You are on page 1of 114

Regulation of  Tissue Oxygenation

Second Edition
ii

Colloquium
Digital Library of Life Sciences

The Colloquium Digital Library of  Life Sciences is an innovative information resource for researchers,
instructors, and students in the biomedical life science community, including clinicians. Each PDF
e-book available in the Colloquium Digital Library is an accessible overview of a fast-moving basic
science research topic, authored by a prominent expert in the field. They are intended as time-saving
pedagogical resources for scientists exploring new areas outside of their specialty. They are also
excellent tools for keeping current with advances in related fields, as well as refreshing one’s under-
standing of core topics in biomedical science.

For the full list of available titles, please visit:


colloquium.morganclaypool.com

Each book is available on our website as a PDF download. Access is free for readers at institutions
that license the Colloquium Digital Library.

Please e-mail info@morganclaypool.com for more information.


Colloquium Series on
Integrated Systems Physiology:
From Molecule to Function to Disease
Editors
D. Neil Granger, Louisiana State University Health Sciences Center
Joey P. Granger, University of Mississippi Medical Center

Physiology is a scientific discipline devoted to understanding the functions of the body. It addresses
function at multiple levels, including molecular, cellular, organ, and system. An appreciation of the
processes that occur at each level is necessary to understand function in health and the dysfunc-
tion associated with disease. Homeostasis and integration are fundamental principles of physiology
that account for the relative constancy of organ processes and bodily function even in the face of
substantial environmental changes. This constancy results from integrative, cooperative interactions
of chemical and electrical signaling processes within and between cells, organs and systems. This
eBook series on the broad field of physiology covers the major organ systems from an integra-
tive perspective that addresses the molecular and cellular processes that contribute to homeostasis.
Material on pathophysiology is also included throughout the eBooks. The state-of the-art treatises
were produced by leading experts in the field of physiology. Each eBook includes stand-alone in-
formation and is intended to be of value to students, scientists, and clinicians in the biomedical
sciences. Since physiological concepts are an ever-changing work-in-progress, each contributor will
have the opportunity to make periodic updates of the covered material.

Published titles
(for future titles please see the website, http://www.morganclaypool.com/toc/isp/1/1)
Copyright © 2016 by Morgan & Claypool Life Sciences

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in
any form or by any means—electronic, mechanical, photocopy, recording, or any other except for brief quotations in
printed reviews, without the prior permission of the publisher.

The Regulation of  Tissue Oxygenation, Second Edition


Roland N. Pittman
www.morganclaypool.com

ISBN: 9781615047208 paperback

ISBN: 9781615047215 ebook

DOI: 10.4199/C00140ED2V01Y201606ISP065

A Publication in the

Colloquium Series on Integrated Systems Physiology: From Molecule to


Function to Disease

Lecture #65

Series Editor: D. Neil Granger, LSU Health Sciences Center, and Joey P. Granger, University of Mississippi
Medical Center

Series ISSN
ISSN 2154-560X print
ISSN 2154-5626 electronic
Regulation of  Tissue Oxygenation
Second Edition

Roland N. Pittman
Virginia Commonwealth University, Richmond, Virginia

COLLOQUIUM SERIES ON Integrated Systems Physiology:


From Molecule to Function to Disease #65
vi

ABSTRACT
This presentation describes various aspects of the regulation of tissue oxygenation, including the
roles of the circulatory system, respiratory system, and blood, the carrier of oxygen within these
components of the cardiorespiratory system. The respiratory system takes oxygen from the atmo-
sphere and transports it by diffusion from the air in the alveoli to the blood flowing through the
pulmonary capillaries. The cardiovascular system then moves the oxygenated blood from the heart
to the microcirculation of the various organs by convection, where oxygen is released from hemo-
globin in the red blood cells and moves to the parenchymal cells of each tissue by diffusion. Oxygen
that has diffused into cells is then utilized in the mitochondria to produce adenosine triphosphate
(ATP), the energy currency of all cells. The mitochondria are able to produce ATP until the oxy-
gen tension or PO2 on the cell surface falls to a critical level of about 4–5 mm Hg. Thus, in order
to meet the energetic needs of cells, it is important to maintain a continuous supply of oxygen to
the mitochondria at or above the critical PO2. In order to accomplish this desired outcome, the
cardiorespiratory system, including the blood, must be capable of regulation to ensure survival of
all tissues under a wide range of circumstances. The purpose of this presentation is to provide basic
information about the operation and regulation of the cardiovascular and respiratory systems, as
well as the properties of the blood and parenchymal cells, so that a fundamental understanding of
the regulation of tissue oxygenation is achieved.

Key Words
cardiovascular system, respiratory system, blood, microcirculation, oxygen transport
vii

Preface to the Second Edition

It has been five years since the first edition of this book was published, and it was thought that
enough new work had been published in the field of oxygen transport and its regulation to justify
preparing this new edition.
Two important developments related to the topic of this book have appeared since the first
edition: a novel mechanism to delineate the role of oxygen in the local regulation of blood flow was
introduced in 2013, and a study of the oxygen dependence of cellular respiration was reported in
2012.
The local regulation of blood flow and oxygen’s role in it have a long history in physiology.
This topic has been subject to intense study for over a century; however, within the past decade
questions have arisen regarding whether current explanations need revision. The standard explana-
tion of functional hyperemia is the metabolic vasodilator hypothesis in which multiple vasodilator
molecules from the active tissue act on nearby arterioles to produce increased blood flow. In the
mid-1990s the novel idea that red blood cells (RBCs), which carry oxygen, can also perform as mo-
bile oxygen sensors was introduced: in one case it was proposed that ATP was released by the RBCs;
in the other, NO (nitric oxide) was released. More recently a very different proposal has been made
in which two small signaling radicals, NO (from the vascular endothelium) and O2- (from the active
parenchymal cells), interact in an oxygen-linked way to modulate the interstitial concentration of
NO, thereby controlling blood flow. It is worth noting that only for this most recent proposal have
real time kinetic data of the vasodilator verified predicted responses. A new figure, with accompany-
ing text, summarizes these different mechanisms.
The oxygen dependence of cellular respiration has long been thought to be an “on/off ” pro-
cess whereby mitochondrial oxygen consumption is independent of oxygen tension (PO2) down to
very low levels (» 1 mmHg). This oft-repeated in vitro finding greatly simplifies the analysis and
interpretation of studies in which tissues are subjected to low oxygen environments, mostly patho-
logic situations. Recently, the novel application of phosphorescence quenching to measure PO2 and
oxygen consumption in micro-environments has revealed that the PO2-dependence of oxygen con-
sumption in situ is quite different, so that oxygen consumption varies with PO2 continuously over
most of the physiologic range. This new and unexpected finding, corroborated by in vitro studies
viii  Regulation of  Tissue Oxygenation

from another laboratory, means that the classic concept of tissue hypoxia needs to be carefully re-
considered and possibly revised, with anticipated consequences for interpretation of physiologic and
clinical data on tissue oxygenation, both in health and disease.
In addition to these new findings included in the present edition of the book, a small num-
ber of corrections which previously escaped the author’s attention have been made right; about 20
new references have been added, and three new figures illustrating the new text have been added.
It is hoped that readers will find the new material to be useful and provocative for their work and
understanding of this key research area.
ix

Contents

1. Introduction........................................................................................................1

2. The Circulatory System and Oxygen Transport.....................................................3


2.1 Design of the Cardiovascular System................................................................... 4
2.2 Hemodynamics..................................................................................................... 5
2.2.1 Flow of Blood through Single Vessels...................................................... 5
2.3 Structure and Function of the Microcirculation................................................... 6
2.4 Transcapillary Exchange of Solutes...................................................................... 7
2.5 Regulation of Blood Flow.................................................................................... 8
2.5.1 Local Regulation of Blood Flow.............................................................. 9
2.5.2 Mechanisms of Local Regulation........................................................... 10
2.5.3 Myogenic Mechanism............................................................................ 10
2.5.4 Metabolically Linked Mechanisms of Blood Flow Regulation.............. 10
2.5.4.1 Metabolic vasodilator hypothesis......................................... 11
2.5.4.2 Erythrocyte as a mobile oxygen sensor hypothesis............... 13
2.5.4.2.1 ATP release from RBCs..................................... 13
2.5.4.2.2 NO release from RBCs...................................... 13
2.5.4.3 Nitric oxide/superoxide radical pair
interaction hypothesis.......................................................... 14
2.5.5 Other Oxygen-Linked Issues of Flow Regulation................................. 15
2.5.6 Conducted Vasomotor Responses........................................................... 16

