once again combined with CoA to form fatty acyl CoA. In this form and in this place, the fatty acid can be oxidized in a systematic way to produce energy. BETA OXIDATION CLEAVES TWO CARBON PIECES OFF FATTY ACIDS Beta oxidation is the process by which fatty acids are processed progressively to release two-carbon segments in the form of acetyl CoA. This series of reactions is summarized in Figure 2.11.6. These reactions produce acetyl CoA and 1 FADH2 and 1 NADH per turn of the beta oxidation cycle. 243 CH2OH OH O Glucose OH OH OH ATP Glucokinase (liver) Hexokinase (muscle) ADP CH2OH OH CH2OPO3 Phosphoglucomutase O OH O Glucose-6-phosphate Glycogen OH OH OPO3 OH OH OH Glucose-1-phosphate Phosphoglucose isomerase CH2OPO3 OH O OH Pi H2C Glycerol OH HC OH H2C OH OH Phosphofructokinase ADP CH2OPO3 OH O HO ATP Glycerol kinase ADP CH2OPO3 Glycerophosphate HC OH H2C OH ATP Fructose-1,6-diphosphatase H2O NAD+ OH NADH + H+ Fructose-6-phosphate CH2OH Fructose-1,6-diphosphate CH2OPO3 Aldolase CH2OPO3 HC O C HC OH O Glycerophosphate dehydrogenase CH OH Triose phosphate isomerase 2 Dihydroxyacetone phosphate CH2OPO3 Glyceraldehyde-3-phosphate Pyruvate FIGURE 2.11.4 Metabolism of glycerol. Glycerol is phosphorylated and then converted into dihydroxyacetone phosphate, an intermediate in glycolysis. EXAMPLE 2.11.1 ATP Yield from Glycerol How much ATP is made from glycerol, under optimal conditions? Glycerol requires 1 ATP per molecule for the conversion into glycerol phosphate. The glycerol phosphate is then converted to dihydroxyacetone phosphate with the production of 1 molecule of NADH. The dihydroxyacetone phosphate can then be oxidized fully to CO2 through glycolysis and the TCA cycle as outlined in Chapters 2.9 and 2.10. In the conversion to pyruvic acid, dihydroxyacetone phosphate generates 1 NADH and 2 ATP molecules per molecule of glycerol. The net effect of converting glycerol to pyruvic acid is 21 ATP 1 2 ATP 1 2 NADH 5 1 ATP 1 2 NADH 5 6 ATP, assuming that the NADH are both oxidized with the generation of 2.5 ATP per NADH. The complete oxidation of pyruvate produces 4 NADH, 1 FADH2, and 1 GTP. When the reducing equivalents are oxidized through the electron transport chain, this produces 4 NADH 3 2:5 ATP=NADH 1 1 FADH2 3 1:5ATP=FADH2 1 1 GTP 3 1 ATP=GTP 5 12:5 ATP Each glycerol molecule liberated from a triglyceride thus produces at most 18.5 molecules ATP per molecule glycerol. Glycerol has a gram molecular weight of 92 g mol21. It produces 18.5 moles ATP/92 g mol21 5 0.20 mol g21. This is a little more than the energy derived from glucose: 32 moles ATP/180 g mol21 5 0.18 mol g21. ATP P roduction III: F att y Acid O xidat ion and A mino Aci d Oxi da tion O O H2 R C C C OH H2 Free fatty acid R C H2 ATP AMP + PPi O H2 C C S CoA AMP + PPi O CH3 H C C H2 CH3 O CH3 N Carnitine CH3 O H C C C C OH H2 H2 CH3 OH C N H2 C R H2 C H2 C C S CoA Fatty acyl CoA O FAD Acyl CoA dehydrogenase C C OH H2 FADH2 O C O C R R H2 C C H H C C Cytoplasm Carnitine acyltransferase I S CoA Trans enoyl CoA H2O Outer mitochondrial membrane Enoyl CoA hydratase Intermembrane space Acylcarnitine translocase OH Inner mitochondrial membrane R Carnitine acyltransferase II CH3 CH3 O H CH3 N C C C C OH H2 H2 CH3 OH H2 C C H O H2 C H C C H2 CH3 O N C CoA 3-Hydroxy acyl CoA NAD+ 3-Hydroxy acyl CoA dehydrogenase O NADH + H+ C C OH H2 O S C Mitochondrial matrix CoA-SH Carnitine CH3 Free fatty acid OH Thiokinase Fatty acyl carnitine CH3 C CoA Thiokinase R H2 C ATP CoA Fatty acyl CoA H2 C C H2 O R H2 C C O H2 C C C S CoA 3-Keto acyl CoA CoA C R C H2 O H2 C C S CoA Thiolase R Fatty acyl CoA O R H2 C C O S CoA + H3C Fatty acyl CoA FIGURE 2.11.5 Involvement of carnitine in the entry of free fatty acids into the inner mitochondrial space. Fatty acids are combined with coenzyme A in a reaction that effectively costs 2 ATP molecules per reaction. The fatty acyl CoA is then converted to fatty acyl carnitine, which can penetrate the inner mitochondrial membrane. In the mitochondrial matrix, the carnitine is removed by a second, different carnitine acyl transferase. Once coupled with coenzyme A, the fatty acyl chain is oxidized, producing FADH2. The fatty acyl chain is further oxidized by adding water and then removing two more hydrogens by the 3-hydroxy acyl CoA dehydrogenase, this time producing NADH. The final step in each turn of the beta oxidation cycle is the production of acetyl CoA and the regeneration of fatty acyl CoA, shortened by two carbon atoms. This shorter fatty acyl CoA reenters the beta oxidation cycle until, at the end, all of the chain is converted to acetyl CoA. In this way, palmitic acid, for example, will produce 8 acetyl CoA molecules and 7 FADH2 molecules and 7 NADH molecules (there are only 7 because the last turn of the cycle produces two acetyl CoA molecules and so the last two carbons do not enter the cycle again to produce FADH2 and NADH). C S CoA Acetyl CoA FIGURE 2.11.6 Beta oxidation of free fatty acid. THE LIVER PACKAGES SUBSTRATES FOR ENERGY PRODUCTION BY OTHER TISSUES During exercise, when fatty oxidation occurs rapidly, the liver packages acetyl CoA in a form that can be readily used by other tissues for the generation of energy. In the liver, two molecules of acetyl CoA combine to form acetoacetate. The acetoacetate that forms can then be converted to β-hydroxybutyric acid and, to a lesser extent, to acetone. All three of these compounds, acetoacetate, β-hydroxybutyric acid, and acetone, are referred to as ketone bodies (see Figure 2.11.7). They leave the liver cell and travel in the blood to the peripheral tissues, which take up the compounds and metabolize them for energy. The combined concentration of the ketone bodies is typically less than about 3 mg%. Despite these low 245 246 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION EXAMPLE 2.11.2 ATP Yield from Fatty Acids Each turn of the beta oxidation cycle produces NADH and FADH2. For palmitic acid, C16H32O2, a total of eight acetyl CoA molecules are produced from seven turns of the beta oxidation cycle. The NADH and FADH2 feed into the electron transport chain to produce ATP. The total ATP produced from beta oxidation is thus 7 NADH=palmitate 3 2:5 ATP=NADH 1 7 FADH2 =palmitate 3 1:5 ATP=FADH2 5 28 ATP=palmitate Each acetyl CoA produced by beta oxidation enters the TCA cycle where it is further oxidized to CO2 and produces reducing equivalents that are used by the ETC to make ATP. Each acetyl CoA molecule produces 3 NADH, 1 FADH2, and 1 GTP molecule during a single turn of the TCA cycle. These are eventually used by the ETC to produce 3 NADH=acetyl CoA32:5 ATP=NADH11 FADH2 =acetylCoA 31:5 ATP=FADH2 11 GTP=acetyl31 ATP=GTP 510 ATP=acetyl CoA Since there are eight acetyl CoA molecules per palmitic acid produced from beta oxidation, the total ATP produced from acetyl CoA from palmitic acid is 10 ATP=acetyl CoA 3 8 acetyl CoA=palmitate 5 80 ATP=palmitate We add this to the ATP produced from beta oxidation and subtract the two ATP needed to start beta oxidation from the initial thiokinase reaction to get 28 ATP from beta oxidation 1 80 ATP from TCA cycle 2 ATP from thiokinase reaction 5 106 ATP=palmitic acid EXAMPLE 2.11.3 Compare ATP Yield from Glucose to that of Tripalmitin In Chapter 2.10, we found that the maximum yield of ATP from the complete oxidation of glucose was 32 ATP per glucose. This corresponds to 7 ATP per glucose from the oxidation of glucose to pyruvate, and 25 ATP from the complete oxidation of pyruvate to CO2 and H2O. The gram molecular weight of glucose is 180 g mol21, so the ATP production is glycerol. We have calculated that glycerol gives rise to at most 18.5 ATP per glycerol, and each palmitic acid produces at most 106 ATP per palmitic. The total for tripalmitin is thus 32 moles ATP=mol glucose 5 32 moles ATP=180 g 5 0:18 mol ATP per g of glucose Thus fat has about 2.3 times as much energy stored per unit weight. 336:5 moles ATP=mole tripalmitin 5 336:5 moles=807:3 g 5 0:42 ATP per g of tripalmitin The gram molecular weight of tripalmitin is 807.3 g mol21, tripalmitin consisting of three palmitic acid molecules and one concentrations, the flux of energy to the metabolizing tissues can be great because the ketone bodies are taken up so quickly. The concentrations of the ketone bodies can occasionally rise to very high levels, a condition known as ketosis. This condition occurs whenever metabolism of fats is emphasized such as in starvation and in diabetes mellitus. In this case, the urine can contain ketones and the presence of acetone is sometimes detectable by its odor in the exhaled air. In the peripheral tissues, acetoacetic acid is taken up and converted in the mitochondria to acetoacetyl CoA by the transfer of a CoA moiety from succinyl CoA, an intermediate in the TCA cycle. Since succinyl CoA is usually converted to succinate with the formation of 1 molecule of GTP, conversion of succinyl CoA to succinate in this reaction removes the potential synthesis of 1 molecule of GTP. So, although no energy in the form of ATP is directly involved in this transfer, it has a net cost of 1 molecule of ATP. The acetoacetyl CoA can then be converted to two molecules of acetyl CoA by thiolase, the same enzyme involved in the production of acetyl CoA from fatty acyl CoA during the beta oxidation pathway of fatty acids. AMINO ACIDS CAN BE USED TO GENERATE ATP Amino acids can be used to build body proteins and they can be broken down to yield energy. In the steady state of the adult, the body store of proteins remains constant and there is a constant throughput of amino acids, equal to the dietary intake, that is converted to metabolic energy. In the typical American diet, about 16% of the calories are provided by dietary protein. Because the hepatic portal blood leaves the intestines and travels to the liver, the liver has the first opportunity to metabolize all the nutrients, including the amino acids absorbed from digested proteins in the intestinal lumen. The liver does several things: it catabolizes a ATP P roduction III: F att y Acid O xidat ion and A mino Aci d Oxi da tion CoA-SH O H3C C Acetyl CoA S 2 H3C C CoA + CoA Thiolase + O HS Acetoacetate Acetyl CoA S O H3C CoA C O H2 C C OH NAD + H+ Beta-hydroxy butyric acid dehydrogenase CO2 NADH OH H3C C H O H2 C O OH C Beta-hydroxy butyric acid H3C C CH3 Acetone FIGURE 2.11.7 Formation of ketone bodies by the liver. During rapid lipid metabolism, when production of acetyl CoA outstrips the liver’s ability to oxidize it, acetyl CoA coalesces to form acetoacetate, beta-hydroxybutyric acid, and acetone. large fraction of the amino acids (57%), releases some unchanged into the general circulation (23%), and utilizes some 20% for synthesis of proteins that either remain in the liver or are released into the blood. Catabolism of amino acids can be broadly categorized into two processes: the breakdown of amino acids to carbohydrate precursors and potentially leading to the formation of glucose; and transformations leading to acetyl CoA that result in the potential formation of ketone bodies. Amino acids that break down into carbohydrate precursors are called glucogenic; those leading to acetyl CoA are called ketogenic. G G G Leucine and lysine are the only exclusively ketogenic amino acids. Isoleucine, threonine, phenylalanine, tyrosine, and tryptophan are both glucogenic and ketogenic. Aspartatic acid, asparagine, glutamic acid, glutamine, alanine, arginine, histidine, glycine, serine, proline, valine, methionine, and cysteine are glucogenic. Because each amino acid has a different side chain, each amino acid is catabolized differently to produce energy and waste products. We will not go through all of these reactions for each of the amino acids. The overall fate of the amino acids is shown in Figure 2.11.8. AMINO ACIDS ARE DEAMINATED TO ENABLE OXIDATION Many amino acids share a common mechanism for the removal of the amino group to form intermediates in the TCA cycle or glycolytic cascade. This is a transamination reaction followed by a dehydrogenation reaction, as shown in Figure 2.11.9. The α-ketoglutarate formed in the transamination reaction in the mitochondria can then enter the TCA cycle. The deamination results in the liberation of ammonia. The reaction sequence shown is one of many such involved in the deamination of a variety of amino acids including phenylalanine, tyrosine, aspartate, cysteine, lysine, arginine, alanine, isoleucine, leucine, and valine. The reactions differ only in the α-keto acid formed following the transamination. UREA IS PRODUCED DURING DEAMINATION AND IS ELIMINATED AS A WASTE PRODUCT The ammonia released during deamination is removed from the blood almost entirely by conversion into urea in the liver. This occurs through another metabolic process called the urea cycle (see Figure 2.11.10). In this process, the ammonia is combined with bicarbonate ion to form carbamoyl phosphate. The complete operation of the cycle requires continual input of aspartate. This can be derived from transamination of oxaloacetic acid by glutamic acid, the reverse of the process shown in Figure 2.11.9. Since glutamate is the product of transamination with several amino acids, it can be replenished. Thus one of the amino groups of urea is derived from ammonia and the other is derived from amino groups on various amino acids, transaminated to glutamate. 247 248 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Aspartic acid Asparagine GDP GTP CO2 Phosphoenolpyruvate Oxaloacetate Phenylalanine tyrosine Phosphoenolpyruvate carboxykinase ADP NADH + H Pyruvate kinase Tryptophan Threonine Cysteine Serine Glycine Alanine Malate dehydrogenase NAD ATP + Pyruvate Malate Fumarate Cytoplasm Outer mitochondrial membrane Dicarboxylic acid carrier Intermembrane space Pyruvate carrier Inner mitochondrial membrane Mitochondrial matrix CoA-SH – OH NAD + Phenylalanine Tyrosine Tryptophan Threonine Isoleucine Lysine Leucine Pyruvate dehydrogenase CO2 NADH + H Acetyl CoA Citrate Oxaloacetate Aconitase Citrate synthase Malate dehydrogenase CoA-SH NADH + H Isocitrate Malate NAD + NAD + Isocitrate dehydrogenase Fumarase TCA cycle CO2 NADH + H Fumarate FADH2 Alpha-ketoglutarate + NAD FAD Succinate dehydrogenase CoA-SH NADH + H Succinyl CoA Glutamic acid Glutamine Proline Histidine Arginine CoA-SH Alpha-ketoglutarate dehydrogenase CO2 Succinate GTP GDP Pi Succinyl CoA synthetase Pyruvate dehydrogenase Isoleucine Valine Methionine FIGURE 2.11.8 Metabolic entry points for the catabolism of the amino acids. Those amino acids that produce acetyl CoA are called ketogenic. These include leucine, lysine, phenylalanine, tyrosine, tryptophan, and isoleucine. Those amino acids that produce carbohydrate precursors that can be converted to glucose are called glucogenic. These include aspartic acid, asparagine, phenylalanine, tyrosine, tryptophan, alanine, cysteine, serine, threonine, glycine, glutamic acid, glutamine, proline, histidine, arginine, isoleucine, valine, and methionine. Only lysine and leucine are exclusively ketogenic. Exclusively ketogenic amino acids are in light blue italic; exclusively glucogenic are in dark blue; both ketogenic and glucogenic are in black. ATP P roduction III: F att y Acid O xidat ion and A mino Aci d Oxi da tion COO– Aspartate +H N 3 COO– CH O C CH2 CH2 COO– COO– Aspartate aminotransferase COO– H2C Glutamate CH2 CH2 COO– C COO– H2C Alpha-ketoglutarate O Oxaloacetate +H 3N COO– C H Cytoplasm Outer mitochondrial membrane Intermembrane space Dicarboxylic acid carrier Inner mitochondrial membrane Mitochondrial matrix COO– H2C Glutamate dehydrogenase CH2 Alpha-ketoglutarate O COO– C COO– H2C CH2 +H 3N C H Glutamate COO– H2O + NH4 NADH + H+ NAD+ FIGURE 2.11.9 Deamination of amino acids by aminotransferase and glutamate dehydrogenase action. 