3. The Respiratory System and Oxygen Transport................................................... 17


3.1 Physical Chemistry of Respiratory Gases........................................................... 17
3.1.1 Gas Laws................................................................................................ 17
3.1.2 Properties of Gases in Liquids: Henry’s Law......................................... 19
3.1.3 Forms in Which Gases Are Carried....................................................... 20
  Regulation of  Tissue Oxygenation

4. Oxygen Transport.............................................................................................. 23
4.1 Gas Exchange and Diffusion.............................................................................. 23
4.1.1 Overall Gas Exchange............................................................................ 23
4.1.2 Diffusion................................................................................................ 24
4.1.3 Fick’s Law of Diffusion.......................................................................... 24
4.1.4 Summary of Diffusion Properties........................................................... 25
4.1.5 Gas Exchange Limited by Diffusion and Perfusion............................... 25
4.2 Oxygen in the Blood.......................................................................................... 26
4.2.1 Blood: Plasma and Red Blood Cells....................................................... 26
4.2.2 Hemoglobin (Heme + Globin)............................................................... 26
4.2.3 Binding of Oxygen to Hemoglobin: Oxygen Saturation
(Dissociation) Curve............................................................................... 27
4.2.4 Allosteric Effectors of Oxygen Binding to Hemoglobin........................ 29
4.2.5 Overall Oxygen Transport...................................................................... 31
4.2.6 Carboxyhemoglobin............................................................................... 31
4.3 Artificial Oxygen Carriers.................................................................................. 33
4.3.1 Hemoglobin-Based Oxygen Carriers..................................................... 34
4.3.2 Perfluorocarbon Emulsions..................................................................... 35

5. Chemical Regulation of Respiration................................................................... 37


5.1 Response to Altered Oxygen.............................................................................. 37
5.2 Central and Peripheral Respiratory Chemoreceptors......................................... 38

6. Tissue Gas Transport......................................................................................... 41


6.1 Utilization of Oxygen by Tissues........................................................................ 41
6.1.1 Mitochondria.......................................................................................... 41
6.1.2 Role of Nitric Oxide............................................................................... 43
6.1.3 Role of Myoglobin in Striated Muscle................................................... 44
6.2 Oxygen Transport in the Microcirculation......................................................... 45
6.2.1 Longitudinal (Axial) Profile of Oxygen in Arterioles............................. 45
6.2.2 Longitudinal (Axial) Profile of Oxygen in a Capillary........................... 47
6.2.3 Tissue Oxygen Transport: Krogh Cylinder Model................................. 49

7. Oxygen Transport in Normal and Pathological Situations:


Defects and Compensations................................................................................ 53
7.1 Description of Oxygen Transport Using Fick’s Principle................................... 53
7.2 Stagnant Hypoxia (Hypoperfusion)................................................................... 54
The Role of Inflammation in Neonatal NEC  xi

7.3 Hypoxic Hypoxia................................................................................................ 55


7.4 Anemic Hypoxia................................................................................................. 55
7.5 Histotoxic Hypoxia............................................................................................. 56
7.6 Summary of Hypoxic Conditions and Responses............................................... 57

8. Matching Oxygen Supply to Oxygen Demand..................................................... 59


8.1 Fick’s Principle................................................................................................... 59
8.2 Convective vs. Diffusive Oxygen Transport........................................................ 60
8.3 Matching Oxygen Supply to Oxygen Demand: Role of Arterioles
and Capillaries.................................................................................................... 60
8.4 Oxygen Profile Along a Capillary: Mass Balance............................................... 61
8.5 Heterogeneity of Blood Flow and Oxygen Delivery.......................................... 64

9. Exercise and Hemorrhage................................................................................... 67


9.1 Exercise.............................................................................................................. 67
9.1.1 Fick’s Principle in Exercise..................................................................... 67
9.1.2 Temporal Phases of Exercise.................................................................. 68
9.1.3 Microvascular Approach to Oxygen Transport during
Muscle Contraction................................................................................ 70
9.1.4 Limited Oxygen Release from Red Blood Cells—Effect of
Transit Time........................................................................................... 70
9.2 Hemorrhage....................................................................................................... 71
9.2.1 Fick’s Principle in Hemorrhage.............................................................. 71
9.2.2 Compensatory Mechanisms in Hemorrhage.......................................... 72
9.2.3 Circulatory Shock and Resuscitation...................................................... 73

10. Measurement of Oxygen..................................................................................... 75


10.1 Oxygen Tension (PO2)......................................................................................... 75
10.1.1 Polarographic Electrodes........................................................................ 75
10.1.2 Phosphorescence Quenching Microscopy.............................................. 76
10.2 Hemoglobin Oxygen Saturation (SO2)................................................................ 79
10.2.1 Spectrophotometry of Hemoglobin........................................................ 79
10.2.2 Resonance Raman Spectroscopy of Hemoglobin................................... 85

11. Summary........................................................................................................... 87

References.................................................................................................................. 89

Author Biography....................................................................................................... 99
xiii

Acknowledgments

First Edition

It is a pleasure for the author to acknowledge the contributions of a number of people to the work
presented in this book. Brian Duling introduced the author to the field of microcirculation, and in
particular to the importance of oxygen transport and the role of oxygen in the regulation of blood
flow. Aleksander (Sasha) Popel, and more recently Aleksander (Alex) Golub, have been long time
colleagues, both of whom have contributed valuable ideas about oxygen transport through numerous
discussions, published articles and research collaborations. Paul Grannis, my graduate dissertation
advisor, served as an early role model for clarity of thought and organization, excellence in teaching,
and a lifelong pursuit of science. The author owes much to the excellent work of the editorial and
production staff of Morgan & Claypool Life Sciences, especially Dana Dreibelbis and Joe Cho.

Second Edition

This new edition is dedicated to the memory of Brian Duling, whose energetic life in science ended
much too soon. Sasha Popel and Alex Golub have continued to provide insight and clarity into
many of the key topics included herein regarding the nature of oxygen in biological systems. The
author is indebted to and offers grateful thanks to Kelsey Hideshima who contributed her artistic
talents to produce the new figures in this edition, offered needed feedback on large sections of the
text, and used her proofreading skills to ensure that typos were kept to a minimum. Any errors
which might have escaped detection by others are the responsibility of the author. Again, the out-
standing work of the editorial and production staff of   Morgan & Claypool Life Sciences, especially
Joe Cho and Jovan Carreon, are gratefully acknowledged.


chapter 1

Introduction

In order to carry out the variety of  activities required of cells which make up an organism, a continuous
supply of energy in the form of adenosine triphosphate (ATP) is required. Cells prefer to make ATP
by the process of oxidative phosphorylation which takes place inside the mitochondria and which has
an absolute requirement for oxygen. Thus, the regulation of tissue oxygenation is a critical feature for
survival of an organism, and various mechanisms have been put in place to ensure that all cells in the
organism are afforded a supply of oxygen which is adequate to carry out cellular activities. The regula-
tion of tissue oxygenation can be studied at several different levels of organization, ranging from the
intact organism, to the collection of organs which make up the organism, to the cells which make up
each organ, and finally to the molecules which are involved in the regulatory processes, from oxygen
itself to various transport and signaling molecules that are at the smallest scale in the regulatory path-
ways. Understanding how oxygen transport works at the molecular scale and integrating the behavior
at one level of  organization to achieve the next level ultimately lead to an overall understanding of  how
tissue oxygenation is regulated. Although much can be learned at each level of organization, we will
find that studies at the level of the microcirculation provide an interface between organ and cellular
behavior, since it is at this level that one first has integration of vascular, blood, and cellular function.
Although the title of this book is Regulation of  Tissue Oxygenation, it is not clear if “regula-
tion” is the appropriate term to use when dealing with the lowest level of organization involving cells
and signaling molecules, the microcirculation. Does the traditional approach of describing regulation
in terms of one or more feedback loops linked to specific chemical mediators apply here or is this a
procrustean approach with no clear endpoint? What variable(s) related to oxygen is/are regulated? Ad-
ditional research at this level is needed to provide a more comprehensive understanding of how tissue
oxygenation is determined and many of the key issues have been the subject of a recent review [92].
What is clear is that the maintenance of an adequate supply of oxygen requires the coordi-
nated operation of the three major systems involved in oxygen transport: cardiovascular system,
respiratory system, and blood. The supply of an adequate amount of oxygen to all cells of the body is
one of the most important functions of the cardiorespiratory system. Because complete descriptions
of the cardiovascular system, the respiratory system and blood can be found elsewhere [7, 13, 67],
this presentation will focus on the most important aspects of these three systems that pertain to the
topic of regulation of tissue oxygenation.