249 250 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION COO– + H3N Aspartate CH CH2 COO– AMP + PPi ATP COO– COO– + + H3N H3N CH (CH2)3 Citrulline CH Arginosuccinate (CH2)3 Arginosuccinate synthetase COO– NH NH CH HCO3– NH2 NH2 Pi Bicarbonate NH CH O C CH2 COO– 2ADP + Pi Urea cycle Lyase 2 ATP Fumarate COO– – PO4 C Ornithine transcarbamylase CH O COO– Carbamoyl phosphate NH4+ synthetase + NH4+ Carbamoyl phosphate + Ammonia H3N Ornithine H3N COO– CH CH COO– (CH2)3 CH NH (CH2)3 C NH2 H2O Arginine NH2 NH O NH2 C NH2 Urea FIGURE 2.11.10 Formation of urea from the urea cycle. Clinical Applications: Nonshivering Thermogenesis Immediately after birth, the baby is thrust into a cool and dry environment. It loses heat rapidly by evaporation. Even when its skin dries, the baby continues to lose heat because it has a large surface area relative to its small body mass, it cannot put on warmer clothes, it has poor thermal insulation in the form of fat, and it has limited ability to generate heat by muscular contractions. However, the baby is able to generate heat by nonshivering thermogenesis, also called metabolic thermogenesis. Babies have specialized fat tissue called brown fat, located mainly in the neck and in the midline of the back. Mitochondria contain cytochromes that contain iron, giving the mitochondria a reddish-brown color. Brown fat contains lots of mitochondria so it takes on a brownish color from mitochondrial pigments. The mitochondria in this fat make ATP primarily from the oxidation of fatty acids that are produced by hydrolysis of stored triglycerides. Under certain circumstances, these mitochondria can become uncoupled. Uncoupling of oxidative phosphorylation in brown fat mitochondria generates heat. Oxidative phosphorylation couples an exothermic reaction, one that releases heat, to an endothermic reaction, one that requires heat. Pumping of H1 ions out of the mitochondrial matrix couples the exothermic oxidation reactions to the energy of the H1 electrochemical gradient. This in turn is coupled to the endothermic synthesis of ATP. The oxidation of foodstuffs (carbohydrates, fats, and amino acids) reduces NAD1 to NADH, whose oxidation powers the H1 pumping by the electron transport chain. The ATP synthase couples the energy stored in the electrochemical potential of H1 to ATP synthesis. This coupling can never be 100% efficient, so that cellular respiration releases some energy ATP P roduction III: F att y Acid O xidat ion and A mino Aci d Oxi da tion as heat. Under certain circumstances a special protein, called the uncoupling protein, short-circuits the synthesis of ATP by allowing H1 ions to cross the inner mitochondrial membrane without making ATP. Then more of the energy stored in the H1 gradient is dissipated as heat and less is captured by ATP. The exothermic reactions are uncoupled from ATP synthesis and the mitochondria generate heat proportionate to the number of mitochondria and the rate of oxygen consumption. content of the diet increases the tissue content of UCP2. Thus these proteins are believed to participate in basal or regulatory thermogenesis, but their exact functions are not yet worked out. The mechanism of brown fat thermogenesis in response to cold is shown in Figure 2.11.11. Adipose tissue is supplied by sympathetic nerves that release noradrenaline onto the fat cells. This stimulates the breakdown of stored triglycerides to fatty acids, which in turn activate UCP1. The effect depends on the tissue content of UCP1, which is also increased by cold exposure. How UCP1 increases the H1 flux across the membrane is also not yet known. UCP1 may form a H1-specific pore or it may transport H1 ions by cycling an H1 carrier, most likely fatty acids. Current work favors the H1 carrier mechanism because some carboxylic acid groups can activate transport without being transported. UCP1, a 32-kDa protein isolated from brown fat mitochondria, was the first uncoupling protein to be described. UCP1 is located mainly in brown fat, whereas UCP2 is expressed in many tissues and UCP3 is found mainly in skeletal muscle. BMCP1 is specific to the brain. UCP1 is activated by fatty acids and inhibited by purine nucleotides (ATP, ADP, GTP, and GDP). Exposure to cold increases UCP1 content in brown fat, and increased caloric Cold Diet Sympathetic nerves Norepinephrine Adenylyl cyclase Cell membrane β-Receptor GDP Gα subunit βγ subunits cAMP ATP + Triglyceride droplet + ATP Lipase, inactive ADP CREB Pi CREB + Active PKA Pi CREB CREB Pi DNA CRE Lipase, active mRNA FA Nucleus Outer mitochondrial membrane –160 mV H+ H+ ++++++ – + + ETC F0 NADH NAD --------- Inner mitochondrial membrane UCP1 F1 + ATP ADP + Pi H+ H+ Heat FIGURE 2.11.11 Postulated mechanisms for nonshivering thermogenesis in brown adipose tissue. Noradrenaline released from sympathetic nerve terminals in adipose tissue binds to β receptors on the surface of the adipocytes, leading to increased cytosolic cAMP by activating adenylyl cyclase. cAMP activates protein kinase A (PKA), which phosphorylates several targets, thereby changing their activity. PKA activates hormone-sensitive lipase that increases triglyceride breakdown and increases cytoplasmic fatty acid concentration. The increased fatty acids activate UCP1, short-circuiting the H1 electrochemical gradient and reducing the synthesis of ATP from ADP and Pi. This uncouples oxidative phosphorylation and the energy of oxidation of fats appears as heat. PKA also phosphorylates a transcription factor, CREB (cyclic AMP response element binding protein), that binds to specific regions of the DNA (CRE—cAMP response element) and activates their transcription into mRNA. This leads to increased numbers of UCP1 protein in the mitochondria, which adapts the body to exposure to cold or some other stress. Diet-induced thermogenesis probably involves increased expression of UCP2. 251 252 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION SUMMARY The oxidation of fatty acids and amino acids produces ATP through some of the same reactions used to produce ATP by the oxidation of carbohydrates. Fatty acids are released from triglycerides by hormonesensitive lipase. The fatty acids are bound in the cytoplasm by fatty acid binding protein and carried into the mitochondria by being complexed with carnitine. Inside the mitochondria, fatty acids undergo beta oxidation in which two carbons at a time are cleaved off the carboxyl end and converted into acetyl Coenzyme A. Each turn of the beta oxidation cycle produces one NADH and one FADH2. These feed reducing equivalents into the electron transport chain, which pumps H1 ions out of the mitochondrial matrix and produces the electrochemical gradient of H1 that drives ATP synthesis through the F0F1ATPase. Beta oxidation also produces acetyl CoA that enters the TCA cycle by combining with oxaloacetate to form citric acid. Each turn of the TCA cycle converts the acetyl CoA into 2 molecules of CO2, 3 NADH, 1 FADH2, and 1 GTP. Thus palmitic acid (16:0) produces 7 NADH and 7 FADH2 from beta oxidation, and 24 NADH, 8 FADH2, and 8 GTP from the complete oxidation of acetyl CoA. Since NADH produces 2.5 ATP molecules and FADH2 produces 1.5, the ATP count from palmitic acid is 28 from beta oxidation and 80 from the TCA cycle. Two ATP molecules are used to prime the palmitic acid in the thiokinase reaction that converts palmitic acid to palmitoyl CoA. The liver packages lipid metabolites into ketone bodies, which collectively consist of acetoacetic acid, betahydroxybutyric acid, and acetone. Build-up of these ketone bodies during starvation or other metabolic conditions such as diabetes is called ketosis. Each amino acid has its own metabolic pathway because they all differ chemically. However, they feed into the main metabolic pathways in a limited number of places. Those amino acids that can be used to make glucose are called glucogenic. Those that can be used to make ketone bodies by producing acetyl CoA are called ketogenic. Many amino acids are both glucogenic and ketogenic. Only leucine and lysine are exclusively ketogenic. Many amino acids are deaminated by a combination of transamination and dehydrogenation. In this reaction, the amino group is transferred from the amino acid to alpha-ketoglutaric acid, forming glutamic acid. The glutamic acid is then converted back to alpha-ketoglutaric acid by glutamate dehydrogenase, resulting in release of ammonia and formation of NADH. The ammonium formed in this way is converted to urea through the urea cycle. The urea cycle begins with the formation of carbamoyl phosphate by condensing ammonium with HCO32. Carbamoyl phosphate then combines with ornithine to form citrulline. Citrulline combines with aspartic acid to form arginosuccinate and arginine in sequence. Arginine gives rise to urea and ornithine to begin the cycle again. REVIEW QUESTIONS 1. What components make up a triglyceride? Where are triglycerides found in the body? Can they be metabolized as is? What enzyme breaks triglycerides down into components? What activates this enzyme? What inhibits the enzyme? 2. How is glycerol oxidized? How much ATP is produced from glycerol, mole per mole? 3. How are fatty acids carried in the cytosol? Where does oxidation of fatty acids take place? How do the fatty acids get into the mitochondria? 4. What is meant by beta oxidation? What are the main products? How many beta oxidation cycles are there for palmitic acid? Stearic acid? Oleic acid? 5. How much ATP is produced from the complete oxidation of palmitic acid, mole per mole? 6. What are the ketone bodies? Where are they produced? 7. What is a glucogenic amino acid? Which amino acids are glucogenic? 8. What is a ketogenic amino acid? Which amino acids are exclusively ketogenic? 9. What is transamination? What are the major receptors for transamination? 10. What is urea? Where is it produced? The Origin of the Resting Membrane Potential Learning Objectives G G G G G G G G Write the Nernst equation for any given ion Define equilibrium potential Be able to calculate the equilibrium potential for any ion Recognize the proper form of the GoldmanHodgkinKatz equation Explain why in the GoldmanHodgkinKatz equation the anion concentration in one compartment appears on the opposite side of the argument from the cation concentration in the same compartment Define and distinguish between slope conductance and chord conductance Write the chord conductance equation Explain what happens to membrane potential when the conductance to a particular ion changes INTRODUCTION In Chapter 2.6, we analyzed the energetics of membrane transport across resting cardiac muscle cells that have a membrane potential of about 280 mV, negative inside. At rest, the concentration gradient favored Na1 and Ca21 entry into the cell and K1 exit. These gradients produced slow leaks of ions that were continually balanced by active transport mechanisms such as the Na,K-ATPase and NaCa exchanger so that the ionic composition of the cytosol stayed constant. The resting membrane potential is extremely important because, first, all cells have a membrane potential (but not the same value!), and modulation of the membrane potential is associated with modulation of cellular activity. Certain cells of the body, called excitable cells, can use rapid changes in their membrane potential as a signal. These cells include nerve cells and muscle cells. Now we ask the question, where did the resting membrane potential come from? To answer this question, we will consider hypothetical membranes. These membranes are not like any biological membrane, but we consider them because they will clarify how the resting membrane potential comes to be what it is. Here we will use concepts of potential and capacitance already covered in Chapter 1.3, and the concept of the electrochemical potential discussed in Chapter 1.7. © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00023-9 3.1 THE EQUILIBRIUM POTENTIAL ARISES FROM THE BALANCE BETWEEN ELECTRICAL FORCE AND DIFFUSION First we consider a hypothetical membrane that is permeable only to Na1 ions. Suppose that [Na1]o, the outside or extracellular sodium concentration, is 145 3 1023 M, and [Na1]i 5 12 3 1023 M. And suppose that initially there is no membrane potential. What happens? As shown in Figure 3.1.1, the diffusion gradient for Na1 favors Na1 entry into the cell, and the initial Na1 influx carries a charge that builds up on the inside of the cell. This produces a potential (recall that separation of charge produces a potential) across the membrane that impedes further Na1 ion movement because the positive charges repel the positively charged Na1 ion. The potential that develops is related to the electric force that now works against further Na1 movement. Eventually the electric force will get so large that the diffusive flow will be exactly counterbalanced, and net flow will stop. This will occur at the Na equilibrium potential. We can calculate what that potential should be in two ways: first by looking at Fick’s law and second by analyzing the energetics. We get the same answer either way. The situation for the development of the Na1 equilibrium potential is shown in Figure 3.1.1. Fick’s law with an electrical force is given as ½3:1:1 Js 5 2D @C D @ψ 2 Cz` @x RT @x This is Eqn [1.7.19]. Here Js is the solute flux, in this case the flux of Na1, D is the diffusion coefficient, C is the concentration of Na1, R is the gas constant, T is the absolute temperature, z is the charge on the ion (11 for Na1), and ` is the faraday, the number of coulombs per mole. At equilibrium, Js 5 0, and we equate the diffusive force and the electrical force: ½3:1:2 RT @C @ψ 5 2z` C @x @x Integrating both sides from outside to inside of the membrane, we get ði ði RT @C @ψ dx 5 2z` dx C @x @x o o ½3:1:3 Ci RT ln 5 z`ðψo 2 ψi Þ Co 255 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 160 160 [Na+]o 145 mM Membrane Na+ flux is down its concentration gradient 130 Concentration (mM) 120 J = –D dC/dx 100 Na+ 90 80 70 60 50 Extracellular fluid Cytosol 120 110 60 50 [Na+]i 12 mM 20 [Na+]i 12 mM 10 0 10 0 160 160 [Na+ ]o 145 mM 140 150 3 120 J = –D dC/dx 100 Na+ – + 70 J = – D/RT CzF dV/dx Concentration (mM) The current builds up charge on the two sides of the membrane that oppose further Na+ flux 130 60 i 70 30 90 80 J = –D dC/dx 90 80 30 110 Na+ carries a charge; its flux makes a current 100 40 20 2 130 40 150 [Na+]o 145 mM 140 1 140 110 150 Concentration (mM) 150 Concentration (mM) 256 4 130 At equilibrium, the electrical force exactly balances diffusion 120 110 100 90 80 70 – + – + – + – + 60 50 40 40 30 30 [Na+]i 12 mM Membrane 140 50 20 [Na+]o 145 mM Extracellular fluid Cytosol [Na+]i 12 mM 20 10 0 10 0 FIGURE 3.1.1 Generation of the Na equilibrium potential across a hypothetical membrane that is permeable only to Na. The [Na]o in this case is high, about 145 mM, whereas [Na]i is 12 mM. Thus the diffusion gradient favors Na1 flux from the outside to the inside of the cell. Since Na1 is electrically charged, flux of only Na1 makes a current across the membrane that separates charges and thus produces an electrical potential. The electric field exerts forces on Na1 that retards its movement. When the diffusive force exactly balances the electrical force, flux is zero and the potential is ENa, the sodium equilibrium potential. This gives the equilibrium membrane potential as ½3:1:4 RT Ci ln 5 ðψo 2 ψi Þ z` Co It is usual for physiologists to take the outside solution as ground (ψo 5 0) because it is potential differences that we are concerned with. Using this standard ψo 5 0, we rewrite Eqn [3.1.4] as ½3:1:5 RT Co ln 5 ðψi 2 ψo Þ z` Ci If you always put the outside ion in the numerator of the argument of the logarithm, then the sign of the membrane potential will be correct. This is a famous equation, the Nernst equation, named for Walther Nernst (18641941), a physical chemist from Berlin, Germany. It calculates the potential at which net flux is zero, which occurs at equilibrium for the ion. The membrane potential at which the diffusive force is exactly balanced by the electrical force is called the equilibrium potential for that ion. For sodium, it is usually symbolized as ENa. The Nernst Equation can also be derived easily from considering the electrochemical potentials. At equilibrium we have Δμ 5 μNai 2 μNao ½3:1:6 0 5 μoNai 1 RT ln½Na1 i 1 z`ψi 2 μoNao 2 RT ln½Na1 o 2 z`ψo T he O rig in of the Res tin g Membra ne P oten tial EXAMPLE 3.1.1 Calculate the Equilibrium Potential for Na1 Inserting values for [Na1]i 5 12 3 1023 M and [Na1]o 5 145 3 1023 M, into Eqn [3.1.5], we calculate 145 3 1023 M ψi ψo 5 8:314 J mol21 K21 3 310 K 3 ln 5 0:0666 V 123 1023 M Note that the potential inside is positive, as it should be to impede further influx of a positive ion. 1 3 9:649 3 104 C mol21 Canceling out the standard free energies, which are equal, we can arrange Eqn [3.1.6] to give ½3:1:7 RT ½Na1 o ln 5 ðψi 2 ψo Þ z` ½Na1 i The Nernst equation is often written with log10 instead of the natural logarithm. At 37 C 5 310 K, Eqn [3.1.7] can be written as ½3:1:8 0:0615 log ½Na1 o 5 ðψi 2 ψo Þ ½Na1 i where the log is now log10. The term 0.0615 is the evaluation of the expression RT/z` and conversion of the natural log, ln, to log10. What this means is that every 10-fold gradient in concentration of a singly charged ion gives an equilibrium potential of 61.5 mV at 37 C. THE EQUILIBRIUM POTENTIAL FOR K 1 IS NEGATIVE Now we suppose that the membrane is impermeable to Na1 and Cl2 ions but it is permeable only to K1. The [K1] concentrations on the two sides of the membrane are those in a muscle cell, namely, [K1]o 5 4 3 1023 M and [K1]i 5 155 3 1023 M. Because of its concentration gradient, K1 will diffuse out of the cell, causing an outward current and accumulation of positive charges on the outside of the cell. The Nernst equation is ½3:1:9 RT ½K1 o ln 5 ðψi 2 ψo Þ z` ½K1 i The result of the calculation using [K1]o 5 4 3 1023 M and [K1]i 5 155 3 1023 M is EK 5 20.0977 V or 297.7 mV. Here the potential is strongly negative inside. In this case, z 5 21 because the Cl2 ion has a negative charge. This is equivalent to inverting the argument of the logarithm. Here we get ψi 2 ψo 5 20.0615 log (100/5) 5 20.080 V. Thus ECl 5 20.080 V 5 280 mV. Figure 3.1.2 shows the