• • • •


chapter 2

The Circulatory System and


Oxygen Transport

The cardiovascular or circulatory system is designed to ensure the survival of all cells of the body
at every moment and it does this by maintaining the immediate chemical environment of each cell
in the body (i.e., the interstitial fluid) at a composition appropriate for that cell’s normal function.
The term “homeostasis” is used to denote the approximate constancy of the internal environment
(Claude Bernard, 1878).
First consider the simple hypothetical case of a single spherical cell suspended in a large
(>100 times the cell volume), well-stirred volume of aqueous medium in equilibrium with room air
and containing other nutrients. Oxygen availability is often a limiting factor for cell survival, and
oxygen is supplied to a cell by passive diffusion. As oxygen molecules diffuse into the cell, they are
consumed, so that there is a progressive fall in oxygen concentration from the surface of the cell to
the lowest concentration which occurs at the center of the cell. For a spherical cell with a typical dif­­­
fusion coefficient for oxygen (»10-5 cm2/s) and an oxygen consumption of resting skeletal muscle
(»10-2 ml O2 cm-3 min-1), the critical size (radius) which is just adequately supplied with oxygen
from the surrounding medium is about 1 mm. Thus, we find that diffusion puts an upper limit on
the size of cells in regard to their need for oxygen.
Although diffusion is an efficient transport process over short distances (<100 µm) as seen by
the average time required for a molecule to diffuse a distance x (t » x 2/2D), how can a much larger
multicellular organism, such as the human body containing about 100 ´ 1012 cells, be adequately
supplied with oxygen? For mammals, the bathing medium for cells is water and total body water
is about 60% of body weight. For a 70-kg person, total body water is distributed among three com­
partments with the following approximate volumes: intracellular »23 l (33% of body weight); in-
terstitial »16 l (22.5% of body weight); and circulating plasma »3 l (4.5% of body weight). Cells
are bathed in interstitial fluid (ISF), but interstitial fluid volume is only a little more than half the
intracellular fluid volume. Thus, ISF cannot be considered a large reservoir of fluid, and its composi-
tion is directly influenced by cellular metabolism.
An organism is faced with the following problem: How can the composition of  ISF be main-
tained near its desired value? The solution of this problem is to introduce a circulatory system which
  Regulation of  Tissue Oxygenation

continuously refreshes the ISF by putting it in intimate contact with “fresh, reconditioned” fluid
(i.e., arterial blood). The circulating blood must be brought close to the cells (<10 µm) since nutri-
ent and metabolic waste exchange takes place by passive diffusion, a transport mechanism which is
most efficient over short distances. Thus, the cardiovascular system uses bulk flow (convection) to
reduce the effective distance between the pumping action of the heart and the various parts of an
organism.
In order for this system to be practical and do its job efficiently, two important conditions
must be satisfied: (1) there must be adequate blood flow through the smallest blood vessels, capillar-
ies, which are in contact with the cells comprising a tissue; and (2) the chemical composition of the
incoming blood must be controlled to be that which is desired in the ISF. The design and operation
of the cardiovascular system fulfill these conditions. Two important functions of the cardiovascular
system are to move material (the carrier is blood) and to move heat (tissue metabolism generates
heat that must be brought from the body’s core to the cutaneous vascular bed at its surface, where it
is radiated away from the body).

2.1 DESIGN OF  THE CARDIOVASCULAR SYSTEM

The systemic circulation and pulmonary circulation are connected in series through the four cham-
bers of the heart, so that all the blood that is pumped from the left ventricle into the systemic organs
eventually makes its way back to the right ventricle from where it is pumped into the lungs. The
systemic organs (tissues) are connected in parallel, and the following statements are consequences
of this parallel architecture: (1) the stroke volume ejected from the left ventricle is divided among
the various organs, and a given volume of blood passes through only one organ before entering the
venous outflow of the organ; (2) the arterial blood entering each organ has the same composition;
(3) the blood pressure at the entrance to each organ is the same; and (4) the blood flow to each organ
can be controlled independently (local regulation of blood flow).
The various organs and tissues can be classified as one of two broad types: (1) blood “recon-
ditioners” and (2) “essential” tissues. The main purpose of the blood “reconditioners” is to maintain
the composition of the ISF relatively constant under all conditions. In general, flows to these tissues
exceed their metabolic needs. Examples of this type of tissue are the lung, which ensures proper ex­
change of oxygen and carbon dioxide; the kidney, which maintains electrolyte composition and fluid
balance; the gut, which oversees nutrient absorption; and the skin, which is involved in temperature
regulation. The “essential” tissues are those whose function is critical at all times. The blood flows
to these tissues typically match their metabolic needs. Examples of this type of tissue are the heart,
The Circulatory System and Oxygen Transport  

which requires a continuous supply of energy to maintain its pumping activity, and the brain, which
requires a continuous supply of nutrients and a need for the washout of metabolic products in order
to maintain consciousness and carry out its critical functions. One can also add skeletal muscle dur-
ing exercise to this list, since its energy requirements and needs for washout of metabolic products
can be substantial.

2.2 HEMODYNAMICS

A requirement for the circulatory system to carry out its function of bringing blood close to cells so
that the exchange of nutrients (e.g., oxygen) and wastes can take place by diffusion is that the blood
be able to flow through the complicated networks of blood vessels in the various organs. In order
to make a viscous fluid such as blood flow, whether through a single vessel, an organ or the entire
systemic circulation, a pressure difference must be applied between the inflow and outflow of the
network. The relationship between volumetric flow, Q, and the applied pressure difference, DP, is
the fluid mechanical equivalent of Ohm’s law for electrical circuits and is expressed as Q = DP/R,
where R is the resistance to the flow of blood. Although the myriad of series and parallel connec-
tions of blood vessels in a tissue is quite complicated, each element—a single vessel segment—is
simple to deal with.

2.2.1 Flow of Blood through Single Vessels


Poiseuille’s law for a viscous fluid quantifies the relationship among the volumetric flow of blood
through a blood vessel, modeled as a circular cylindrical tube, the geometric properties of the tube
and the flow properties of the blood. Poiseuille’s law (1846) is usually expressed as:

Q = p a 4 DP /8hl , (2.1)

where Q is volumetric flow, the factor p/8 arises from the circular cross-section, a is the radius of
tube, l is the length of tube, h is the viscosity of the blood, and DP is the pressure difference between
the ends of the tube, also called the driving pressure or perfusion pressure. It is noteworthy that the
fourth power dependence of flow on radius means that blood flow is quite sensitive to changes in
radius, which can vary in the circulatory system as vasomotor tone in vessels controlling flow (i.e.,
mainly arterioles) changes. It should also be noted that vessel length is generally constant for a given
vessel and that viscosity is a property of  blood related to the ease with which it can be made to flow.
  Regulation of  Tissue Oxygenation

From the relationship among Q, DP, and R, one finds that R depends on the geometry of the vessel
and the viscosity of the blood as

R = 8hl /p a 4. (2.2)

The average velocity of blood through a vessel can also be expressed in terms of the above
factors. Conservation of flow leads to the conclusion that volumetric flow is equal to the product of
average velocity, v, and the cross-sectional area of the vessel, pa 2:

Q = p a 2v . (2.3)

Thus, the average velocity can be expressed as

v = a 2 DP /8h l . (2.4)

For a Newtonian fluid flowing through a vessel or tube of circular cross-section, the radial
dependence of velocity is described as “parabolic” due to the quadratic dependence of velocity on
radial position, r:

v(r ) = v0 [1 - (r /a )2 ], (2.5)

where a is the internal radius of the tube, and v0 is the maximum velocity that occurs on the axis
(r = 0); the minimum velocity is zero at the wall (r = a; called the “no slip” condition).

2.3 STRUCTURE AND FUNCTION OF THE


MICROCIRCULATION

The microcirculation deserves special attention since it is across the walls of these vessels that the
exchange of oxygen, among other substances, takes place [101]. Furthermore, the arterioles, also
known as the “resistance” vessels, are the primary site for control of blood flow. Thus, the blood
vessels of the microcirculation play important roles in both the convective (arterioles) and diffu-
sive (capillaries) transport of oxygen. These blood vessels are classified as arterioles, capillaries and
venules and vary in diameter from about 100–200 µm for the largest arterioles and venules down
to about 5 µm for capillaries. In terms of their structure, all these vessels possess an inner layer of
The Circulatory System and Oxygen Transport  

endothelial cells. In addition, the arterioles have a circumferential layer of vascular smooth muscle
with which they can control blood flow and its distribution within organs. Venules typically have
thinner layers of smooth muscle.
The primary function of the circulatory system is to exchange substances between blood and
tissue, and these exchange processes take place in the microcirculation. The classes of vessels play-
ing a role there are the arterioles (resistance vessels which regulate flow), capillaries (the primary
exchange vessels) and venules (exchange and collecting vessels). The amount of flow through the
capillaries appears to be regulated to maintain adequate tissue oxygenation. The regulation of blood
flow appears to be accomplished by the coordination of several different mechanisms which affect
the flow of blood through precapillary vessels.

2.4 TRANSCAPILLARY EXCHANGE OF SOLUTES

The transport mechanism of passive diffusion is a rapid and efficient mode of molecular exchange
over the small distances (tens of micrometers) between the blood supply (capillaries) and tissue cells.
Fick’s first law of diffusion (1855) describes the net rate of transfer of a substance from a location of
high concentration to one of lower concentration:

DN /D t = D A (D c/D x ) = P A D c, (2.6)

where DN/Dt is the amount of the substance exchanged per unit time, D is the diffusion coefficient
for the substance through the capillary wall, A is the surface area available for diffusion (propor-
tional to the number of blood-perfused capillaries), Dc is the concentration difference across the
capillary wall or Dc = c(blood) - c(ISF), Dx is the thickness of the capillary wall (~1 µm), and P is the
permeability of the capillary wall defined as D/Dx.
In regard to the permeability characteristics of the capillary wall, the wall is composed of a
single layer of endothelial cells about l µm thick. For lipid-soluble substances (e.g., oxygen), the
entire wall surface is available for diffusion. For water-soluble substances (e.g., glucose), there are
small aqueous pathways equivalent to cylindrical pores 80–90 Å in diameter through which they
may pass. Total pore area is about 1/1000 (i.e., 0.1%) of the surface area of a typical capillary. The
permeability of the wall to a particular substance depends upon the relative size of the substance and
the pore (“restricted” diffusion).
During times of increased activity in a tissue, there is a need for delivery of more nutrients to
the active tissue, as well as a need to eliminate accumulated metabolic wastes that result from the
  Regulation of  Tissue Oxygenation

increased metabolism of the tissue. The amount of a substance which is exchanged between blood
and tissue can be increased by having more of the anatomically present capillaries perfused with
blood. This increases the surface area available for exchange and reduces the distance that exchanged
molecules must diffuse, both of which increase the efficiency of diffusion. There is some controversy
regarding whether it is the number of blood-perfused capillaries that is important or, in the case
of oxygen exchange, whether it is the surface area of the capillary wall in contact with moving red
blood cells. Under resting, baseline conditions, the equivalent of only a fraction (about 1/3 to 1/2) of
the capillaries in a given tissue are being perfused at any given moment. During times of increased
demand for nutrients and especially oxygen (e.g., heart and muscle tissue during exercise), more
capillary pathways can be opened to flowing red blood cells. Whether a given capillary is open or
closed depends on the contractile state of a region of smooth muscle (probably a terminal arteriole)
located near the entrance to a capillary [65]. This view of capillary recruitment is the one proposed
by Krogh [69], although in recent years this view has been replaced by the idea that most capillaries
are perfused by red blood cells at any given time, and increased activity simply increases blood flow
in capillaries of the active tissue [97].

2.5 REGULATION OF BLOOD FLOW

Since the convective supply of oxygen depends directly on blood flow, the regulation of tissue oxy-
genation depends critically on the regulation of blood flow. The cardiovascular system controls
blood flow to individual organs (1) by maintaining the input pressure to each organ within narrow
limits by the mechanisms designed to regulate arterial pressure and (2) by allowing each organ to
adjust its vascular resistance (R) to blood flow to an appropriate value. The cardiac output (CO) is
distributed among the various organs according to their respective resistances so that flow (Q) in an
organ is given by:

Q = (TPR/R ) CO, (2.7)

where TPR is total peripheral resistance of the systemic circulation. There are three major mecha-
nisms that control the function of the cardiovascular system: local, neural and humoral. They can
work independently of each other, but there are also interactions among them. The local mechanisms
are intrinsic to a tissue and will be described in more detail below. The neural mechanisms involve
the central nervous system and rely primarily on the release of norepinephrine from the sympathetic
nerve endings of the autonomic nervous system. Finally, the humoral mechanisms rely on circulating
The Circulatory System and Oxygen Transport  

vasoactive hormones, such as angiotensin II and epinephrine. It is important to recognize that the
vasoregulation occurs in the resistance vessels. In the context of the regulation of tissue oxygenation, it
is most appropriate to focus on the mechanisms that control blood flow at a local level.

2.5.1 Local Regulation of  Blood Flow


The local mechanisms for regulating blood flow are intrinsic to the various tissues and can operate
independently of neurohumoral influences [17, 110]. Local regulatory processes allow each tissue in
the body some measure of autonomy to satisfy its current and particular requirements in regard to
blood flow. Because the various organs and tissues of the body are connected in parallel, the cardiac
output can be redistributed among the tissues should their relative need change by altering the re-
sistance (R) to blood flow in the affected tissues.
The site of  local regulation of  blood flow is the microcirculation, which is composed of a network
of blood vessels—arterioles, capillaries, and venules—whose functions are regulation of tis­­sue perfusion
and exchange of substances between blood and tissue. Although the topology of vascular networks is
typically quite complex, as a first approximation, one can think of most networks as a collection of mi-
crocirculatory “units” connected in parallel, where each unit is composed of a feeding arteriole, several
capillaries arising from the arteriole and a venule which collects the blood after molecular exchange
has taken place between it and the interstitial fluid. Because of the parallel structure of the network,
which is a collection of these microcirculatory units, it is possible to redistribute blood flow from one
region to another within a tissue to accommodate any alterations in local metabolic needs.
Examples of local blood flow control processes are autoregulation, reactive hyperemia and
active (or functional) hyperemia. The term “autoregulation” in this context refers to the tendency
for organ blood flow to remain constant in the face of local changes in arterial or perfusion pressure.
Autoregulation is observed in virtually every vascular bed. It is most pronounced in the brain and
kidney and is prominent in the heart, skeletal muscle, intestine, and liver. Recall that flow (Q) equals
perfusion pressure (DP = difference between inflow arterial pressure and outflow venous pressure,
Pa - Pv) divided by vascular resistance (R) so that, as DP rises through the autoregulatory range
(Pa » 80–160 mm Hg in brain and kidney), R must increase to maintain constant flow. Reactive
hyperemia refers to the elevated blood flow observed in an organ when flow is restored following a
period of circulatory arrest (i.e., occlusion of the blood supply). Hyperemia is literally an excess of
blood in a region. The magnitude of the hyperemia is related both to the duration of the occlusion
period and to the pre-occlusion blood flow. Active (or functional) hyperemia refers to the increase
in blood flow which accompanies an increase in the metabolic activity of an organ or tissue. It has
been described in skeletal and cardiac muscle, brain, intestine, stomach, salivary glands, kidney, and
10  Regulation of  Tissue Oxygenation

adipose tissue. The name of the hyperemia depends upon the specific function of the tissue (e.g.,
contraction hyperemia for muscle or secretory hyperemia for various glands). Each one of these
examples of local regulatory processes can be linked to the regulation of tissue oxygenation.

2.5.2 Mechanisms of  Local Regulation


Two major mechanisms have been proposed to account for the local regulatory phenomena de-
scribed above: the myogenic mechanism and the metabolic mechanism. Although these mecha-
nisms appear to act independently, the expression of each mechanism varies among tissues and some
combination of each one is probably operative, depending on the particular intervention, i.e., altered
perfusion pressure, flow, or tissue activity.

2.5.3 Myogenic Mechanism


The myogenic mechanism, in essence, states that vascular smooth muscle actively contracts in re-
sponse to stretch, in an attempt to maintain circumferential wall tension, T, relatively constant in
the resistance vessels. The relationship among wall tension (T ), intravascular pressure (P), internal
radius (a), and vessel wall thickness (w) is given by the law of Laplace (1805) for a cylindrical elas-
tic tube: T = Pa/w. Thus, elastic blood vessels exposed to an increased intravascular pressure will
become passively distended. The smooth muscle in the vessel wall responds by active contraction
(leading to vasoconstriction) which tends to return wall tension near its baseline value and vascular
caliber below its original value. The myogenic mechanism is sometimes referred to as pressure-
related control of blood flow [16].