concentration differences for Na1, K1, and Cl2 and the equilibrium potential for each of these ions. It is important to remember what the equilibrium potential is. It is the potential at which the electrical force exactly balances diffusion so that the net flux of the ion across the membrane is zero. Since there is no flux, there is no change in the concentrations on the two sides of the membrane and the ion is at equilibrium. Such hypothetical membranes as the ones we have been considering that are permeable only to Na1, or only to K1 or only to Cl2 do not exist. Real membranes have some nonzero permeability to all of these ions. So what is the membrane potential across a membrane that is permeable to all three? The short answer is: it depends on how permeable the membrane is to each of the ions. What we need is some expression that tells us what the magnitude of the membrane potential will be, given the equilibrium potentials and the relative permeabilities of the ions. INTEGRATION OF THE NERNSTPLANCK ELECTRODIFFUSION EQUATION GIVES THE GOLDMANHODGKINKATZ EQUATION When there is a flux of solute, and the solute is charged, there is a current. The relationship between current density and flux is ½3:1:11 Ii 5 z`Ji We can repeat this calculation for any ion to obtain its equilibrium potential. We must remember what goes into the equations, however. As an example, consider that the membrane is not permeable to either K1 or Na1, but is freely permeable only to Cl2. Suppose further that Cl2 is 100 3 1023 M outside the cell and 5 3 1023 M inside the cell. The Nernst equation for Cl is where the subscript i denotes the particular ion that is carrying the current. The total current is the sum of all the ionic currents. Now we substitute in for the flux using Eqn [3.1.1] to get @Ci Di @ψ 1 Ci zi ` ½3:1:12 Ii 5 zi `Ji 5 2zi ` Di @x @x RT RT ½Cl2 o ln 5 ðψi 2 ψo Þ z` ½Cl2 i This is the NernstPlanck electrodiffusion equation. We can obtain an expression for Ii by integrating this equation over the thickness of the membrane (from ½3:1:10 257 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Membrane 160 150 [K+]i 155 mM [Na+]o 145 mM 130 [Cl–]o100 mM 0 100 90 80 70 –20 Extracellular fluid Cytosol 60 50 –40 –60 40 ECl = –80 mV 30 20 10 ENa = +66.7 mV +20 120 110 +60 +40 140 Concentration (mM) 258 [K+]o 4 mM 0 [Na+]i 12 mM [Cl–]i 5 mM –80 –100 EK = –97.7 mV FIGURE 3.1.2 Equilibrium potentials for Na, K, and Cl in a muscle cell. A positive equilibrium potential is needed to prevent Na1 ions from entering the cell from the extracellular fluid; a negative equilibrium potential is needed to prevent Cl2 ions from entering; a negative equilibrium potential is needed to prevent K1 ions from exiting the cell. The solid lines represent ion concentrations. x 5 0 to δ, where δ is the thickness of the membrane), but this integration can be accomplished in closed form only if we assume that the potential gradient is linear. This is called the constant field assumption. The integration gives ðzi `Em =RTÞ z2 `2 Di Ci e 2 Co Em Ii 5 i ½3:1:13 RTδ eðzi `Em =RTÞ 2 1 as detailed in Appendix 3.1.A1. This is often written in an alternate form by multiplying both numerator and denominator by eð2Z`=RTÞEm : ½3:1:14 Iion 5 ðD=δÞ½ðz2 `2 =RTÞEm ½Ci 2 Co eð2z`=RTÞEm ½1 2 eð2z`=RTÞEm This is the GoldmanHodgkinKatz (GHK) current equation. It relates the current carried by each ion to its concentration on both sides of the membrane (Ci and Co) and to the membrane potential, Em. Now at steady-state resting membrane potential, the total current across the membrane must be zero, or otherwise Em would be changing: ½3:1:15 Itotal 5 INa 1 IK 1 ICl 5 0 By substituting in for the expression for INa, IK, and ICl, we can derive an expression for Em: RT PK ½K1 o 1 PNa ½Na1 o 1 PCl ½Cl2 i ln ½3:1:16 Em 5 ` PK ½K1 i 1 PNa ½Na1 i 1 PCl ½Cl2 o This is the GHK equation. It describes the resting membrane potential when only Na1, K1, and Cl2 are permeant, but it can be expanded to include other ions. The inside concentration of Cl2 appears in the numerator with the outside concentrations of Na1 and K1 because z for Cl2 is 21.0 and z for Na1 and K1 is 11.0. Additional ions can be added to this equation if they contribute significantly to the currents across the membrane. This equation shows that the resting membrane potential results from the concentrationweighted permeabilities across the membrane because each permeability is multiplied by the concentration of the ion. Appendix 3.1.A1 presents a full derivation of this equation. SLOPE CONDUCTANCE AND CHORD CONDUCTANCE RELATE ION FLOWS TO THE NET DRIVING FORCE The GHK current equation (Eqn [3.1.14]) describes the current carried by any given ion in terms of its concentration on both sides of the membrane and the membrane potential. If we assume that the permeability, D/δ, is constant, we can calculate the current carried by each ion as a function of the membrane potential. The currents carried by K1 and Na1 for a muscle cell containing 155 mM [K1]i and 12 mM [Na1]i and 4 mM [K1]o and 145 mM [Na1]o as predicted from the GHK current equation are shown in Figure 3.1.3. At the equilibrium potential for each ion, there is no current. This equilibrium potential is also called the reversal potential, because at this point the current changes from negative (positive charges enter the cell—by convention this is an inward current) to positive (positive charges exits the cellthis is an outward current). T he O rig in of the Res tin g Membra ne P oten tial At high Em the IK approaches ohmic because [K+]i dominates the outward current and K+ o carries little inward current The slope of the chord connecting the curve to its reversal potential is the chord conductance 30 30 IK INa Slope of the curve at any Em 20 (dI/dEm) gives the slope conductance The reversal potential20 is the Em at which current falls to zero. For K this is –98 mV 10 10 Em 0 0 –10 –10 –20 –20 –30 –0.2 The dashed line is Ohm's law which occurs only –0.1 0.0 if 0.1 [K+]i = [K+]o = 155 mM 0.2 Em –30 –0.2 Em (V) Reversal potential for Na is +67 mV –0.1 0.0 0.1 0.2 Em (V) FIGURE 3.1.3 Currents carried by K1 (left) and Na1 (right) as predicted by the GHK current equation if the membrane was permeable only to K1 (left) or to Na1 (right). The relationship between current and voltage can be described by a conductance. Ohm’s law states ½3:1:17 Ii 5 E 5 gi E Ri where E is the potential difference that drives current flow (NOT the electric field intensity!), Ri is the resistance to the ion i, and gi is the conductance. Resistances have units of ohms. Conductances have units of ohm21, which is a siemen, equal to 1 AV21. According to Ohm’s law, the conductance is defined for a line that passes through the origin. The origin is the point at which there is no net driving force for current flow. For ions that are not uniformly distributed across the membrane, however, the point of no net current flow occurs at the reversal potential. Thus we define a chord conductance: ½3:1:18 Ii 5 gi ðEm 2 Ei Þ where gi is the chord conductance at Em, Em is the membrane potential at any point in the currentvoltage curve, and Ei is the equilibrium potential 5 reversal potential for a single ion, i. This conductance is the slope of the chord joining the curve to its reversal potential, as shown in Figure 3.1.3. Thus the chord conductance is not constant but varies with membrane potential, Em. We can also define a slope conductance that relates I to E. This is obtained by differentiating Eqn [3.1.17]: ½3:1:19 g5 dIi dEm Neither the slope conductance nor the chord conductance is a constant. Even for a straight cylindrical pore that is specific for some ion, the conductance varies with voltage because the current is carried only by one ion and its concentration is not the same on the two sides of the membrane. THE CHORD CONDUCTANCE EQUATION RELATES MEMBRANE POTENTIAL TO ALL ION FLOWS There are two ways to answer the problem of finding the resting membrane potential when the membrane is permeable to several ions. One way focuses on the permeabilities as we have already defined them, and a second focuses on the conductances, which are related, but not identical, to the permeabilities. We will use conductances here because it makes it easier to understand other electrical phenomena in cells. We begin with the fact that the total current at the resting membrane potential is zero. This is Eqn [3.1.15]: ½3:1:15 Itotal 5 INa 1 IK 1 ICl 5 0 Substituting in from Eqn [3.1.18] for the individual currents, we get ½3:1:20 I 5 gNa ðEm 2 ENa Þ 1 gK ðEm 2 EK Þ 1 gCl ðEm 2 ECl Þ At rest, I 5 0, and so we collect terms in Em on the lefthand side to find ½3:1:21 ðgNa 1 gK 1 gCl ÞEm 5 gNa ENa 1 gK EK 1 gCl ECl Solving for Em, we obtain 259 260 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION gNa gK gCl ENa 1 EK 1 ECl gNa 1gK 1gCl gNa 1gK 1gCl gNa 1gK 1gCl ½3:1:22 Em 5 This is the chord conductance equation. What it says is that the resting membrane potential is the conductance-weighted average of the equilibrium potentials for the ions that have any conductance across the membrane. We can clearly enlarge Eqn [3.1.22] to include other ions such as Ca21. Both the chord conductance equation and the GHK equation indicate that all permeant ions contribute to the resting membrane potential. Those ions with the largest conductances, or the largest permeabilities, are correspondingly greater determinants of the resting membrane potential. INa 5 gNa ð285 mV 2 66:7 mVÞ 5 2gNa 3 151:7 mV Which way does the Na1 ion go? Its concentration is higher outside the cell than inside. A potential of 166.6 mV inside would be necessary to stop the diffusive flow down its concentration gradient, but the resting membrane potential is 285 mV. Thus the negative membrane potential further drives Na1 influx. Na1 goes into the cell, but the current is negative, as indicated by the sign and the fact that gNa is always positive. The current is negative because that is our convention for determining the direction of current. An inward current (flux of a positive ion) is a negative current. From Eqn [3.1.11], we see that the current is in the same direction as the flux: ½3:1:11 THE CURRENT CONVENTION IS THAT AN OUTWARD CURRENT IS POSITIVE As mentioned earlier, the reversal potential is the potential at which the current reverses direction: it changes from positive to negative. It is important here to remember the sign convention for these currents. First, current is taken as the direction of positive charge flow. Second, a positive current is taken as an outward current—it goes from the inside of the cell to the outside. Third, a negative current is taken as an inward current. This convention becomes apparent when we consider the individual ionic currents. According to Eqn [3.1.18], the individual ionic currents are INa 5 gNa ðEm 2 ENa Þ ICl 5 gCl ðEm 2 ECl Þ Recall that g stands for a conductance. Conductances are always positive. We consider the example given above for the concentrations and equilibrium potentials. These values are recapitulated in Table 3.1.1. Suppose that the resting membrane potential is 285 mV. The Na current could then be calculated as TABLE 3.1.1 Concentrations of Ions and Their Equilibrium Potentials n [Ion]out 1 [Ion]in 23 Na 145 3 10 1 23 M K 4 3 10 Cl2 100 3 1023 M M Ei 23 12 3 10 M 23 155 3 10 5 3 1023 M M Thus inward Na flux is also a negative flux. The convention for positive flux is also from inside the cell to outside. In the case of K1, the current is given as IK 5 gK ð285 mV 1 97:7 mVÞ 5 gNa 3 8:7 mV which is positive. It takes 297.7 mV to stop K1 exit from the cell. The resting membrane potential of 285 mV is insufficient to stop K1 exit, so at rest there is some outward IK. Because it is an outward current, it is positive. The JK is similarly positive. In the case of Cl2, the current is ICl 5 gCl ð285 mV 1 80 mVÞ 5 2gCl 3 5 mV IK 5 gK ðEm 2 EK Þ ½3:1:23 Ii 5 zi `Ji 166.6 mV 297.7 mV 280.0 mV So the current is negative. Here 280 mV is enough to stop Cl entry into the cell. This is the ECl and you can see that if the membrane potential was 280 mV, ICl would be zero. But the membrane potential is 285 mV, which is more negative than ECl; thus the negative inside potential forces Cl2 out of the cell. There is an outward flux of Cl2, which is a positive flux. But the charge on Cl2, zCl 5 21 and so the current carried by Cl2 is opposite to its flux! (see Eqn [3.1.11]). The negative current is an inward current carried by the outward flux of Cl2. This current convention and the convention that membrane potential is defined as Em 5 ψi 2 ψo are true conventions in that the opposite conventions do not violate any physical law. These conventions are equivalent to the orientation of the x-axis perpendicular to the surface of the cell membrane. The convention is that x 5 0 is on the inside of the cell and x 5 δ, where δ is the thickness of the membrane, is on the outside surface. Figure 3.1.4 illustrates this convention and what it means for the gradients in C and ψ. T he O rig in of the Res tin g Membra ne P oten tial The concentration gradient is positive, so the diffusive force is negative or inward 160 + – 150 + – + – + – + – 140 130 [Na+]o145 mM Concentration (mM) 120 110 100 90 80 ψo = 0 mV + dC =- + dx - + – 70 + – 60 + – Extracellular fluid 50 + The voltage gradient is positive, so the40 electrical force is negative + or inward 30 20 10 0 dψ +40 Diffusive force +20 [Na]o – [Na]i 0 δ–0 Cytosol –20 Electrical force –40 –60 – – ψ i = –85 mV ψo – ψ Em = –85 mV – + i δ-0 dx ENa = +66.7 mV – + = +60 [Na+]i 12 mM –100 x=0 x =δ X-axis The concentration gradient is negative so the diffusive force is positive or outward + 160 150 + 140 + +60 – [K+]i 155 mM +40 – + – + + dC = dx - + – + – 70 + – 60 + – 50 + – + – 130 Concentration (mM) 120 110 100 90 80 ψo = 0 mV Extracellular fluid 40 30 dψ 20 dx 10 0 = ψ +o – ψi +δ – [K]o – [K]i +20 δ–0 Cytosol 0 –20 [K+]o 4 mM –60 –100 x =δ Electrical force –40 – ψ i = –85 mV – 0 Diffusive force Em = –85 mV EK = –97.7 mV x=0 X-axis FIGURE 3.1.4 Cartoon of the concentration and voltage gradients and their resulting parts of the ion flux. Top, situation with Na1 ions. The [Na1]o is about 145 mM, whereas the [Na1]i is about 12 mM. Thus the concentration gradient is positive (it slopes up with x; the direction of the x-axis is from inside to outside of the cell as indicated at the bottom of the figure). Fick’s law gives the diffusive flux as being proportional to the negative of the gradient. Thus the diffusive force favors a negative flux, which is directed inward. The outside potential, ψo, is taken as zero. At rest, ψi is about 285 mV. Thus the voltage gradient is also positive. The flux produced by the electrical force is also proportional to the negative of the electrical gradient. Thus the positive voltage gradient produces a negative flux, which is inward. Thus for Na1, both the diffusive force and electrical force produce an inward flux and an inward current. The bottom panel illustrates the situation with K1 ions. [K1]o is just 4 mM and [K1]i 5 155 mM, so the concentration gradient is negative. Thus the diffusive force is positive or directed outward. The negative electrical potential inside makes a positive electrical potential gradient, which favors a negative K1 movement, or an inward current and flux. Thus here the diffusive force and the electrical force oppose each other, but their magnitudes are not equal. Another way of looking at the equilibrium potential is that it is the magnitude of the diffusive force expressed in electrical terms. The net effect on movement is obtained by subtracting the two forces. 