2.5.4 Metabolically Linked Mechanisms of  Blood Flow Regulation


Tissue cells continuously utilize ATP as an energy source to maintain cellular function. The two most
common ways in which ATP can be produced are by oxidative phosphorylation and glycolysis. Be-
cause oxidative phosphorylation is the primary pathway for most cells to generate ATP, cells have a
continuous need for oxygen. In the presence of an adequate supply of oxygen (normoxia), the adeno­
sine diphosphate (ADP) produced from the hydrolysis of ATP is rephosphorylated as part of the pro-
cess of oxidative phosphorylation, and the contribution of glycolysis to ATP production is negligible.
When the activity of a tissue increases, blood flow through that tissue is observed to increase
to a degree that is proportional to the increase in activity. Blood flow increases because the vascular
smooth muscle of the arterioles relaxes in response to an increase in the concentration of one or
The Circulatory System and Oxygen Transport  11

more vasodilator molecules. Several different hypotheses have been put forward to account for the
observed vasodilation, but at the present time none of these possibilities has been able to account
quantitatively for the increased perfusion, i.e., active or functional hyperemia. Figure 1 illustrates
four of the most prominent ideas to explain the activity-induced increase in blood flow. Each of
these hypotheses focuses on a different mechanism involving one or more different vasodilators.

2.5.4.1 Metabolic vasodilator hypothesis

The oldest of the hypotheses dates back to 1876 when it was proposed that increasing the activity
of a tissue (e.g., skeletal muscle contraction) led to the elaboration of a vasodilator molecule, sub-
stance X, from the active parenchymal cells (see Fig. 1A). Over the years many attempts were made
to identify “substance X” to no avail. By chemical analysis of arterial and venous blood associated
with the active tissue, substances whose concentration was higher in venous blood were viewed as
candidates for the vasodilator produced by the active parenchyma. In fact, a number of molecules
were identified in this way, many of which had vasodilator properties. Examples are adenosine and
lactic acid, associated with the development of hypoxia in the active cells. Other substances, such
as potassium ion, carbon dioxide, and osmolarity, are associated with the increase in activity, but not
directly with cellular hypoxia. Other potential chemical mediators of the vasodilation have also been
proposed, but so far no one substance or even combination of substances has been able to account
for the observed vasodilation. Because multiple potential chemical mediators of the vasodilation
have been proposed, the concept of  “redundancy” has been introduced [62, 73, 83] whereby some
unknown combination of vasodilators interact in a way to provide the needed amount of blood flow
and consequent oxygen supply to the active parenchymal cells. Although “redundancy” exists in the
sense that there are undoubtedly multiple vascular smooth muscle and parenchymal cells involved
in the functional hyperemic response, for there to be simultaneously present multiple chemical
mediators of vasodilation, each with its distinct mechanism of action and dose-response relation-
ship, the integration of these disparate inputs in just the correct way to match oxygen supply to
oxygen demand would appear to be unreasonably complicated for such an important function. The
increased tissue activity leads to an increase in ATP utilization and the resultant increase in oxida-
tive and glycolytic metabolism lead to reduced PO2 and increased production of lactic acid. If the
supposed cellular hypoxia is the cause of an increased release of adenosine (through ATP degrada-
tion) and lactic acid, then restoration of blood flow (i.e., the functional hyperemia) should return
the active cells to their baseline state of being adequately oxygenated, thereby removing the hypoxic
stimulus that caused the production and subsequent elaboration of the vasodilators in the first place.
In order for this hypothesis to be valid, all the parenchymal cells should constantly be on the brink
12  Regulation of  Tissue Oxygenation

Figure 1: Schematic diagram of metabolically linked mechanisms for blood flow regulation in re-
sponse to an increase in tissue activity. A common feature for all four mechanisms is that the increased
activity causes more oxygen to be consumed, lowering intracellular and interstitial PO2. This decreased
PO2 is then related ultimately to an increased concentration of the particular vasodilator involved in the
vasomotor response, followed by an increase in blood flow. A. Metabolic mechanism; B. NO/O2- radical
pair mechanism; C. ATP release from RBCs mechanism; D. NO release from RBCs mechanism.
The Circulatory System and Oxygen Transport  13

of hypoxia, a state which does not appear to be borne out by experimental observation. In addition
to the logical inconsistency regarding removal of the hypoxic stimulus, the fact that the metabolic
vasodilator hypothesis cannot be falsified, as an unobserved vasodilator can always be proposed to
explain any discrepancy in response, this hypothesis does not stand on a firm footing.

2.5.4.2 Erythrocyte as a mobile oxygen sensor hypothesis

In the mid-1990s the intriguing idea was proposed that the carrier of oxygen, the erythrocyte or
red blood cell (RBC), was involved in the vasodilatory response to increased tissue activity. Two
different hypotheses utilizing the release from the RBC of either ATP [28] or nitric oxide [117]
were published at about the same time. A common feature of both proposals is that an increase in
parenchymal cell activity leads to a reduction in PO2 within the active cells, as well as in the adja-
cent interstitial fluid, and this reduction in PO2 is propagated by diffusion, ultimately to the RBCs
flowing through nearby capillaries which provide oxygen to the parenchymal cells. As oxygen is
then released from the hemoglobin molecules inside the RBCs, a conformational change takes place
whereby the hemoglobin goes from the oxygenated R (relaxed) state to the deoxygenated T (taut)
state. From this point the two hypotheses differ in regard to the identity of the vasodilator and its
mode of release. These ideas are illustrated in Fig. 1, panels C and D, respectively.

2.5.4.2.1 ATP release from RBCs  The RBCs require a supply of ATP to run the Na-K
ATPase (Na-K pump) which maintains the ionic gradients across the RBC membrane. Because
RBCs do not contain mitochondria, they produce ATP by glycolysis and have an ATP concentra-
tion in the millimolar range. As the hemoglobin molecule transitions from the R to the T state,
this conformational change is transmitted to the RBC membrane and leads to the release of ATP
through a band 3 channel [60]. The ATP then moves through the plasma to the vascular endo-
thelium, where it binds to P2Y purinergic receptors, causing the production and release of NO,
prostacyclin (PGI2), and endothelium-derived hyperpolarizing factor (EDHF). These substances
then produce a conducted electrical response which is conducted upstream to the arterioles, whose
vasodilation is the cause of the observed functional hyperemic response (see Fig. 1C).

2.5.4.2.2 NO release from RBCs  According to this version of the “RBC as a mobile O2
sensor” hypothesis, as the RBCs pass through the lungs, NO produced by nitric oxide synthase in
the endothelial cells (eNOS) is taken up into the RBC cytoplasm and is carried by the hemoglobin
14  Regulation of  Tissue Oxygenation

as S-nitroso hemoglobin (SNOHb) at the cysteine-93 locus on the beta chains of hemoglobin.
As in the ATP release hypothesis, when the hemoglobin becomes deoxygenated, there is a con-
formational change from the R to the T state and this causes NO to be released from the hemo-
globin. NO then diffuses out of the RBC and to the arteriolar smooth muscle, where it produces
vasodilation. The most serious issue with this hypothesis is that it is difficult to understand how the
NO can escape from the RBC, given the very high affinity of the unliganded heme group for NO
(see Fig. 1D). An interesting alternative to this hypothesis is that the source of the NO is NO2- ion
which enters the RBC from the plasma and is acted on by deoxygenated hemoglobin which takes
on the role of a nitrite reductase, enzymatically converting NO2- to NO [35]. How the free NO is
able to exit the RBC cytoplasm without being scavenged by the hemoglobin remains an obstacle to
understanding how this mechanism works in vivo.

2.5.4.3 Nitric oxide/superoxide radical pair interaction hypothesis

The endothelial cells continuously produce nitric oxide (NO) through the action of the constitutive
enzyme eNOS on the substrates l-arginine and molecular oxygen. NO is a vasodilator and, for it to
be “the” chemical mediator of functional hyperemia, a complementary molecule should be released
from the parenchymal cells to neutralize some of the NO produced, thereby modulating the con-
centration of NO in the interstitium and, hence the degree of vasodilation. An obvious candidate
for this complementary signaling molecule is superoxide (O2-), produced into the extracellular space
by NAD(P)H oxidase in the plasma membrane of the parenchymal cells (e.g., sarcolemma of skel-
etal muscle cells) from its substrates oxygen and NADPH. Chemically, the radical pair NO and O2-
react with each other rapidly and selectively (spin-prohibited interaction with other non-radicals),
so that the NO concentration can be altered quickly by slight changes in the production of O2-.
Under resting or inactive conditions, the parenchymal cell would be expected to be well oxygen-
ated, so that the production of O2- would be high and thus neutralize much of the NO produced
by the endothelium, resulting in a state of partial constriction and relatively low blood flow. When
the activity of the parenchymal cells increase (e.g., increased skeletal or cardiac muscle contraction,
increased neural activity in the brain, and digestion of a meal in the gastro-intestinal system) their
oxygen consumption would increase, thereby lowering the local PO2 and reducing O2- production;
under these conditions NADH also moves from the cytosol to the mitochondria, reducing the
other substrate for NAD(P)H oxidase. NO concentration in the surrounding interstitium would
then increase, resulting in vasodilation and increased perfusion. When the parenchymal cell activity
wanes and oxygen consumption decreases, PO2 will increase and NADH will be shuttled from the
The Circulatory System and Oxygen Transport  15

mitochondria to the cytosol, restoring the O2- production, [NO]ISF and blood flow to the “resting”
level. A detailed description of this mechanism is presented by Golub and Pittman [39–41] and
Fig. 1B provides a pictorial view of the processes involved.