261 262 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION SUMMARY Separation of charges produces an electric field defined as the force experienced by a test positive unit charge. The field at any point is also characterized by a potential, defined as the work necessary to move a positive unit test charge from infinite separation to that point. The potential surrounding positive charge is positive, and that surrounding negative charge is negative. Thus the separation of charges across a membrane produces a potential difference between the two sides of the membrane. Ions move in response to concentration differences and in response to electrical forces. If a membrane is permeable only to a single ion, and the concentration of the ion is different on the two sides of the membrane, then the ion will diffuse from its high concentration toward its low concentration. Since it is charged, the ion movement makes a current and there is a separation of charge. This separation of charge produces an electrical force that opposes further diffusion. The electric field is the force per unit test charge, and this is the negative spatial derivative of the potential. The membrane potential at which diffusion is exactly balanced by the electrical forces is called the equilibrium potential. It can be calculated by using the Nernst equation: RT Co ln 5 ðψi 2 ψo Þ z` Ci The equilibrium potentials for K1, Na1, and Cl2 depend on their concentrations on both sides of the membrane and they are typically different from each other. In the Nernst equation, put the outside concentration in the numerator to agree with the sign convention. Real membranes are permeable to a variety of ions, but with different permeabilities. The permeability can be expressed either as a permeability relating flux to driving force in concentration units or as conductance relating current to driving force in voltage units. The resting membrane potential occurs at a net zero membrane current. This occurs at a membrane potential that is the conductance-weighted average of the equilibrium potentials, as described by the chord conductance equation: Em 5 gNa gK ENa 1 EK ð gNa 1 gK 1 gCl Þ ð gNa 1 gK 1 gCl Þ 1 gCl ECl ð gNa 1 gK 1 gCl Þ Thus the resting membrane potential is closest to the equilibrium potential of the ion with the highest conductance. The resting membrane potential also can be expressed in terms of the permeabilities: it depends on the concentration-weighted permeabilities as expressed in the GHK equation. Membrane potential is defined as Em 5 ψi 2 ψo. An outward flow of positive ions is taken as a positive current. The current for any ion is given as Ii 5 gi ðEm 2 Ei Þ where gi is the conductance to the particular ion, in siemens. This is a chord conductance which is distinguished from a slope conductance. Here Ei is the equilibrium potential for the ion in question. Conductances are always positive. A negative current means that the current is directed inward. REVIEW QUESTIONS 1. What is the current of an ion at its equilibrium potential? What would the current look like as a function of membrane potential if the ion had the same concentration on both sides of the membrane? 2. If the conductance of the membrane to K1 was increased, what would happen to the equilibrium potential for K1 (EK)? What would happen to Em? 3. Why is Clo in the denominator of the GHK equation when Nao and Ko are in the numerator? 4. What potential is calculated by the Nernst equation? 5. What potential is calculated by the GHK equation? 6. What is a siemen? 7. What is meant by an inward current? Outward current? What sign does an inward current have? If Cl2 exits the cell, does it make a positive or negative current? APPENDIX 3.1.A1 DERIVATION OF THE GHK EQUATION The flux of an ion through a water-filled channel should be given by the generalized Fick’s law given by Eqn [3.1.1]: ½3:1:A1:1 Js 5 2D @C D @ψ 2 Cz` @x RT @x The flux can be obtained by integrating this equation. To begin, we multiply both sides of the equation by an integrating factor, ρ, and we choose ρ so that the right-hand side of the equation becomes an exact differential: @C Cz`ρ @ψ 2 Js ρ 5 2D ρ ½3:1:A1:2 @x RT @x We choose ρ so that the terms in brackets are an exact differential. Thus we want ½3:1:A1:3 ρ @C Cz`ρ @ψ dðρCÞ @C @ρ 1 5 5ρ 1C @x RT @x dx @x @x From comparing the extreme left of Eqn [3.1.A1.3] to the extreme right, we see that the condition that multiplication by ρ transforms the equation into an exact differential is met if ½3:1:A1:4 @ρ z`ρ @ψ 5 @x RT @x T he O rig in of the Res tin g Membra ne P oten tial We rearrange this to get +60 @ρ z` 5 @ψ ρ RT ½3:1:A1:5 160 150 140 A solution to Eqn [3.1.A1.5] is 130 ρ 5 eðz`=RTÞψ 120 We may insert this result back into Eqn [3.1.A1.2] to obtain Js eðz`=RTÞψ 5 2D ½3:1:A1:7 dðCeðz`=RTÞψ Þ dx which may be rewritten as Js eðz`=RTÞψ dx 5 2DdðCeðz`=RTÞψ Þ ½3:1:A1:8 The flux can be obtained by integrating this equation from x 5 0 (one side of the membrane) to x 5 δ, the other side of the membrane for a membrane with thickness δ: ðδ ðδ ½3:1:A1:9 Js eðz`=RTÞψ dx 5 2D dðCeðz`=RTÞψ Þ 0 The numerator in this equation is the integral of an exact differential and can be immediately evaluated between the boundaries. This gives Js 5 2D½CðδÞeðz`=RTÞψðδÞ 2 Cð0Þeðz`=RTÞψð0Þ Ðδ ðz`=RTÞψ dx 0 e The denominator in this equation can be evaluated only if ψ(x) is known. However, generally ψ(x) is unknown. We can integrate the denominator if we assume that ψ is a linear function of x. This is true if the electric field between x 5 0 and δ is constant. Recall that the electric field intensity is the negative derivative of the potential. For this reason, the assumption of a linear potential gradient is called the constant field assumption. If the potential is linear, then we can write ½3:1:A1:12 ψ 5 ψð0Þ 1 ψð0Þ 2 ψðδÞ x 02δ The situation of C(0), C(δ), ψ(0), and ψ(δ) is illustrated in Figure 3.1.A1.1. The difference in potential across the membrane, ψ(0) 2 ψ(δ), is the the membrane potential Em. Here ψ(0) is the inside potential and ψ(δ) is the outside potential. Thus Eqn [3.1.A1.12] becomes ½3:1:A1:13 100 – + – + – + – + – dC ψo = ψ (δ ) = 0 mV + =– dx + 90 – 80 Extracellular fluid 70 + – + – 60 + – 50 + – 40 + – + – 30 20 dψ 10 = – ψ 5 ψð0Þ 1 Em x δ +40 +20 Co – Ci 0 δ – 0 Cytosol –20 –40 –60 ψi = ψ (0) Ci δ – 0 dx 0 ψ – +ψ o i –100 0 In this case, we limit ourselves to the steady-state condition. In this case, Js does not vary with distance across the membrane; it is constant. Therefore Js may be removed from the integral and we get Ðδ 2D 0 dðCeðz`=RTÞψ Þ ½3:1:A1:10 Js 5 Ðδ ðz`=RTÞψ dx 0 e ½3:1:A1:11 110 Concentration (mM) ½3:1:A1:6 Co + x=0 x =δ X–axis FIGURE 3.1.A1.1 Concentration and potential profile across the membrane along with the convention for the location of the X-axis. This result can be substituted into the denominator of Eqn [3.1.A1.11] to give ðδ ðδ ðz`=RTÞψ e dx 5 eðz`=RTÞðψð0Þ2ðEm =δÞxÞ dx 0 0 5e ðz`=RTÞψð0Þ ðδ e2ðz`=RTÞðEm =δÞx dx 0 ½3:1:A1:14 Evaluation of the integral gives eðz`=RTÞψð0Þ ðe2ðz`=RTÞEm 21Þ 2ðz`=RTÞðEm =δÞ ½3:1:A1:15 Inserting the result of Eqn [3.1.A1.15] back into Eqn [3.1.A1.11] Js 5 2D½CðδÞeðz`=RTÞψðδÞ 2 Cð0Þeðz`=RTÞψð0Þ ðz`=RTÞEm 21 m =δÞÞ½e ðeðz`=RTÞψð0Þ Þ=ð2ðz`=RTÞðE ½3:1:A1:16 This can be simplified to Js 5 ðD=δÞ½ðz`=RTÞEm ½CðδÞeðz`=RTÞψðδÞ 2 Cð0Þeðz`=RTÞψð0Þ e2ðz`=RTÞψð0Þ ½e2ðz`=RTÞEm 2 1 ½3:1:A1:17 Multiplying through by the exponent in the numerator, and recalling that Em 5 ψ(0) 2 ψ(δ), we convert Eqn [3.1.A1.17] into ½3:1:A1:18 Js 5 ðD=δÞ½ðz`=RTÞEm ½CðδÞe2ðz`=RTÞEm 2 C0 ½e2ðz`=RTÞEm 2 1 263 264 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION This equation is typically rewritten by multiplying numerator and denominator by 21, with no change in meaning: ðD=δÞ½ðz`=RTÞEm ½Cð0Þ 2 CðδÞe2ðz`=RTÞEm ½1 2 e2ðz`=RTÞEm ½3:1:A1:19 Js 5 This flux equation holds true for any ion including Na1, K1, and Cl2. The current carried by each ion is given by Eqn [3.1.19] by multiplying the flux by z`. We identify C(0) with the inside concentration of the ion, and C(δ) with the outside concentration, as described in Figure 3.1.A1.1 which of course are different for each of the ions. For each ion, we can write an equation for the current that has the form Iion 5 ðD=δÞ½ðz2 `2 =RTÞEm ½Ci 2 Co e2ðz`=RTÞEm ½1 2 e2ðz`=RTÞEm ½3:1:A1:20 We now insert the expressions for each of the individual currents. Here, however, we make an additional substitution that the diffusion coefficient for each ion divided by the thickness of the membrane, D/δ, is equal to the permeability of the membrane to each ion. The result is PNa ð`2 =RTÞEm ½Nain 2 ½Naout e2ð`=RTÞEm Itotal 5 0 5 1 2 e2ð`=RTÞEm PK ð`2 =RTÞEm ½Kin 2 ½Kout e2ð`=RTÞEm 1 1 2 e2ð`=RTÞEm PCl ð`2 =RTÞEm ½Clin 2 ½Clout eð`=RTÞEm 1 1 2 e ð`=RTÞEm ½3:1:A1:25 Solving this equation, we find 0 5 PNa ½½Nai 2 ½Nao e2ð`=RTÞEm Thus we can write ðD=δÞ½ð`2 =RTÞEm ½½Nai 2 ½Nao e2ð`=RTÞEm INa 5 ½1 2 e2ð`=RTÞEm ½3:1:A1:21 ½3:1:A1:22 IK 5 ðD=δÞ½ð`2 =RTÞEm ½½Ki 2 ½Ko e2ð`=RTÞEm ½1 2 e2ð`=RTÞEm ðD=δÞ½ð`2 =RTÞEm ½½Clin 2 ½Clout eð`=RTÞEm ½1 2 eð`=RTÞEm ½3:1:A1:23 1 PK ½½Ki 2 ½Ko e2ð`=RTÞEm ½3:1:A1:26 1 PCl ½½Clo 2 ½Cli e2ð`=RTÞEm where we have multiplied the numerator and denominator of the Chloride contribution to the current by exp (2`/RT Em) and rearranged the terms. Solving for Em, we finally obtain ICl 5 Note that the exponent in the numerator and denominator of ICl is positive instead of negative! This is because the z term in the exponent in Eqn [3.1.A1.20] is 21 for Cl. The total current across the membrane is the sum of the currents carried by each ion. At steady state (the resting membrane potential), the current is zero. Thus we can write ½3:1:A1:24 Itotal 5 INa 1 IK 1 ICl 5 0 ½3:1:A1:27 Em 5 RT ½PNa ½Nao 1 PK ½Ko 1 PCl ½Cli ln ` ½PNa ½Nai 1 PK ½Ki 1 PCl ½Clo This is the GHK equation. It tells us that the resting membrane potential (at which point the net current across the membrane is zero) is determined by the concentrations of ions on both sides of the membrane and by their respective permeabilities. The ion with the largest permeability dominates the membrane potential by moving the potential closer to its equilibrium potential. The Action Potential Learning Objectives G G G G G G G G G G G G G Draw a picture of a motor neuron and identify soma, dendrites, axon, myelin sheath, and terminals Describe what is meant by depolarization and hyperpolarization and what currents achieve it Explain what is meant by the “all or none” law Define latency, absolute refractory period, relative refractory period, and overshoot in an action potential on a nerve Define rheobase and chronaxie and how they can be determined from a strengthduration curve Recognize the Weiss relation between strength and duration Draw a graph showing the conductance changes with time during an action potential on a nerve Explain what is meant by the “activation gate” and “inactivation gate” of the Na channel Describe the state of the activation gate and inactivation gate during an action potential Describe how the inactivation gate is reset Distinguish between unitary current and ensemble current Using the whole-cell iv curve, explain the effect of small depolarizations below threshold and of depolarizations above threshold Explain why stronger stimuli need shorter durations to achieve an action potential CELLS USE ACTION POTENTIALS AS FAST SIGNALS Certain cells in the body are capable of initiating and propagating an action potential over their surface. These cells are called excitable cells and they include muscle and nerve cells. The action potential is a brief, pulse-like change in the membrane potential. Because it can be propagated rapidly over the surface of the cell, it conveys a fast signal from one place to another. We will use a motor neuron as an example of an excitable cell. THE MOTOR NEURON HAS DENDRITES, A CELL BODY, AND AN AXON Motor neurons are large cells in the ventral horn of the spinal cord as shown in Figure 3.2.1. They have a © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00024-0 3.2 number of processes called dendrites that bring signals to the motor neuron. The motor neuron also has one large process, the axon, that connects the motor neuron on one end with a muscle fiber on the other. Action potentials move along the axon so that activity in the motor neuron alters activity in the muscle. Axons from neurons can be myelinated or unmyelinated. Myelin refers to a sheath that covers the axon, but not entirely. In the peripheral nervous system, Schwann cells make the myelin by wrapping themselves around the axon, forming a multilayered structure of multiple cell membranes of the Schwann cell. In the central nervous system, oligodendroglial cells make the myelin. The sheath is not continuous in either the peripheral or central nervous system. At the end of each Schwann cell, there is a gap in the myelin. This gap is called the Node of Ranvier (see Figure 3.2.2). Like all cells, the motor neuron has a nucleus located in its cell body or soma. The soma is also sometimes referred to as the perikaryon from the Greek root “peri” meaning “around” or “surrounding” and “karyon” meaning “nut” or “kernel,” and referring to the nucleus. There is only one nucleus in the motor neuron and it is the site of mRNA transcription. PASSING A CURRENT ACROSS THE MEMBRANE CHANGES THE MEMBRANE POTENTIAL Figure 3.2.3 shows a highly schematic diagram of how the membrane potential can be measured by inserting a microelectrode through the membrane of the axon. The resting membrane potential in these axons is about 260 mV. Recall that the membrane potential is defined as ψi 2 ψo. It is possible to pass current across the membrane through the arrangements shown in Figure 3.2.3. In one case, the battery is hooked up so that current will pass into the cell. Recall here that current is defined as the direction of positive charge flow and an outward flow is positive. An inward current, shown in the middle of Figure 3.2.3, will depolarize the cell. The cell is already polarized; an inward current would make the cell less polarized and thus it would depolarize it. If the battery was connected with the opposite polarity, the resulting current would be outward and this would make the membrane more polarized. This is a hyperpolarizing 265 current. 266 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION AN OUTWARD CURRENT HYPERPOLARIZES THE MEMBRANE POTENTIAL When an outward current is passed across the membrane, the recorded membrane potential is a distorted version of the stimulus. As the magnitude of the current is increased (stimulus intensity is increased), the recorded Dorsal Dorsal root White matter hyperpolarization grows larger (see Figure 3.2.4). It takes some time for the current to reach a new steady-state membrane potential. In addition, the magnitude of the hyperpolarization depends on the distance from the stimulating electrode to the recording electrode. These phenomena are consequences of the electrical characteristics of the axon, its cable properties. In brief, the cable properties define a time constant and a length constant which describe how voltage builds up (or falls off) with time and with distance. The membrane parameters which influence these cable characteristics include the membrane resistance, the membrane capacitance, and the electrical resistance of the axoplasm. Gray matter THE RESULT OF DEPOLARIZING STIMULUS OF ADEQUATE SIZE IS A NEW PHENOMENON—THE ACTION POTENTIAL Motor neuron Ventral root Ventral FIGURE 3.2.1 Location of the motor neuron in the spinal cord. The spinal cord is shown in cross-section. The dorsal aspect is toward the back; ventral is toward the front. The dorsal and ventral roots are paired, with one on each side of the cord, but only one side is shown here. The motor neuron is shown in dark blue. The motor neuron’s cell body is located in gray matter in the ventral horn, and its long axon leaves the cord via the ventral root and continues on to a muscle where it makes a neuromuscular junction. These cells produce an action potential that propagates along the axon, excites the nerve at the neuromuscular junction, and conveys that excitation to the muscle in order to activate the muscle. Depolarization to a small degree produces a change in Em that mirrors hyperpolarization—the change is a distorted version of the stimulus. With larger depolarization, however, a new phenomenon arises. When Em exceeds a threshold value, there is an abrupt rise in Em, reaching positive values near 130 mV. Just as quickly, Em returns to near normal values. This abrupt change in the membrane potential brought about by depolarization is the action potential (see Figure 3.2.5). Dendrites Axon terminals Node of Ranvier Soma Axon Myelin sheath FIGURE 3.2.2 Parts of the motor neuron. Dendrites are multiple processes of the neuron that bring signals to the cell body, or soma. A single long axon exits the cell on one pole and reaches all the way to its target cell, the muscle fiber. The long axon is covered by a myelin sheath made by Schwann cells. The sheath is interrupted at regular intervals at the nodes of Ranvier. Negative (inward) current depolarizes the membrane Positive (outward) current hyperpolarizes the membrane Battery + − Battery − + Voltmeter FIGURE 3.2.3 Arrangement of electrodes to record membrane potential (left); to inject a depolarizing current (middle) or to pass a hyperpolarizing current (right). IM refers to “membrane current,” or current across the membrane. + − IM IM ++++++++++++++++++++++++++++++++++++++++++++++++++++++++++ −−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−− Axoplasm The Action P otential THE ACTION POTENTIAL IS ALL OR NONE When the strength of depolarization is increased further, the resulting wave form in Em remains remarkably constant. That is, if the depolarization is enough to trigger an action potential, the resulting action potential is independent of the stimulus. The action potential is “all or none.” The action potential is not graded, meaning that it does not vary from a minimum to a maximum; rather, a given stimulus will either produce an action potential or it will not, and the resulting shape of the plot of Em against time is approximately the same. THE LATENCY DECREASES WITH INCREASING STIMULUS STRENGTH +60 +40 Once initiated, the action potential moves away from its point of origin, and this is called propagation. At some point away from its origin, the action potentials are indistinguishable. However, near the stimulus there is a slight difference, having to do with the time it takes to reach threshold, the latency. The membrane potential is not a completely faithful representation of the stimulus: the rise in Em lags behind the stimulus rise and the fall in Em lags behind the stimulus fall. The membrane distorts the stimulus. When the stimulus strength is increased above threshold, the time to reach threshold, the latency, decreases. The relationship between latency and stimulus strength is shown in Figure 3.2.6. +20 0 Em (mV) −20 −40 −60 −80 Stimulus intensity 0 1 2 3 4 5 6 7 Time (ms) 8 9 10 FIGURE 3.2.4 Results of hyperpolarizing stimuli of varying intensity on the axon membrane potential, Em. The effect of hyperpolarizing stimuli on the membrane potential is a distorted version of the stimulus wave form. The rise time and fall time of Em are part of the cable properties of the axon. THRESHOLD IS THE MEMBRANE POTENTIAL AT WHICH AN ACTION POTENTIAL IS ELICITED 50% OF THE TIME The threshold for a nerve is the membrane potential which must be reached in order to “fire” an action Myelinated squid axon Rabbit motor neuron +60 Absolute refractory period +40 +60 +20 +40 0 Em (mV) −20 Em Relative refractory period Overshoot +20 0 (mV) −40 –20 −60 –40 −80 –60 Hyperpolarizing after potential –80 Stimulus intensity Stimulus intensity Stimulus–response latency 0 1 2 3 4 5 6 7 Time (ms) 8 9 10 FIGURE 3.2.5 Results of depolarizing stimuli of varying intensity on the axon membrane potential. At low stimulus strength, Em is a distorted version of the stimulus, as in the case of hyperpolarization. When the depolarization exceeds a threshold value, there is an abrupt rise in Em, reaching positive values. This abrupt change in the membrane potential brought about by depolarization is the action potential. 