2.5.5 Other Oxygen-Linked Issues of  Flow Regulation


Several other observations related to the regulation of blood flow, and hence convective oxygen de-
livery, will be considered here as they have a direct impact on the regulation of tissue oxygenation.
The question arises as to whether tissue PO2 is closely regulated and whether oxygen plays a direct
role in oxygen-linked flow regulation by acting directly on the vascular smooth muscle of resistance
vessels. Duling and Berne [22] found that tissue PO2 was regulated within a narrow range, even
when the PO2 of the superfusion solution flowing over the tissue under observation (e.g., hamster
cheek pouch or cremaster muscle) was varied over a relatively wide range of tens of mm Hg. Thus,
raising the PO2 of the superfusion solution led to arteriolar constriction, but a relatively constant
tissue PO2, suggesting that some components of the tissue and/or the arteriolar wall were sensitive
to oxygen and communicated with the arterioles to limit blood flow and oxygen delivery to a de-
sired level. Furthermore, subsequent experiments using a bicarbonate-buffered superfusion solution
showed that in the presence of carbon dioxide, tissue PO2 was still regulated but at a higher PO2 than
in the absence of carbon dioxide [19]. Duling [19] suggested that there were two possibilities that
might explain this finding: (1) the oxygen supply might be regulated by a direct effect on vascular
smooth muscle or (2) oxygen acts only through cellular metabolism and that the rate of oxygen de­­liv­
ery to cells, rather than the absolute PO2, would be regulated. Further in vitro [93] and in vivo [20]
experiments raised serious doubts that vascular smooth muscle had the requisite sensitivity to oxygen,
so that the blood and nearby tissue would need to become severely hypoxic (PO2 of only a few mm Hg)
before arteriolar smooth muscle would respond with relaxation. Cytochrome c oxidase, which has a
high affinity for oxygen (low KM or P50 of ~1 mm Hg) and is thus responsive to oxygen levels over a
narrow range, was considered to be the oxygen “sensor” in the above studies. Duling [21] later con-
sidered that other oxidases and oxygenases, with higher KM’s or P50’s and which serve other oxygen-
linked processes, might take on the role of oxygen sensor.
Is tissue PO2 really regulated in the sense of being a controlled variable in a feedback loop
which involves a chemical mediator? There are several oxygen-linked systems which have been
proposed for modulating blood flow [17]. Because the maintenance of tissue oxygenation is such
an important feature for survival of the organism, it seems necessary that some mechanism must
exist to ensure an adequate oxygen supply to all cells of the organism. Perhaps there is some overall
16  Regulation of  Tissue Oxygenation

coordination of events which regulates blood flow to ensure delivery of oxygen and nutrients and
removal of metabolic wastes, and leads to an oxygen level consistent with maintaining a balance
between energy demand and production.

2.5.6 Conducted Vasomotor Responses


It has been found that the local vasomotor responses can spread from their point of origin to up-
stream and downstream sites by electrical conduction through gap junctions between endothelial
and vascular smooth muscle cells [2, 110]. Since the metabolic responses most closely associated
with the regulation of tissue oxygenation will be expressed and sensed initially in the terminal
branches of the microvascular network (i.e., capillaries and terminal arterioles), their spread to up-
stream sites will typically lead to increased blood flow, and hence oxygen supply, through increased
vasodilation of arterioles. Thus, the local signals confined to perhaps tens of micrometers can exert
their influence over a much wider spatial domain of hundreds to thousands of micrometers, thereby
recruiting many vessels in the network to participate in the hyperemia. The sensitivity of the vascu-
lar wall to various locally produced vasoactive substances (classic metabolic mechanism), shear stress
(flow-induced release of NO from the endothelium), and stretch (myogenic mechanism) appears to
vary along the vascular network and from organ to organ. The conducted vasomotor responses thus
act to coordinate and integrate the regulation of tissue oxygenation.

• • • •
17

chapter 3

The Respiratory System and


Oxygen Transport

Before describing the regulation of tissue oxygenation, it is instructive to consider the respiratory
system and how blood flowing through the pulmonary capillaries is oxygenated. Prior to that dis-
cussion, it is necessary to review the physical chemical basis by which oxygen is carried in the blood
and how it moves between the air spaces and the blood.

3.1 PHYSICAL CHEMISTRY OF RESPIRATORY GASES

3.1.1 Gas Laws


The quantitative relationship among the pertinent variables in a gas confined to a volume, V, is
given by the Ideal Gas Law:

PV = nRT , (3.1)

where P is the pressure due to gas molecules in the volume, V; n is the number of moles of gas in the
volume; R is the molar gas constant (= 0.082 atm l K-1 mol-1); and T  is the absolute temperature
(K). For the respiratory gases at normal physiological pressures and temperatures, the Ideal Gas
Law adequately describes the relationship among P, V, n, and T. The Ideal Gas Law is a combina­
tion of Boyle’s law (1622, PV = constant) and Charles’ (1678) and Gay-Lussac’s laws (1809) to
represent the relation among the pressure, volume and temperature of a given mass of ideal gas. A
related and useful relationship is Avogadro’s law (1811) which states that equal volumes of different
gases at the same temperature and pressure contain the same number of molecules. At one atmo-
sphere pressure (760 mm Hg at sea level) and a temperature of 0°C (273 K), the volume of 1 mole
of an ideal gas is 22.4 l.
Dalton’s law of partial pressures defines the partial pressure of a gas in a gas mixture as the
pressure that the gas would exert if it occupied the total volume of the mixture in the absence of the
other components (i.e., no interactions between gas molecules). Dalton’s law follows directly from
18  Regulation of  Tissue Oxygenation

the Ideal Gas Law since it states that the pressure exerted by a gas is proportional to the number of
moles of that gas. Thus,

Pi = Fi P B, (3.2)

where Pi is the partial pressure of gas component i, Fi is the mole fraction of component i (or
ni/ntotal, where ni is the number of moles of component i among the total number of moles of all
gases in the mixture, ntotal) and PB is the ambient barometric pressure. The total pressure of a gas
mixture is thus the sum of the partial pressures of all the components. For the four gases normally
found in the alveoli, we have

PO2 + P  CO2 + P  N2 + P  H2O = P B (3.3)

because, by definition, PB is the total pressure of all gases in a mixture in contact with the atmo-
spheric pressure. For instance, at sea level, barometric pressure is 760 mm Hg, and oxygen makes
up 21% of dry air. Thus, the partial pressure of oxygen in dry air is PO2 = 0.21 ´ 760 mm Hg =
160 mm Hg.
It is useful to consider a few additional definitions and conventions related to respiratory
gases. Pressure (P ) is generally expressed in atmospheres (atm) or mm Hg. At sea level, 1 atm is
equivalent to 760 mm Hg. The unit “mm Hg” is sometimes referred to as “torr.” Barometric pres-
sure decreases at higher altitudes (e.g., at 5,000 ft, PB is 632 mm Hg). The absolute (Kelvin) scale
is generally assumed in all equations involving temperature (T ). To convert °C to K, add 273 to the
Celsius reading.
Since respiratory gases are typically in a humidified environment, the vapor pressure of wa-
ter, PH2O, deserves special consideration. For a volume of water at temperature, T, there will be
some water molecules in the gas phase above those in the liquid phase due to evaporation. When
equilibrium is reached between the gas and liquid phases (i.e., rate of evaporation = rate of con-
densation or return of gaseous water molecules to the liquid phase), the partial pressure due to
the water molecules in the gas phase is defined to be the vapor pressure of water and its value de-
pends on the temperature; values for different temperatures have been tabulated. For instance, PH2O
(20°C) equals 17.5 mm Hg and PH2O (37°C) equals 47.0 mm Hg; PH2O increases with increasing
temperature. Tracheal air is 100% humidified so that PH2O = 47 mm Hg in the trachea and beyond
(i.e., alveolar gas, blood) at normal body temperature.
The two symbols BTPS and STPD represent two conventional conditions used when dis-
cussing respiratory gases. BTPS stands for body temperature (37°C), ambient pressure, and gas
The Respiratory System and Oxygen Transport  19

saturated with water vapor, whereas STPD stands for standard temperature (0°C or 273 K) and
pressure (760 mm Hg) and dry (no water vapor). Pulmonary ventilation (1/min) is usually measured
at BTPS, whereas gas volumes in blood are usually expressed at STPD. To convert a gas volume at
BTPS to one at STPD, one needs to multiply the former volume by (273/310) (PB - 47) / 760.
Finally, a few words about gas mixtures are in order. Consider a dry gas mixture containing
the following fractional amounts of O2, CO2 , and N2: FO2, FCO2 , and FN2, respectively. If barom­
etric pressure is PB, then the partial pressure of oxygen is PO2 = FO2 PB. Similar expressions hold for
the other gases. If this same gas mixture is placed in a humidified volume (i.e., saturated with water
vapor), then PO2 = FO2 (PB - PH2O), where FO2 is the fraction of the gas that is oxygen in a dry
mixture. The partial pressures of all the gases in the mixture must add up to PB, but PH2O is already
set by the ambient temperature. One can think of the other gases as being diluted by the presence
of water vapor.

3.1.2 Properties of Gases in Liquids: Henry’s Law


Gases are soluble to varying degrees in liquids and, as a general rule, the solubility decreases with
increasing temperature. Henry’s law (1803) relates the amount of gas dissolved in a liquid to the
partial pressure (P) of the gas: C  = a P, where C is the concentration of the gas in solution, and a is
the Bunsen solubility coefficient, which is specific for a given gas and liquid.
When a liquid is put in contact with a gas phase of partial pressure, Pgas, gas will dissolve in
the liquid until equilibrium is reached between the two phases. The condition for equilibrium of a
gas between a gas phase and a liquid phase is that the partial pressures of the gas are equal in the
two phases, not the concentrations, as is the more familiar case for non-gaseous solutes. Thus, at
equilibrium, Pliquid = Pgas rather than Cliquid = Cgas. This is an important principle in understand-
ing gas exchange between the alveolar space (gas phase) and the pulmonary capillary blood (liquid
phase). In general, gases diffuse between sites where there is a difference in partial pressure (e.g.,
red blood cell and plasma, ISF, and cell cytoplasm), not in concentration. Thus, gas exchange takes
place between the two phases as long as there is a partial pressure difference between them. Once
equilibrium is reached, net gas exchange ceases.
Table 1 gives Henry’s law solubility coefficients for several common gases in four different
media.
Except for CO2, these gases are quite insoluble in aqueous media. The partial pressure used
above in Henry’s law is the pressure in the gas phase above a solution in equilibrium with that gas.
Dissolved gas concentrations are often expressed as “volume percent (vol% = ml gas/100 ml liq-
uid).” Thus, 10 ml of O2 dissolved in 100 ml of  water would yield a concentration of 10 vol%. Key
20  Regulation of  Tissue Oxygenation

Table 1:  Solubilities of some common gases

Water Plasma Red Cells Whole Blood

Oxygen 0.0232 0.0209 0.0260 0.0230

Carbon dioxide 0.545 0.490

Nitrogen 0.0127 0.0137

Carbon monoxide 0.0182


Solubilities at 38°C are based on Henry’s law: [X] = a X P, where [X] is the concentration of the gas in ml gas/ml solution, a X is
solubility of gas X in ml gas/(ml solution atm) and P is in atm. Values for red blood cells and whole blood do not include bound
gas. Data from Roughton [89].

assumptions for Henry’s law are that (1) the gas does not chemically react with the solvent and
(2) the gas does not bind to other molecules in the solution.

3.1.3 Forms in Which Gases Are Carried


All gases are carried in solutions in the dissolved form, and Henry’s law describes the relationship
between gas content or concentration, [X ], and the partial pressure of the gas, P X:

[X ] = α X P X, (3.4)

where a X is the solubility coefficient of gas X. It is important to remember that only free gas mol-
ecules (those that are physically dissolved) contribute to the partial pressure of a gas in solution.
Gases bound to proteins (e.g., O2, CO2, CO) or present in a chemically modified form (e.g., CO2
as HCO3-) do not contribute to the partial pressure of the gas and hence are not “seen” by regulatory
mechanisms that rely upon free gas molecules.
Gases can be bound to protein molecules, such as hemoglobin or plasma proteins. Examples of
this form of gas carriage are oxygen or carbon monoxide bound to the heme groups of the hemoglobin
molecule or carbon dioxide bound to the terminal amino groups of hemoglobin or plasma proteins.
Some gases can be carried in a chemically modified form. For instance, most (~90%) of the
CO2 in blood is carried as HCO3-. Carbon dioxide taken up by the blood is converted to bicarbon-
ate in the red blood cells (with the help of the enzyme carbonic anhydrase) and plasma as blood
The Respiratory System and Oxygen Transport  21

passes through the tissues, and the process is reversed when the blood passes through the pulmonary
circulation.
The total concentration of a gas in a liquid is the sum of the concentrations of the various
forms in which it is carried and is given by the expression:

[gas]total = [gas]dissolved + [gas]bound + [gas]chemically modified. (3.5)

• • • •
23

chapter 4

Oxygen Transport

Some general comments about gas exchange and diffusion will be made, followed by a description
of how oxygen is carried in the blood. The binding of oxygen to hemoglobin will be discussed, in-
cluding the oxygen saturation (or dissociation) curve and factors (allosteric effectors) which cause it
to shift. Next, a discussion of the effects of carbon monoxide on oxygen binding will be presented.
Finally, a description of artificial oxygen carriers will be presented. Most of these topics are covered
in standard textbooks [7, 13, 67, 133] and monographs on oxygen transport [132].

4.1 GAS EXCHANGE AND DIFFUSION

4.1.1 Overall Gas Exchange


Table 2 gives the partial pressures of the four respiratory gases in dry air, moist tracheal air, alveoli,
and arterial and venous blood.

Table 2:  Partial pressures of gases in gas and blood phases

TOTAL AND PARTIAL PRESSURES OF GASES (mm Hg)

Moist Mixed
Alveolar Arterial
Dry Air Tracheal Venous
Gas Blood
Air Blood

PO2 159.1 149.2 104 100 40

PCO2 0.3 0.3 40 40 46

PH2O 0.0 47.0 47 47 47

P N2 600.6 563.5 569 573 573

P TOTAL 760.0 760.0 760 760 706


24  Regulation of  Tissue Oxygenation

The composition of alveolar gas depends upon the composition of inspired gas, composition
of gas in the functional residual capacity (FRC), minus the O2 taken up by the blood plus the CO2
added from blood. Details of how the listed composition arises are discussed in standard mono-
graphs of respiratory physiology on the topic of ventilation/perfusion defects [7, 13, 67, 133].

4.1.2 Diffusion
Diffusion takes place in the gas phase by the random motion of gas molecules.
Graham’s law of diffusion (1833) states that the rate of diffusion of a gas is inversely propor-
tional to the square root of its molecular weight (D~MW -½). Thus, the relative rates of diffusion
of CO2 and O2 are equal to Ö(32/44) or 0.85. Diffusion coefficients in the gas phase are Dgas »
10-1 cm2/sec. In the liquid phase, diffusion rates of gases are generally 10,000 times smaller than
those in gaseous environments due to the much shorter mean free path between collisions with
other molecules (e.g., the solvent); thus, Dliquid » 10-5 cm2/s [68]. This is not a severe handicap,
however, since the distances over which gas transfer must take place in the liquid phase are generally
short (about 100 times shorter than that in the gas phase).