0 1 2 3 4 5 6 7 8 Time (ms) 9 10 FIGURE 3.2.6 Effect of stimulus strength above threshold on resulting wave forms in a myelinated squid motor neuron axon. Note that the wave form here differs somewhat from that in mammalian motor neurons. It is shorter in the squid and shows a hyperpolarization after potential that is largely lacking in mammalian motor neurons. 267 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION potential. Operationally, it is defined as the membrane potential at which the nerve will fire an action potential 50% of the time. However, it is possible to initiate an action potential even when the membrane potential is below threshold, but the probability of this happening is much reduced. Similarly, it is possible to fail to initiate an action potential even when the membrane potential is above threshold, but with a much reduced probability. The probability of initiating an action potential is a steep function of the membrane potential. THE NERVE CANNOT PRODUCE A SECOND EXCITATION DURING THE ABSOLUTE REFRACTORY PERIOD While the first action potential is occurring, it is impossible to begin a second action potential, no matter how powerful the second stimulus. The time during which the nerve is refractory to a second stimulus is called the absolute refractory period. It typically lasts for 12 ms. Following the absolute refractory period is a second, relative refractory period. This typically lasts some 4 ms or so, and during this time it is possible to stimulate the nerve cell to make another action potential, but it is more difficult to do so than in the resting neuron. That is, the threshold is elevated during the relative refractory period. THE ACTION POTENTIAL REVERSES TO POSITIVE VALUES, CALLED THE OVERSHOOT The first recordings of action potentials with intracellular electrodes were accomplished in 1939 and 1940 by A.L. Hodgkin and Andrew Huxley and by K.S. Cole and H.J. Curtis, respectively, using the squid giant axon. They found results similar to those shown in Figure 3.2.6. In particular, they found that the membrane potential not only went to zero, but it took on positive values. This is called the overshoot. Given the fact that the equilibrium potentials for all known ions were negative except for Na1 and Ca21, these pioneers suggested that the membrane could have a positive potential by becoming selectively permeable to Na1. THE STRENGTHDURATION RELATIONSHIP IS HYPERBOLIC The relationship between the current necessary to reach threshold and the duration of the current for a compound action potential is shown in Figure 3.2.7. This is related to the strengthlatency relationship, as the latency is the time required to elicit an action potential for a given strength of stimulus. The strengthduration relationship shown in Figure 3.2.7 asks the question, for a given strength of stimulus, how long must it continue to produce an action potential? For some strengths of stimulus, no duration is sufficient—you never get an action potential. For a critical strength of stimulus, you get an action potential only if the stimulus lasts infinitely long. This strength of stimulus is 50 40 Threshold current (mA) 268 30 20 10 0 0 0.2 0.4 0.6 Stimulus duration (ms) 0.8 1.0 FIGURE. 3.2.7 Strengthduration relationship in human peripheral nerve. The median nerve was stimulated using surface electrodes 1 cm in diameter taped to the skin over the median nerve at the wrist, 4 cm apart oriented along the course of the nerve. Stimulus was a square wave with rise and fall times of 10 μs. The antidromic compound sensory action potential was recorded at the index finger using ring electrodes with 23 mm diameter set 23 cm apart on the finger. Stimulus intensity was reduced 25% until the amplitude of the compound action potential (the aggregate of a bundle of axons) was reduced to 30% of maximum (From Mogyoros, I, Kiernan, MC, and Burke, D. Strengthduration properties of human peripheral nerve. Brain 119:439447, 1996). called the rheobase (see Figure 3.2.8). Weiss described this relationship by the equation: ½3:2:1 Q 5 Irh ðt 1 τSD Þ where Q is the charge, I is the rheobase current, t is the time the current is on, and τ SD is a strengthduration time constant, often called the chronaxie. Because the total charge is the current times the time, this equation becomes ½3:2:2 I 5 Irh ðt 1 τSD Þ τSD 5 Irh 1 Irh t t This last equation describes a rectangular hyperbole offset from the x-axis by Irh. Also, when t 5 τ SD, the current I 5 2 Irh; the current is twice the rheobase. These relationships are shown diagrammatically in Figure 3.2.8. We will seek to further understand this relationship after we establish how depolarization induces an action potential. VOLTAGE-DEPENDENT CHANGES IN ION CONDUCTANCE CAUSE THE ACTION POTENTIAL INCREASE IN ION CONDUCTANCE BEGINS AFTER THE MEMBRANE BEGINS TO DEPOLARIZE K.S. Cole and H.J. Curtis in 1939 studied the impedance properties of the squid giant axon and discovered a marked decrease in the impedance (equivalently, an increase in the conductance) of the squid axon The Action P otential 140 Current calculated as I = Irh (t + τSD) / t with Irh = 10 mA and τSD = 0.2 s Threshold current (mA) 120 The chronaxie is the time at which threshold current is twice the rheobase The rheobase, Irh, is the current that requires infinite time to elicit an action potential 100 80 60 τSD 40 20 Irh 0 0.0 0.2 0.4 0.6 Stimulus duration (ms) 0.8 membrane during the upstroke of the action potential. This increase begins only after the membrane potential rises many millivolts above the resting membrane potential. They argued that the foot of the action potential resulted from the discharging of the membrane from local currents from elsewhere in the cell. At the inflection point on the rising phase of the action potential, the cell generated its own net inward current. Such a current must be carried by an ion, and the most likely candidate is Na1, because of its high extracellular concentration. THE ACTION POTENTIAL IS ACCOMPANIED BY NA 1 INFLUX To test the idea that increases in Na1 conductance might cause the action potential, Alan Hodgkin and Bernard Katz replaced some of the NaCl in seawater with choline chloride. Replacing the Na1 reduced the upstroke of the action potential and markedly reduced the size of the action potential. Later experiments using radioactive tracers showed that Na1 influx accompanies the action potential. THE CHORD CONDUCTANCE EQUATION PREDICTS THAT CHANGES IN CONDUCTANCE WILL CHANGE THE MEMBRANE POTENTIAL In Chapter 3.1, we developed the chord conductance equation that shows that the resting membrane potential is the conductance-weighted average of the equilibrium potentials for all ions (see Eqn [3.2.3]). At rest, the membrane potential is closer to the K1 equilibrium potential because conductance K1 is higher than the conductance of the other ions. Em 5 ½3:2:3 gNa gK ENa 1 EK ðgNa 1 gK 1 gCl Þ ðgNa 1 gK 1 gCl Þ gCl ECl 1 ð gNa 1 gK 1 gCl Þ 1.0 FIGURE 3.2.8 Definition of rheobase and chronaxie. The rheobase current is the current which takes an infinite time to elicit an action potential—it is obtained by extrapolation or curve fitting. The chronaxie is the time at which threshold current is twice the rheobase. These two parameters are parameters of Eqn [3.2.2] that can be obtained by fitting the equation to experimental data. In the figure, the current was calculated according to Eqn [3.2.2] using a rheobase of 10 mA and a chronaxie of 200 ms. Note that the calculated curve resembles the form of the experimental curve shown in Figure 3.2.7. During the rising phase of the action potential, the conductance to Na1 increases, changing Em from its resting, polarized value toward more positive values. If the conductance to Na1 becomes large enough, relative to the conductances for K1 and Cl2, then Em will be driven toward the equilibrium potential for Na1, ENa. Since ENa is positive, Em will be driven to positive values and will exhibit the overshoot. GNA INCREASES TRANSIENTLY DURING THE ACTION POTENTIAL; GK INCREASES LATER AND STAYS ELEVATED LONGER The results described earlier show that the rapid depolarization and overshoot in the action potential is due to a transient increase in membrane conductance, and this is accompanied by an Na1 influx. The rapid repolarization of the membrane afterward is due to shutting off the increased Na1 conductance and increasing the K1 conductance. These results are consistent with the chord conductance equation. Hodgkin and Huxley succeeded in calculating gNa and gK during different parts of the action potential in the squid axon, and their results are summarized schematically in Figure 3.2.9. Although these results were determined in the squid, the principles remain the same for action potentials in mammalian excitable cells. CONDUCTANCE AND EQUILIBRIUM POTENTIALS FOR NA 1 AND K 1 ACCOUNT FOR ALL OF THE FEATURES OF THE ACTION POTENTIAL The origin of the action potential can be explained on the basis of the equilibrium potential for Na1 and K1 and the time course of their conductances. In the squid axon, the resting potential is on the order of 260 mV, with EK about 275 mV. When the membrane is depolarized by some means, a conductance pathway for Na1 begins to open. At this time, Na1 is far away from its equilibrium 269 270 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Em +60 ENa 30 +40 +20 Ionic gNa 0 Em (mV) –20 20 conductance gK 10 –40 (ms cm–2) –60 –80 FIGURE 3.2.9 Changes in gNa and gK during the propagated action potential as calculated by Hodgkin and Huxley. EK 0 1 2 potential, so the driving force for Na1 is large and the opening of the conductance pathway causes a Na1 influx and a further depolarization of the membrane. This additional depolarization further opens additional conductance pathways. This positive feedback explains the explosive increase in Na1 conductance and the rapid increase in the membrane potential toward ENa. 3 4 5 6 Time (ms) G The pathway for Na1 conductance next undergoes a time-dependent inactivation. This by itself reduces the inward Na1 current, except that the inactivation is accompanied by a reduction in Em (i.e., a repolarization of the membrane) and a subsequent increase in the driving force. The Na1 current is complicated, being the product of the conductance and the net driving force (see Eqn [3.1.18]). Repolarization of the membrane would occur with a return of Na1 conductance to normal, but it is hastened by a delayed increase in gK followed by its gradual return to resting levels. The increased gK increases an outward current carried by K1, and this current more quickly repolarizes the cell. Because gK remains elevated even when gNa has returned to normal, the cell is hyperpolarized for several ms following the action potential until gK returns to normal. G GNA IS A FUNCTION OF A NA 1 SELECTIVE CHANNEL G The identification of two major components, INa and IK, in the ionic currents by Hodgkin and Huxley, has allowed electrophysiologists to characterize the gating properties of the channels that pass these currents. The present view is that the Na1 channel has at least four distinct components: G The Selectivity Filter Enables the Na1 Channel to Pass Na1 Preferentially The relative permeabilities of ions in ion channels can be calculated by measuring the reversal potential when the ion composition is changed. The results show that the Na1 channel actually conducts H1 ions much more readily than it does Na1 ions. The current through the Na1 channel, however, is dominated by Na1 because of its much higher concentration. The [H1] is around 1027 M, whereas [Na1] is around 0.1 M, or 106-fold higher than [H1]. It is 7 8 0 9 10 believed that the selectivity relies on a combination of hydrated ion size and the free energy of hydration. Most other ions have a lower permeability through the Na1 channel. Thus the Na1 channel is selective for Na1. The Na1 Channel Possesses an Activation Gate At rest the membrane has a low gNa because the Na1 channel is blocked by its activation gate. This is a part of the Na1 channel that can be moved to allow Na1 to conduct through the channel. The opening of this gate is voltage dependent: the activation gate begins to open only when the membrane depolarizes. This voltage-dependent opening of the activation gates causes the explosive increase in Na1 conductance in the rising phase of the action potential. The Na1 Channel Possesses an Inactivation Gate Na1 channels also inactivate. This is the function of a separate gate that closes according to time and voltage. When it closes, the Na1 channel is nonconducting and, most importantly, it is inactivatable. Conductance through the Na1 channel requires that both the inactivation gate and the activation gate are open. When the inactivation gate is closed, opening of the activation gate alone does not allow the channel to conduct ions. Specific Toxins Bind to the Na1 Channel The puffer fish produces a potent toxin, tetrodotoxin, or TTX, that binds to Na1 channels and blocks them. As a result, action potential conduction in nerve and muscle is blocked, with lethal consequences. Another natural toxin, saxitoxin, has similar properties. It derives its name from the Alaskan butter clam, Saxidomus. Eating a single contaminated shellfish can be fatal. The Na1 channel contains regions that bind these toxins. THE INACTIVATION GATES MUST BE RESET BEFORE ANOTHER ACTION POTENTIAL CAN BE FIRED An action potential can occur in a nerve only when the Na1 channels can open. If they are blocked, for example, by tetrodotoxin or saxitoxin, then no action potentials are possible. Similarly, if the channels are in the inactivatable state because the inactivation gate is The Action P otential Na+ channel + Selectivity filter + + TTX binding site + – – Activation gate – Inactivation gate – TTX Extracellular fluid + – + – Cytosol TEA – – + filter Selectivity + K+ channel Activation gate – + – + – FIGURE 3.2.10 Hypothetical and conceptual model of the voltage-gated Na1 channel and K1 channel. The selectivity filter allows Na1 to pass but not K1 (top). When TTX (black circle) binds to its site, the channel is blocked. Closure of either the activation gate or inactivation gate will block Na1 conductance. The K1 channel also possesses an activation gate and is blocked by tetraethylammonium (TEA). closed, they also cannot be activated by opening the activation gates. Following a normal action potential, the Na1 channels are reset when the activation gates close and the inactivation gates open. This takes some time. Part of the refractoriness of the nerve cell immediately following the action potential is due to the lower conductance of the membrane to Na1 because the inactivation gates are closed. Figure 3.2.10 shows a conceptual model of the Na1 and K1 channels with their activation and inactivation gates and their binding sites for specific blockers. Figure 3.2.11 shows how these channels change during the action potential. Figure 3.2.12 shows the timing of opening of the activation and inactivation gates and that Na1 entry is possible only when both are open. CONDUCTANCE DEPENDS ON THE NUMBER AND STATE OF THE CHANNELS Figure 3.2.11 shows a cartoon of the states of the Na1 and K1 channels in axon membranes that give rise to the action potential. Out of necessity, only representative channels are shown in the figure. They are meant to represent what a population of channels is doing. The overall current carried through a patch of membrane is given by ½3:2:4 Ii 5 NPo i where I is the current, N is the number of channels, Po is the probability of a channel being open, and i is the unitary current, the current carried by the open channel under physiological conditions. Cells possess thousands of channels, each of whose behavior is stochastic, meaning that their opening and closing are not deterministic but probabilistic. The average behavior of many channels or the behavior of a large number of them is predictable, but the behavior of a single channel appears to be erratic and unpredictable. The sum of the currents of a population of channels is called the ensemble current. PATCH CLAMP EXPERIMENTS MEASURE UNITARY CONDUCTANCES It is possible to study the behavior of single channels using a patch clamp technique, shown in Figure 3.2.13. In this method, one clamps the voltage across the patch and measures current across it at that voltage. One can also step the potential from some holding value to a new value and measure the currents. Figure 3.2.14 shows the unitary Na1 and K1 currents in successive sweeps from patch clamp recordings in neuroblastoma cells (left) and squid giant axon (right). Distinction can be made between INa and IK by judicious choice of ionic conditions and use of specific inhibitors. The Na1 channels open briefly upon depolarization and then do not open later on. Note that the individual channels open and close rather erratically. It is not possible to predict exactly when a particular channel will either open or close. The presence of a large number of channels, however, will smooth out these discontinuities. The ensemble average is obtained by averaging many sweeps. This describes the average behavior of a single channel over many experiments, which is equivalent to the average behavior of a population of channels in a single sweep. The average behavior shows that Na1 channels open upon depolarization and then close, or inactivate. K1 channels, on the other hand, open after a slight delay and stay open during depolarization. Repolarization closes the K1 channels. THE CURRENTVOLTAGE RELATIONSHIP FOR THE WHOLE CELL DETERMINES THE THRESHOLD FOR EXCITATION Now that we know the ionic basis of both the resting membrane potential and the action potential, we are in a better position to understand why there is a critical level of depolarization. The whole-cell currentvoltage relationship in a cell is approximated by the curve shown in Figure 3.2.15. This curve results mainly from the iv relationship shown for the K1 channels and for the Na1 channels shown in Figure 3.1.3. These currents were for the open channels, and the whole-cell iv relationship is not just a sum of these, but also reflects the probability of the channel opening at the particular membrane potential. The resulting curve has negative currents at hyperpolarized membrane potentials, positive current at slight depolarizations, and negative currents at further depolarizations. Recall our convention for current sign: a positive current is an outward current, 271 272 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION 3 2 1 – + + – + + – Cytosol – + – ECF + – – + – Cytosol – + – + Cytosol ECF – + Na+ + Na+ channel Na+ channel Na+ channel ECF +40 mV –55 mV –70 mV Na+ Na+ – – + K+ + K+ – FIGURE 3.2.11 Proposed conceptual changes in the Na1 and K1 channels that give rise to the action potential. At rest (1), fluxes through both Na1 and K1 channels are small and the membrane is polarized. The Na1 activation gate is closed and the inactivation gate is open. The K1 activation gate is closed. During threshold depolarization (2), the Na1 activation gate opens and inward Na1 current further depolarizes the cell, leading to the upstroke of the action potential and the overshoot. Inward Na1 slows as Na1 approaches its equilibrium potential. Meanwhile, the K1 activation gate opens, leading to an outward K1 current (3). This outward K1 current repolarizes the cell and the Na1 inactivation gate closes (4). The continued outward K1 current may lead to a hyperpolarization. The Na1 activation gate closes (5). Upon repolarization, the K1 activation gate closes and the Na1 inactivation gate opens, resetting to the resting condition. The diagram here shows a single channel, but membrane conductance is governed by an ensemble of channels whose states are not necessarily identical. + – + – K+ channel + – + – + – channel – – – + – + – Cytosol ECF + 6 Na+ channel – channel –70 mV 5 Na+ channel + K+ –80 mV 4 ECF K+ + –50 mV + – + Na+ channel + – + – + + + – Cytosol – + – Cytosol – + – + – ECF Na+ Na+ + + + – + – – – + – + – K+ K+ – + – + K+ channel – + – and current is in the direction of positive ion flow. Thus a positive current removes positive ions from the cell, which hyperpolarizes the cell. Thus at membrane potentials lower than the resting membrane potential there is a negative total current carried by Na1 ions into the cell, the result being a depolarization back toward the resting membrane potential. The total current is zero at the resting membrane potential. At the resting membrane potential, the total iv curve crosses the x-axis. At slight depolarizations, the current is positive, meaning that positive ions exit the cell and cause a return towards the resting membrane potential. At the uniform threshold, the iv curve again crosses the x-axis. At this point, the current shifts from positive to negative. A slightly more positive + – + K+ channel – – + – + – + – + K+ channel – – + – + membrane potential shifts the current from positive, which returns the membrane to the resting membrane potential, to negative, caused by influx of Na1, and the membrane depolarizes. This depolarization results in progressively more negative currents. This is the influx of Na1 ions that constitutes the rising phase of the action potential. The point where the iv curve crosses the x-axis on the high potential side is the “uniform threshold,” meaning that if the membrane was uniformly depolarized to this point, an action potential would ensue 50% of the time. It turns out that the membrane is seldom uniformly depolarized: the potential varies with distance from the point of excitation. As a consequence, the actual threshold is generally higher than the uniform threshold. The Action P otential Upstroke of the action potential Resetting of the resting condition Repolarization of the membrane Rest A I Time 1 FIGURE 3.2.12 Timing of the opening of the Na channel activation gate (A) and inactivation gate (I). The activation gate is closed at rest while the inactivation gate is open. The activation gate opens briefly during excitation, and Na1 can cross the membrane because both gates are open. The inactivation gate closes, partly contributing to repolarization of the membrane. This first closes the activation gate, followed by a later resetting of the inactivation gate to the open state. Blue indicates closed state; white indicates open state. Pull Suction Cell 1. Polished glass patch clamp makes contact with the cell membrane, forming a high resistance seal 2. Gentle suction seals the membrane to the patch clamp pipette 3. Pulling the patch clamp pipette away breaks the membrane and forms the excised patch. Current measured from the patch pipette inside to the outside must go through the patch FIGURE 3.2.13 Method for making an excised cell patch. A polished glass microelectrode is brought near to a cell while positive pressure is applied to keep the microelectrode contents uncontaminated by the cell bathing solution. As the microelectrode approaches the cell, the resistance increases, indicating close approach. Positive pressure is turned off and the microelectrode advances to make contact with the membrane. Clean membranes will form a high-resistance seal. Suction is then applied and the microelectrode is reversed to pull off a small patch of membrane. In this configuration, any current that passes across the microelectrode tip must pass through the patch. In this way, single channel currents can be measured. The patch may have none, one, two, or several channels. Experiments can also be performed using cell-attached patches. THRESHOLD DEPOLARIZATION REQUIRES A THRESHOLD CHARGE MOVEMENT, WHICH EXPLAINS THE STRENGTHDURATION RELATIONSHIP Depolarization to threshold requires enough charge to change the voltage across the membrane to the threshold voltage. When stimulation is at a point, the threshold voltage is actually higher because the negative current in the patch of membrane above threshold is partially offset by positive currents in nearby membrane that is below threshold. The condition for threshold is that the total current of the cable is inward. The area of membrane that supplies inward current must be large enough so that outward currents supplied by the rest of the membrane are counterbalanced. The minimum length of fiber that must be depolarized to threshold is called the liminal length. This idea is shown diagrammatically in Figure 3.2.16. 273 274 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Voltage clamp record shows the holding potential of –130 mV was stepped to –40 mV at time zero in neuroblastoma cells Holding potential of –100 mV was stepped to +50 mV in squid giant axons +50 mV Sweeps show long openings and outward currents carried by K+ –40 mV –100 mV Individual sweeps show discrete channel openings –130 mV Downward deflection is an inward current. Here it is carried by Na+ 2 pA 5 ms 1 pA 0.5 pA 0 5 10 Time (ms) 15 20 Ensemble average shows opening early after depolarization and later closing of the Na+ channel 0 5 10 Time (ms) 15 20 Ensemble average shows delayed opening and continued conductance until membrane is repolarized FIGURE 3.2.14 Patch clamp currents of Na1 channels in neuroblastoma cells (left) and K1 channels of squid giant axon (right). (Left) Adapted from data of C.M. Baumgarten, S.C. Dudley, R.B. Rogart and H.A. Fozzard, Unitary conductance of Na1 channel isoforms in cardiac and NB2a neuroblastoma cells. Am. J. Physiol. 269: C1356C1363, 1995. (Right) Adapted from data of F. Bezanilla and C.K. Augustine cited in B. Hille, Ionic Channels in Excitable Membranes, Sinauer, 1992. THE AMOUNT OF CHARGE NECESSARY TO REACH THRESHOLD EXPLAINS THE STRENGTHDURATION RELATIONSHIP Figure 3.2.16 indicates that depolarization to threshold requires movement of sufficient charge according to the capacitance of the membrane: ΔV 5 Δq/C. Thus reaching threshold (a ΔV from the resting potential) requires a defined amount of charge movement. This explains the inverse relationship between stimulus strength (its current) and the duration, according to the Weiss Equation (see Eqn [3.2.1]). SUMMARY Resting nerve cells are polarized with a negative resting membrane potential caused by greater K1 conductance in the resting cell. Application of an outward current further polarizes the membrane, and the recording membrane potential is a distorted version of the stimulus. Application of an inward current depolarizes the membrane. If depolarization reaches threshold, nerve cells fire an action potential. The action potential is a brief, pulselike change in the membrane potential which can move from one area of the cell membrane to another and so it can be used to signal distant parts of the neuron. The Action P otential Depolarization above resting membrane but below threshold causes a positive current that depolarizes the membrane back towards the resting potential +1 Depolarization above threshold results in a negative current—an inward flow of + ions—that further depolarizes the membrane, leading to the action potential 0 Resting membrane potential Current -1 Uniform voltage threshold –2 Hyperpolarization below resting potential causes negative currents—an inward flow of + ions that depolarize the membrane back to the resting potential –3 –4 –5 –100 –80 –60 –40 –20 0 20 40 60 80 Membrane potential (volts) FIGURE 3.2.15 Whole-cell currentvoltage relationship. Positive currents above the equilibrium potential for K1 are mainly due to K1 exit, mainly due to the increased driving force for K1. Negative currents are mainly due to Na1 entry, due to reduction in the driving force for K1. Hyperpolarization below the resting membrane potential causes a negative current (inward flow of positive ions) that depolarizes the cell back towards the resting membrane potential. Slight depolarizations produce a positive current (outward flow of positive ions) that repolarizes the cell back towards the resting membrane potential. Thus slight hyperpolarization or depolarization returns the cell to rest. Larger depolarizations cause a negative current due to inward flow of Na1 ions that further depolarize the cell in the action potential. +1 0 Current –1 –3 +1 –4 0 Outward current in nearby slightly depolarized patch tends to limit depolarization –1 Current –2 –2 –5 –100 –80 –60 –40 –20 0 20 40 60 80 Membrane potential (volts) –3 Depolarizing, inward current –4 –5 –100 -80 + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + Membrane potential (volts) + + + + + + + + + + + + + + + area - - Less +depolarized + + + + + + + + + + + + + + + + + + Slightly depolarized area + + + + -- ++ ++ ++ + ++ + ++ + + ++ ++ ++ + + ++ + + - + + + + + + + + + + + + + + + + - + + + + + + + + + + + + + + + + + + + + + + + + - + + +movement + + + within + -- + +Charge + + + + + + + + + + + + depolarization + reduces fiber + - - - + +the + + + + of current injection+ + + + + + + + + point at the + + + + + + + +++ ++ + + + + + + ++ + + + + + + + + + + + + -60 -40 -20 0 20 40 60 80 + FIGURE 3.2.16 Consequence of spatial differences in membrane potential. Excitation occurs at a location in the membrane that causes a depolarization. This depolarization is sufficiently large to produce a negative or inward current that would further depolarize the cell. However, nearby patches of membrane are depolarized less and they are located on the part of the iv curve that carries a positive current. This positive current lowers the depolarization at the point of excitation. 275 276 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Action potentials are all or none: you either get one or you don’t. Increasing stimulus strength shortens the delay before the start of the action potential, but it does not alter its peak or its duration. The threshold is the potential at which an action potential is triggered 50% of the time, but this depends on the rate of depolarization. For some period of time after the start of an action potential, nerve cells will not begin another action potential. This period is the absolute refractory period. The relative refractory period follows the absolute refractory period. During the relative refractory period, the threshold for initiating a second action potential is higher. During the action potential, the membrane potential overshoots zero and becomes positive. Since only Na1 and Ca21 have positive equilibrium potentials, the membrane potential can become positive only if the permeability to one of these increases. The ionic currents giving rise to the action potential for squid giant axons were described by Alan Hodgkin and Andrew Huxley, who received the Nobel Prize for the work in 1963. They developed a kinetic model that explained the action potential in terms of a single voltage-gated K1 channel and a Na1 channel governed by voltagedependent and time-dependent activation and inactivation gates. During rest, the Na1 inactivation gate is open and the activation gate is closed. Upon depolarization, the activation gate opens and Na1 conductance increases markedly. This causes an inward current (carried by Na1) that further depolarizes the membrane and causes more Na1 channels to open. The explosive increase in gNa causes the rising phase of the action potential. With time, the Na1 inactivation gate closes, the K1 activation gate opens, and the Na1 activation gate closes. Opening of the K1 activation gate and closing of the Na1 inactivation gate cause an outward current that repolarizes the membrane. With further time, the Na1 inactivation gates reopen and this reestablishes the resting condition. The squid axon is presented here as an example of how to think about excitable cells. Each excitable cell, and in fact each part of an excitable cell, can have different sets of channels with distinct activation and inactivation properties. The height of the action potential and its duration depend on these characteristics, as well as the electrical characteristics of the cell and the ionic conditions inside and outside the cell. Channels have discrete states with discrete conductances. Each channel undergoes transitions between conducting and nonconducting or subconducting states. Such discontinuous conductance changes cannot be described by continuum mathematics but rely on probabilistic descriptions. Continuum mathematics can describe the average or ensemble behavior of a population of channels, provided the population is sufficiently large. Depolarization to threshold requires the movement of sufficient charge to cause the depolarization. This charge can be delivered over a short time at high current or long time at low current. This is the basic nature of the strengthduration relationship. At lower currents there is more time for redistribution of charges within the nerve fiber so that the relationship is not purely reciprocal. The strengthduration relationship is adequately described by the Weiss equation: I 5 Irh(t 1 τ SD)/t, where Irh is the rheobase and τ SD is the chronaxie. REVIEW QUESTIONS 1. If a Na-selective channel was to open on the membrane of a motor neuron at rest, which way would current flow? Would this depolarize or hyperpolarize the cell? 2. If a K-selective channel was to open on the membrane of a motor neuron at rest, which way would current flow? Would this depolarize or hyperpolarize the cell? 3. What is an action potential? Why does the membrane potential become positive during the action potential? 4. What do we mean when we say that action potentials are “all or none”? What is the absolute refractory period? Relative refractory period? 5. What is the “activation gate” of the Na1 channel? When is it open? When does it close? What is the “inactivation gate” of the Na1 channel? When is it open? When does it close? 6. What is tetrodotoxin? Why does it block action potentials? 7. What is a patch clamp? Why is it useful? What is an ensemble current? 8. How do gNa and gK change during the action potential? What causes these changes? 9. What is the total current across the membrane at the resting membrane potential? Why does a slight depolarization, below threshold, come back to the resting membrane potential? Why does a slight hyperpolarization return to the resting membrane potential? 10. Why is there an inverse relationship between current and time to reach threshold? APPENDIX 3.2. A1 THE HODGKINHUXLEY MODEL OF THE ACTION POTENTIAL ALAN HODGKIN AND ANDREW HUXLEY’S GOAL WAS TO ACCOUNT FOR THE ACTION POTENTIAL BY MOLECULAR MECHANISMS Hodgkin and Huxley’s goal was to explain the ionic fluxes and conductance changes during the action potential in terms of molecular mechanisms. After trying some different mechanisms, they concluded that not enough was known to determine a unique mechanism. Instead, they tried to develop an empirical kinetic description which would allow them to calculate electrical responses and which would correctly predict the shape of the action potential and its conduction velocity. Today this model is referred to as the HodgkinHuxley or HH model. The Action P otential THE HH MODEL DIVIDES THE TOTAL CURRENT INTO SEPARATE NA 1 , K 1 , AND LEAK CURRENTS An alternative way of expressing Eqn [3.2.A1.3] is as follows: ½3:2:A1:5 dn ðnN 2 nÞ 5 dt Tn Hodgkin and Huxley wrote the Na1 and K1 currents in terms of their maximum conductances which are multiplied by coefficients that vary continually between 0 and 1. The overall conductance, then, varies between 0 and the maximum conductance. All of the kinetic properties of the conductances are embedded in the characteristics of the coefficients. The conductances in the model vary with voltage and time but not with concentration of either Na1 or K1. where nN is the steady-state value of n at any particular voltage and Tn is a time constant. The values of nN and Tn are given by αn nN 5 αn 1 β n ½3:2:A1:6 1 Tn 5 αn 1 β n THE HH MODEL OF THE K 1 CONDUCTANCE INCORPORATES FOUR INDEPENDENT “PARTICLES” THE HH MODEL OF NA 1 CONDUCTANCES USES ACTIVATING AND INACTIVATING PARTICLES On depolarization, the increase in gK follows an S-shaped curve, whereas gK decreases exponentially upon repolarization. Hodgkin and Huxley proposed that the gating of the K1 channel could be modeled by four identical membrane “particles.” The probability that each channel is in the position to allow K1 conductance is n. All four particles must be correctly situated to allow conductance. The probability that all four are positioned for conductance is n4. The K1 current is given as Similar to the case with the K1 conductance model, HH empirically modeled the Na1 conductance with four hypothetical gating particles that make first-order transitions between conductive and nonconductive states. Because the Na1 conductance has two opposing actions, activation and inactivation, Hodgkin and Huxley used two kinds of gating particles, called m and h. Here the probability of an open configuration is m and h, respectively. The probability that all four gates are open is m3h. In this case, the Na1 current is given as ½3:2:A1:1 ½3:2:A1:7 IK 5 n4 gKmax ðEm 2 EK Þ Here the probability n depends on time and voltage. In the HH model, values of n are determined by a firstorder reaction, written as ½3:2:A1:2 αn 12n"n ½3:2:A1:3 dn 5 αn ð1 2 nÞ 2 β n n dt Hodgkin and Huxley determined empirical relationships between the rate constants, αn and β n, and the membrane potential. For the squid axon at 6 C, these were: ½3:2:A1:4 The same formalism for transitions between open and closed states of the n gates for the K1 channels also governs transitions between the open and closed states of the m and h gates of the Na1 channel: βn where αn and β n depend on voltage. The reaction in Eqn [3.2.A1.2] can be written in differential form as αn 5 0:01ð10 2 ðEm 2 Er ÞÞ eð102ðEm 2Er ÞÞ=10 2 1 β n 5 0:125eð2ðEm 2Er Þ=80Þ where the membrane potential is in millivolts and the rate constants have units of M s21. These are not the exact forms originally used by Hodgkin and Huxley, because their sign convention for membrane potential is the reverse of that used today. In addition, Hodgkin and Huxley derived their empirical equations based on voltage clamp experiments in which the degree of variation from resting membrane potential, here symbolized as Er, was clamped. Equation [3.2.A1.4] is transformed to agree with today’s conventions. INa 5 m3 hgNamax ðEm 2 ENa Þ αm ½3:2:A1:8 12m"m βm αh 12h"h βh The rate constants all depend on voltage, according to the following empirical relations: αm 5 ½3:2:A1:9 0:1ð25 2 ðEm 2 Er ÞÞ eð252ðEm 2Er Þ=10Þ 2 1 β m 5 4eð2ðEm 2Er Þ=18Þ αh 5 0:07eð2ðEm 2Er Þ=20Þ 1 β h 5 ðð302ðE 2E ÞÞ=10Þ m r 11 e Relationships among mN, hN, Tm, and Th are given in analogy to Eqn [3.2.A1.6]. CALCULATION OF GNA(T) AND GK(T) FOR A VOLTAGE CLAMP EXPERIMENT To calculate the time dependence of Na1 and K1 conductances, we need to know the set of rate constants 277 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION which describe the transitions between states of the conductances in the HH model: (αn, β n, αm, β m, αh, β h). Further, we need to know this set at the two voltages. In the voltage clamp experiment, the nerve is held at some voltage and then is rapidly switched to another voltage. As a result, the conductances go through a transition from one state to another. Because the gates have different kinetics, the overall behavior can be complex. Each of the gates relaxes between the steady-state value before the voltage jump, and the second steady-state value at the new voltage. The equation describing the time course from an initial value of n, m, and h (n0, m0, and h0) and a final value of n, m, and h (nN, mN, and hN) can be derived from integrating Eqn [3.2.A1.3]: 1.0 9 n∞ 8 0.8 7 6 n∞ 0.6 5 4 0.4 τn 3 2 0.2 1 dn dt 5 αn ð1 2 nÞ 2 β n n 0 5@ ½3:2:A1:10 10 τn 278 0 –100 1 3n α t n ln4 2n5 5 2 Tn αn 1β n ½3:2:A1:11 2 2 0 50 Em (mV) αn 2 nAðαn 1 β n Þ αn 1 β n dn 5 ðαn 1 β n Þdt ðαn =αn 1 β n Þ 2 n ðn ðt dn 5 ðαn 1 β n Þdt n0 ðαn =αn 1 β n Þ 2 n 0 Integrating, we obtain 0.0 –50 FIGURE 3.2.A1.1 The dependence of τ n and the steady-state value of n (nN) as a function of the membrane potential, Em. These values of n control the conductance through the K1 channel in the HH formalism. nN, and Tn, we can calculate the n(t). All of these values depend on the α and β for the given gates. The procedure is entirely analogous for m and h. RESULTS OF THE CALCULATIONS Figure 3.2.A1.1 shows the voltage dependence of nN and Tn. n0 3 n 2 n N 55 2 t ln4 nN 2 n0 Tn n 5 nN 2 ðnN 2 n0 Þe2ðt=Tn Þ Here we have identified Tn and nN according to Eqn [3.2.A1.6]. It is easy to justify the assignment of nN: simply let dn/dt in Eqn [3.2.A1.3] go to zero at infinite time (at which point the steady-state value of n should have been reached). From that constraint, we find nN 5 αn/(αn 1 β n). The last equation tells us that at t 5 0, n 5 n0 and at t 5 N, n 5 nN. In between, n relaxes between n0 and nN with an exponential time course. If we know n0, The voltage clamp experiments performed by Hodgkin and Huxley involved changing the membrane potential essentially instantaneously from the resting potential (265 mV) to some set potential and clamping it there. What happens is that the resting values of n, m, and h (called n0, m0, and h0) relax to their new values at the clamped voltage, which we will call nN, mN, and hN. The resulting time course of n, m, and h generates a time course for the conductances, calculated according to Eqns [3.2.A1.1] and [3.2.A1.7]. For a 188-mV clamp from 265 to 123 mV, the relevant values are given in Table 3.2.A1.1 (see Figure 3.2.A1.2). These values are substituted into equations of the form of Eqn [3.2.A1.11] to derive n(t), m(t), and h(t), from which the instantaneous conductances can be calculated TABLE 3.2.A1.1 Values of n, m, and h for Voltage Clamp from 265 to 123 mV K Channel Na Channel n0 5 0.3177 m0 5 0.0529 h0 5 0.5961 nN 5 0.9494 mN 5 0.9953 hN 5 0.0009 Tn 5 1.2028 ms Tm 5 0.1577 ms Th 5 1.0022 ms gKmax 5 36 ms cm22 gNamax 5 120 ms cm22 Derived from Figures 3.2.A1.1 and 3.2.A1.2. The Action P otential 1.0 m∞ 0.8 1.0 10 1.0 0.8 8 0.8 τm 0.4 0.4 4 0.4 0.2 0.2 2 h∞ 0.6 τm (ms) 6 0.6 0.6 m∞ τm (ms) τh 0.2 h∞ 0.0 –100 0 –100 0.0 –50 0 50 0.0 –50 Em (mV) 0 50 Em (mV) FIGURE 3.2.A1.2 The dependence of τ m and mN on the membrane potential. These values control the opening of the activation gate of the Na1 channel in the HH fomalism. 50 Time after voltage clamp from –65 mV to +23 mV 45 gK or gNa (ms cm–2) 40 35 gK 30 25 20 15 10 gNa 5 0 0 1 2 3 4 5 Time (ms) 6 7 8 FIGURE 3.2.A1.3 Calculated changes in gNa and gK during a voltage clamp from 265 mV to 123 mV using the HH formalism. according to Eqns [3.2.A1.1] and [3.2.A1.7]. The results of these calculations for the given voltage clamp are shown in Figure 3.2.A1.3. Of course, Hodgkin and Huxley had it much more difficult than this, because they had to find the original equations and parameters to fit their voltage clamp results, whereas here we are simply confirming that the equations and parameters they found do, indeed, look like their voltage clamp records. It is important to remember that the HodgkinHuxley formalism is an empirical model, designed to fit the data. Although there are deficiencies in the model, it succeeds admirably well in predicting the wave form of the action potential. It is now generally accepted that the basis of the action potential is a rapid switching on of the Na1 conductance, followed by its inactivation and more slowly turning on of the K1 conductance. 279 3.3 Propagation of the Action Potential Learning Objectives G G G G G G G G G Define propagation of the action potential Define conduction velocity Describe how conduction velocity varies with axon diameter and with myelination Using the formula for a parallel plate capacitor, explain how myelin decreases membrane capacitance Using the formula for resistances in parallel, explain how internal resistance of the axoplasm varies with axon diameter Define the space constant and time constant Describe how the space constant and time constant vary with axon diameter and myelination Qualitatively account for how membrane capacitance and axoplasmic resistance explain the dependence of conduction velocity on myelination and axon diameter Describe saltatory conduction and explain what “jumps” from node to node THE ACTION POTENTIAL MOVES ALONG THE AXON Consider Figure 3.3.1, which shows an axon of a motor neuron that has been impaled at intervals by recording electrodes. If an action potential is begun at the far left by depolarization to threshold, each succeeding recording electrode records an action potential. Note that the successive action potentials have similar waveforms but they are observed at each electrode at successively later times. The action potential is propagated along the surface of the nerve. The action potential moves over the surface of the cell, appearing some distance away after some elapsed time. THE VELOCITY OF NERVE CONDUCTION VARIES DIRECTLY WITH THE AXON DIAMETER The action potentials shown in Figure 3.3.1 do not have identical waveforms due to the stimulation artifact that dies out with distance along the axon. After this initial stimulation artifact decays away, all subsequent action potentials are essentially identical. The identical 280 waveform of the action potential as it travels over the axon is a variant of the “all-or-none” description of the action potential. As the action potential appears later at longer distances from the point of initiation, we can define a conduction velocity of action potential propagation equal to the distance between the recording electrodes divided by the delay in time between action potentials recorded at the two sites. The velocity of action potential conduction has been determined for myelinated and unmyelinated fibers of different sizes (see Table 3.3.1). Within each category of nerve fiber, myelinated or unmyelinated, the conduction velocity varies with the diameter of the nerve. For myelinated fibers, the conduction velocity varies approximately in proportion to the diameter. In unmyelinated fibers, the conduction velocity varies approximately with the square root of the diameter. THE ACTION POTENTIAL IS PROPAGATED BY CURRENT MOVING AXIALLY DOWN THE AXON How is the action potential conducted down the length of the axon? Recall that the action potential is triggered by a depolarization of the membrane to threshold. In order for an action potential to move from one place to another along the axon, the depolarization that triggers the action potential must precede it. Depolarization of the membrane proceeds electrotonically or passively. As shown in Figure 3.3.2, the local depolarization of the neuron’s axon membrane spreads out from the origin of the depolarization. THE TIME COURSE AND DISTANCE OF ELECTROTONIC SPREAD DEPEND ON THE CABLE PROPERTIES OF THE NERVE Recall from Figures 3.2.4 and 3.2.5 that a hyperpolarizing or depolarizing stimulus was not faithfully reproduced in the axon: the signal was distorted. This distortion is a consequence of the cable properties of the nerve. The cable properties of the nerve refer to the passive or electrotonic properties and not to the active properties that give rise to the action potential. Each length of axon is characterized by an Ohmic resistance to current across the membrane (Rm), a capacitance (Cm), a resistance through the external solution that bathes the membrane (Ro), and a resistance through the © 2017 Elsevier Inc. All rights reserved. DOI: https://dx.doi.org/10.1016/B978-0-12-800883-6.00025-2 Propa gat ion of t he Ac ti on Potential Voltmeter – + + – + + + + + + + + + + + – + + + + + + + + + + + – + + + + + + + + + + + + + + + + + + + + + – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – – Axoplasm Direction of action potential propagation Em +60 +60 +60 +60 +40 +40 +40 +40 +20 +20 +20 Em 0 (mV) Em (mV) –20 Em 0 0 (mV) –20 +20 0 (mV) –20 –20 –40 –40 –40 –40 –60 –60 –60 –60 –80 –80 –80 0 2 4 6 8 10 Time (ms) 0 2 4 6 8 10 Time (ms) –80 0 2 4 6 8 10 Time (ms) 0 2 4 6 8 10 Time (ms) FIGURE 3.3.1 Appearance of action potentials at later times down the axon from the point of stimulation. The output of each voltmeter is shown below it. Depolarized region TABLE 3.3.1 Velocity of Nerve Impulse Conduction as a Function of Axon Size Nerve Fiber Type Diameter Conduction (µm) Velocity (m s21) Physiological Function Aα 1222 70120 Somatic motor Aδ 15 1230 Pain, sharp C 0.51.2 0.22 Pain, ache Axoplasm axoplasm that fills the axon (Ri). A schematic diagram of this electrical model is shown in Figure 3.3.3. If we pass a constant current across the membrane between nodes A and B, for example, so that some new membrane potential E is established, we should expect that the membrane potential Ex at some point x away from the current source will depend on the distance from the current source and, because of the capacitances, it will also depend on the time since the current was turned on. The cable properties determine this dependence on position x and time t. CAPACITANCE DEPENDS ON THE AREA, THICKNESS, AND DIELECTRIC CONSTANT The membrane acts much like a parallel plate capacitor. The expression for the capacitance of a parallel plate capacitor is given as ½3:3:1 C5 κεo A δ Passive spread of current depolarizes adjacent patches of membrane Axoplasm In the next patch of membrane, passive spread continues Newly depolarized regions Axoplasm FIGURE 3.3.2 Passive spread of a depolarization to adjacent areas of membrane. Depolarization of a patch of membrane spreads to adjacent areas. If the depolarization reaches threshold in the nearby patch, an action potential will be initiated. If there is no action potential, the spread of depolarization will decay away with time and distance from the original depolarized area. where C is the capacitance (in F 5 C V21), κ is the dielectric constant characteristic of the material between the plates (a dimensionless ratio), ε0 is the electrical permittivity of the vacuum 5 8.85 3 10212 C2 J21 m21, 281 282 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION External solution A Rm Ro E C Cm Membrane B Ri D F Axoplasm FIGURE 3.3.3 Schematic diagram of the electrical model of an axon. A is the area (m2), and δ is the distance between the plates (m). Since 1 J 5 1 V C, the units of C come out in C V21. From this equation, it should be clear that the capacitance depends directly on the area of the membrane. It is common to normalize the capacitance by dividing by the area. The specific capacitance Cm 5 C/A is this normalized capacitance. and GM be the specific conductance per unit area. Similarly, let R be the total resistance, which is the inverse of the conductance, and RM be the specific resistance. Then the resistance of a patch of membrane is given as The real membrane is not a parallel plate capacitor, but a concentric, coaxial capacitor. This does not materially affect the analysis presented here, as detailed in Appendix 3.3.A1. 1 1 1 1 5 5 G Gm A Gm A CHARGE BUILDUP OR DEPLETION FROM A CAPACITOR CONSTITUTES A CAPACITATIVE CURRENT The relationship among charge, volts, and capacitance is ½3:3:2 C5 Q V where C is the capacitance, Q is the charge (in coulombs) on the capacitor, and V is the voltage difference across the capacitor. When the capacitor is charging or discharging, there is a capacitative current given as ½3:3:3 i5 dQ dV 5C dt dt The capacitive current is given this special name because current does not pass through the dielectric, but charge builds up on one side and is taken away from the other, so effectively there is a charge movement across the capacitor without a charge flow through the dielectric. G 5 Gm A ½3:3:4 R5 Here the specific resistance, Rm, is given in units of Ω cm2, and R is in Ω. These units may seem peculiar, that the resistance per unit area is given as Ω cm2, but this is a consequence of Ohm’s Law that gives the current as being inversely proportional to the resistance. THE AXOPLASMIC RESISTANCE DEPENDS ON THE DISTANCE, AREA, AND SPECIFIC RESISTANCE The resistance of an electrolyte solution such as that in the axon is typically given as its specific resistance, ρ. This is the resistance between two faces of a cube 1 cm on a side. Since resistances in series add, the resistance of a length of solution is just the length times the specific resistance. The resistance of an area of electrolyte solution is the specific resistance divided by the area, as in Eqn [3.3.4]. So the equivalent resistance of the axoplasm is given as ½3:3:5 THE TRANSMEMBRANE RESISTANCE DEPENDS ON THE AREA OF THE MEMBRANE Adding membrane area is like adding resistances in parallel—the overall resistance actually decreases. The conductances add, whereas the inverse of the resistances add. The total conductance of a patch of membrane is the conductance per unit area times the area. Let G be the total conductance of a membrane of area A, Rm A Ri 5 ρi d A where d is the distance and A is the cross-sectional area. Since Ri has the units Ω, ρi has the units Ω cm. THE EXTRACELLULAR RESISTANCE ALSO DEPENDS ON THE DISTANCE, AREA, AND SPECIFIC RESISTANCE The resistance provided by the extracellular electrolyte solution is entirely analogous to the axoplasmic Propa gat ion of t he Ac ti on Potential resistance. However, the area involved here is not precisely known and it is large. Because the area appears in the denominator of Eqn [3.3.5], typically the outside resistance is small compared to the axoplasmic resistance. We will treat Ro as being zero, so that the voltage everywhere along the axon on the outside is zero. CABLE PROPERTIES DEFINE A SPACE CONSTANT AND A TIME CONSTANT Consider part of the schematic diagram of Figure 3.3.3 shown in Figure 3.3.4. Analysis of the currents as a function of distance will allow us to characterize the axon in terms of its cable properties. These include a space constant and a time constant. One of Kirchoff’s circuit laws states that the sum of all currents out of any node must be zero. This is just another way of saying that there is a conservation of total charge. Applying this principle to node (x) in Figure 3.3.4, we have iðxÞ 5 iðx 1 dxÞ 1 im ½3:3:6 where i(x) is the current passing down the axon into the node at (x), im is the current that passes through the membrane at this node, and i(x 1 dx) is the current that passes down to the next node at (x 1 dx). The current across the membrane, im, has two parts: one part that passes through the resistance and a second part that either charges or discharges the capacitor. External solution Ro The current im can be written as ½3:3:7 im 5 where V(x) is the membrane potential at position x and Vr is the membrane potential at rest at which the net membrane current is zero. The first part of the right-hand side of this equation is just Ohm’s law for the current through the resistance across the membrane. The second part is from Eqn [3.3.3] and describes that part of the current that either charges or discharges the capacitor. Substituting Eqn [3.3.7] into Eqn [3.3.6], we have ½3:3:8 iðxÞ 5 iðx 1 dxÞ 1 VðxÞ 2 Vr dV A 1 Cm A dt Rm This equation can be rearranged, using A 5 2πa dx as the surface area of the membrane element, where a is the radius of the axon. Insertion of the area in Eqn [3.3.8] and rearranging, we obtain ½3:3:9 iðx 1 dxÞ 2 iðxÞ VðxÞ 2 Vr dV 522πa A 22πaCm A dx dt Rm In the limit as dx-0, the left-hand side of this equation is the definition of the derivative. Taking this limit, we derive di VðxÞ 2 Vr dV 522πa 22πaCm dx dt Rm ½3:3:10 There is another relationship between i and V(x) that we can use here, and that is Ohm’s law through Ri. There would be no longitudinal current unless there is a voltage gradient in x. Ohm’s law gives the longitudinal current, i, in Eqn [3.3.10] as i5 R= VðxÞ 2 Vr dV A 1 Cm A dt Rm VðxÞ 2 Vðx 1 dxÞ Ri Vðx 1 dxÞ 2 VðxÞ ðρi dx=πa2 Þ πa2 Vðx 1 dxÞ 2VðxÞ i52 limdx-0 dx ρi i52 Rm ½3:3:11 A C = CmA i52 im V(x) Ri = Substituting this result for i into Eqn [3.3.10], we derive ρd A V(x + dx) i(x + dx) i(x) πa2 dV ρi dx Axoplasm FIGURE 3.3.4 Currents at a patch of membrane area of the axon. C is the capacitance of the membrane; Cm is the specific capacitance; R is the resistance across the axon membrane; and Rm is the resistance per unit area; im is the current across the membrane; i(x) is the current down the axon at node x. Ri is the internal resistance of the axoplasm and ρi is its specific resistance. ½3:3:12 2 πa2 d2 V V 2 Vr dV 5 22πa 2 2πaCm 2 dt ρi dx Rm which can be rearranged to ½3:3:13 V 2 Vr 5 Rm πa2 d2 V Rm dV 2πaCm 2 dt 2πaρi dx2 2πa If we let V0 5 V 2 Vr, this equation has the form ½3:3:14 V 0 5 λ2 d2 V 0 dV 0 2 τ dx2 dt 283 284 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Here λ is identified as a space constant and τ is a time constant, so named because they govern the spatial and time derivatives of the voltage when a constant current is injected across the membrane. By comparison with Eqn [3.3.13], their values are given as ½3:3:15 sffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffi Rm πa2 Rm a 5 λ5 ρi 2πa ρi 2 τ 5 Rm Cm Note that at steady state, when dV/dt 5 0, the space constant defines the way in which the voltage varies with distance. This would be the case when a constant current has been passed across the membrane for a sufficiently long time to charge all the capacitors to their steady-state level. In this case, Eqn [3.3.14] becomes ½3:3:16 V 0 5 λ2 d2 V 0 dx2 The relevant solution to this differential equation is ½3:3:17 V 5 ðV0 2 Vr Þeð2x=λÞ 1 Vr This equation means that the voltage falls off exponentially from the point of current application to the nerve to its value some large distance away. Here the resting variable V0 refers to the voltage at x 5 0 and Vr is the resting voltage some large distance away. The equation looks this way because of the boundary conditions that V 5 V0 when x 5 0, and V 5 Vr when x 5 N. These boundary conditions arise because of the way that we set up the situation: the current at a specified point on the axon is constant and it is applied for sufficiently long times that the capacitors are all charged and the capacitative currents go to zero. Under these conditions, at steady state, the voltage at the point of current application will be V0 and the voltage far away (actually, infinitely far away) will be the resting membrane potential. This condition is called electrotonus. Note that the space constant described in Eqn [3.3.15] consists of two resistances: the resistance to current flow across the membrane and the resistance to current flow down the interior of the axon. If the resistance across the membrane becomes larger, as occurs in myelinated fibers, then the space constant will be larger. If the resistance of the axoplasm is smaller, as occurs when the axon is larger (and therefore has a larger cross-sectional area, which participates in the determination of Ri), then the space constant will be larger. Figure 3.3.5 shows the voltage as a function of distance from the point of application of an inward current sufficient to depolarize the cell to 240 mV without the engagement of the active behavior of the axon, say in a TTX-poisoned nerve. In myelinated fibers, the voltage decreases slowly away from the point of current injection because λ, the space constant, is large. In contrast, the steady-state voltage in unmyelinated fibers decays rapidly away from the point of injection of the current. From Eqn [3.3.17], the space constant is the distance for the voltage difference, (V 2 Vr), to fall from V0 2 Vr to within 1/e of V0 2 Vr, which is 0.367 3 [V0 2 Vr]. This can be readily seen by letting x 5 λ, at which point V 2 Vr 5 [V0 2 Vr] 3 1/e. The larger space constant in myelinated fibers means that a depolarization maintains a higher voltage at longer distances, because more of the current travel down the axon to change the membrane potential as opposed to going through the membrane. Equation [3.3.15] predicts that the space constant will vary with the square root of the axon diameter. This is the case with unmyelinated fibers. In myelinated fibers, the value of Rm increases proportionate to the radius. So the overall length constant would be proportional to the square root of the square of the radius or approximately proportionate to the radius. Substituting in for Rm and ρi from Eqns [3.3.4] and [3.3.5] into Eqn [3.3.15], and using the area terms 2πad for the surface area (relevant to Rm) or πa2 for the crosssectional area (relevant to Ri), we obtain sffiffiffiffiffiffiffiffiffiffi Rm a λ5 ρi 2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2πadR a 5 ðπa2 =dÞRi 2 ½3:3:18 sffiffiffiffiffiffiffiffiffiffiffiffiffi dR 5 ðRi =dÞ 5 sffiffiffiffiffiffiffiffiffiffi R 2 d Ri It is important to note that the units of these resistances are different because of the way in which they have been normalized. Rm is the resistance per unit area, and so its units are Ω cm2. The specific resistivity, ρi, is in units of resistance per unit area per unit length, so its units are Ω cm. Insertion of these units in Eqn [3.3.18] gives the proper units of distance in the calculation of λ. R and Ri are resistances, in Ω. Since these resistances depend on the distance down the axon, the distance used is included in the calculation of R and Ri. Sometimes it is convenient to express the resistances as resistances per unit length. These variables are commonly noted as rm and ri, with units of Ω cm and Ω cm21, respectively. This is equivalent to identifying ½3:3:19 rm 5 dR Ri ri 5 d in Eqn [3.3.18]. The utility of this nomenclature is that the space constant is then given simply as rffiffiffiffiffi rm ½3:3:20 λ5 ri The temporal properties of cables can be made readily apparent by using a space clamp. This is accomplished Propa gat ion of t he Ac ti on Potential –5 –4 Ro Ro –3 Cm Cm Ro 4– Cm Ro Cm Extracellular space –1 Ro Ro Ro 1 Cm Cm 2 Ro Cm 3 Ro Cm 4 Ro Cm 5 Ro Cm Iin Rm Rm Ri Rm Ri Rm Ri Rm Ri Rm Ri Ri Rm Ri Rm Ri Rm R3 Rm Ri Ri Axoplasm –30 V0 –40 Myelinated fiber Em (mV) –50 Unmyelinated fiber –60 –70 1/e(V0 – V) V –80 λ –90 –6 –4 –2 0 2 4 6 Displacement from X = 0 (cm) FIGURE 3.3.5 Steady-state voltage as a function of distance from the point of injection of current in myelinated and unmyelinated nerve fibers. In unmyelinated nerves, the voltage drops off quickly with distance, implying a small space constant because Rm is small and Ri is large. In myelinated fibers, the voltage drops off much less with distance, indicating a large space constant. experimentally by passing a thin wire down the axon, effectively making Ri 5 0. This removes the spatial dependence of V(x,t) caused by passing a current. Here dV/dx 5 0 and Eqn [3.3.14] becomes ½3:3:21 V0 5 2 τ dV 0 dt This is a familiar equation for exponential decay. The applicable solution is ½3:3:22 V 5 ðV0 2 Vr Þe2ðt=τÞ 1 Vr where the time constant, τ 5 RmCm does not vary very much from myelinated to unmyelinated fibers because myelination increases Rm according to the thickness of the myelin layer, but it decreases Cm inversely with the thickness of the myelin layer. The time constant describes the characteristic time for the capacitor to charge or discharge from the initial value to the final value. Here V0 2 Vr refers to the difference between the membrane potential initially and the membrane potential at steady state after injection of current. Note that, as with the space constant, the time constant is defined as the time necessary for the voltage difference to decay to 1/e 5 0.367 of the difference to its new value. Each patch of membrane is identical in this way, but the distributed nature of the capacitors and resistances makes the response of each membrane dependent on its distance from the current source. In this way, the patches of membrane nearer the current source charge faster than exponentially, and those further away charge slower than exponentially. This phenomenon is easily modeled using electronic networks. THE CABLE PROPERTIES EXPLAIN THE VELOCITY OF ACTION POTENTIAL CONDUCTION As we discussed in the section “The Action Potential Is Propagated by Current Moving Axially Down the Axon,” the action potential occurs at a specified location x because the membrane is depolarized to threshold at that location. If the depolarization occurs earlier at location x, then the action potential also occurs earlier at that location and the velocity of nerve impulse conduction is faster. Myelinated nerves have a longer space constant. Thus the depolarization occurring some distance x away 285 286 QUANTITATIVE HUMAN PHYSIOLOGY: AN INTRODUCTION Action potential occurs only at nodes of Ranvier because this is the location of the regenerative Na+ and K+ channels Jumping of the current from node to node is "saltatory conduction" Waveform propagates passively with decrement between nodes Axon Myelin sheath Node of Ranvier Passive spread of current is fast because of large Rm, small Cm, and small Ri FIGURE 3.3.6 Saltatory conduction. The inward current accompanying an action potential travels down the axon to depolarize adjacent parts of the cell. In myelinated fibers, little of the current is lost across the membrane because the transmembrane resistance is high (myelin insulates the axon) and little is lost to discharging the capacitor of the axon because the capacitance is low. Enough current remains to depolarize the axon at the next node of Ranvier where voltage-gated Na1 channels reside. In this way, myelin increases conduction velocity. In between nodes, the “action potential” does not involve active currents across the membrane: it is a passive spread of the waveform of the action potential. Its decrement shown here is exaggerated for illustrative purposes. from the point of stimulation is greater in the myelinated nerve. This is due to the fact that: G G G more of the current can travel down the axon, because its resistance is less due to its larger size; less of the current crosses the membrane, because its resistance is larger in the myelinated fibers; less current is required to depolarize the membrane, because the membrane capacitance is smaller due to the greater thickness of the myelin (see Eqn 3.3.