4.1.3 Fick’s Law of Diffusion


Fick’s first law states that the amount of gas transferred per unit time (DN/Dt) across a membrane of
thickness Dx is proportional to the area (A) available for exchange and the partial pressure difference
(DP) of the gas across the membrane. The constant of proportionality (K ) is called Krogh’s diffusion
coefficient (see below) to distinguish it from D:

DN /Dt = KA D P /D x. (4.1)

For gas exchange across an alveolus in the lung, A and Dx are the same for all gases; different
transfer rates result from differences in K and DP. For the lung, Dx is about 0.5 µm—a very thin
barrier; and A is about 70 m2—a very large surface area. Krogh’s diffusion coefficient (K = aD) is
equal to the diffusion coefficient (D) times the solubility (a) of a gas in the fluid through which the
gas diffuses. For example, CO2 is 24 times more soluble than O2 in water. Thus, the rate of CO2
diffusion is 0.85 ´ 24 = 20 times as rapid as that for O2 given the same partial pressure difference.
Oxygen Transport  25

Table 3.  Solubility, diffusion coefficient, and Krogh’s diffusion coefficient


for various gases relative to values for oxygen

gas a D K = aD

O2 1 1 1

CO2 24 0.85 20

CO 0.8 1.07 0.86

N2O 16 0.85 14

N2 0.5 1.07 0.53

4.1.4 Summary of  Diffusion Properties


Table 3 summarizes the solubilities and diffusion coefficients for common respiratory gases relative
to those factors for oxygen (i.e., the values for a given gas are divided by the value of that variable for
oxygen), where a is the solubility (Henry's law coefficient; for oxygen a = 0.0296 ml O2/(ml atm)),
D is the diffusion coefficient (liquid phase; for oxygen D = 2.41 ´ 10-5 cm2/s), and K = aD is
Krogh’s diffusion coefficient (liquid phase; for oxygen K = 4.28 ´ 10-5 ml O2/(cm min atm)).
Note that all the D values are about the same as the molecular weights of these gases are
similar. Thus, differences in diffusion through the liquid phase are determined primarily by the
solubility coefficient.

4.1.5 Gas Exchange Limited by Diffusion and Perfusion


A quantitative description of the gas exchange characteristics of the lungs leads to the conclusion
that the exchange of most gases is limited by perfusion (i.e., blood flow). This is the case for oxygen
so that the blood flowing through the pulmonary capillaries comes into equilibrium with the PO2
in the alveolar gas after traversing about one-third the length of the pulmonary capillaries. One
can carry out an analysis of gas exchange by using Fick’s first law to determine the gas transport
that takes place between an alveolus and a small volume of blood as it traverses the gas exchange
region of the lung. A similar analysis is carried out for oxygen in the peripheral circulation in Chap-
ter 8. The analysis for oxygen exchange (uptake) in the lung follows the calculations represented by
26  Regulation of  Tissue Oxygenation

Eqs. (8.3) – (8.12), in which PISFO2 should be replaced by alveolar PO2 (PAO2) and PinO2 should be
replaced by venous PO2 (PvO2). For simplicity, gases that only exist in the physically dissolved form
(e.g., He, Ar, N2O) are usually considered, so that one does not have to deal with the complications
added by binding to proteins in the blood or carriage in a chemically modified form. This analysis
can be found in many monographs on respiratory physiology, as well as the more involved case of
oxygen exchange where oxygen binding to hemoglobin inside the red blood cells must be taken into
account [57, 133].

4.2 OXYGEN IN THE BLOOD

4.2.1 Blood: Plasma and Red Blood Cells


For purposes of discussing oxygen transport by the blood, we will consider blood to be composed of
two phases: plasma and red blood cells (RBCs). The fractional volume of blood occupied by RBCs
is called the hematocrit, and its value is a little less than 50% in human adults (~40% for females
and ~45% for males). Oxygen is carried in the blood in two forms: (1) dissolved in plasma and RBC
water (about 2% of the total) and (2) reversibly bound to hemoglobin (about 98% of the total).
At physiological PO2 (40 < PO2 < 100 mm Hg), only a small amount of oxygen is dissolved in
plasma since oxygen has such a low solubility. At elevated PO2 (breathing 100% oxygen or during
hyperbaric oxygenation), however, the physically dissolved form of oxygen can become significant.
Henry’s law states that the amount of oxygen dissolved in plasma is directly proportional to PO2:
[O2] = a PO2, where a = 0.003 ml O2 (100 ml plasma)-1 mm Hg-1. Thus, at a PO2 of 100 mm Hg
(typical value for arterial blood), 100 ml of plasma contains 0.3 ml O2 (or 0.3 vol%). Oxygen is car-
ried in two forms inside RBCs: it is dissolved in RBC water (about 70% of RBC volume is water)
in accordance with Henry’s law, and a much larger amount of oxygen is reversibly bound to the
hemoglobin contained within the RBCs.

4.2.2 Hemoglobin (Heme + Globin)


The protein hemoglobin is a molecule which is responsible for carrying almost all of the oxygen in
the blood. It is composed of four subunits, each with a heme group plus a globin chain. The heme
group is composed of a porphyrin ring which contains an iron (Fe) atom in its center. Normally, the
Fe is in the +2 redox state (ferrous) and can reversibly bind oxygen. There are at least six genes that
control globin synthesis in humans, resulting in the formation of six structurally different polypep-
tide chains that are designated a, b, g, d, x, and z chains. All normal and most abnormal hemoglobin
Oxygen Transport  27

molecules are tetramers consisting of two different pairs of polypeptide chains, each chain forming
a monomeric subunit.
The blood of a normal adult human contains at least six different species of hemoglobin
molecules, all of which have the same principal structure and function. Hemoglobin A (A for adult)
makes up 92% of the total hemoglobin concentration in a normal adult human. To date, approxi-
mately 200 structurally different human hemoglobin variants have been reported. These abnormal
hemoglobins (relative to hemoglobin A) often have different oxygen-binding properties.
Hemoglobin A (HbA) is composed of two a chains and two b chains, symbolically written as
a2 b2. Its molecular weight is 64,400 Da. Each a chain has 141 amino acids, and each b chain has
146. The concentration of Hb inside red blood cells is 330 g/l (= 33 g%). At a hematocrit of 45%,
this yields a blood [Hb] of 150 g/1 or 15 g%. The structure of the Hb molecule has been elucidated
by the x-ray crystallographic work of Perutz and his co-workers [18, 34]. The a and b chains are
arranged in ab pairs, and any conformational change in one polypeptide chain is transmitted to the
others in the molecule. There are two different arrangements of the subunits within the tetramer
that are much more stable than all others. One of these two quaternary conformations predominates
when the iron atoms are saturated with oxygen (oxy structure), and the other predominates when
these binding sites are vacant (deoxy structure). The deoxy structure is characterized by the presence
of inter- and intrasubunit salt bridges which give it a constrained or taut (T ) configuration. The
oxy conformation is obtained when the salt bonds are broken so as to give the tetramer a relaxed (R)
quaternary structure.
When the iron atom in the heme group becomes oxidized (loses an electron), its valence state
changes from +2 (ferrous) to +3 (ferric). The hemoglobin is then called methemoglobin (metHb)
or ferrihemoglobin (Fe+3 will not bind oxygen). Ordinarily, about 1% of the hemoglobin in a red
blood cell is in this form. The level of metHb is maintained at this low level primarily by the enzyme
NADH-methemoglobin reductase. It is important that the level of metHb be kept low since it will
not reversibly bind oxygen and thus cannot carry oxygen.

4.2.3 Binding of Oxygen to Hemoglobin: Oxygen Saturation


(Dissociation) Curve
The hemoglobin molecule has four binding sites for oxygen molecules: the iron atoms in the four
heme groups. Thus, each Hb tetramer can bind four oxygen molecules. From the molecular weight
of Hb, one can calculate that 1 g of Hb can combine with 1.39 ml of oxygen. Actually, some of the
Hb normally in red blood cells cannot bind oxygen (it is either metHb or HbCO), and the em-
pirically determined oxygen-binding capacity of hemoglobin (CHb) is 1.34 ml O2 per gram Hb. In
28  Regulation of  Tissue Oxygenation

100 ml of blood, there is about 15 g of Hb, so that 100 ml of blood has the capacity to bind 20.1 ml
of oxygen. This quantity is called the oxygen-binding capacity of  blood (CB). Note that CB is pro-
portional to the hematocrit of the blood.
If one begins with a deoxygenated sample of blood and allows it to equilibrate in steps with
gas mixtures of increasing PO2, the binding sites for oxygen will become progressively occupied
until, at a high enough PO2, all of them will contain oxygen. A curve representing the equilibrium
binding of O2 to blood is shown in Figure 2. The curve is known as the oxygen saturation curve
or the oxygen dissociation curve and expresses the relationship between PO2 and the bound oxygen
content.
P50 is defined as the PO2 at which oxygen saturation is 50%. The standard conditions under
which oxygen binding is measured are T = 37°C, pH = 7.4, and PCO2 = 40 mm Hg. The frac-
tional oxygen saturation of Hb is the amount of oxygen combined with Hb divided by the oxygen-
binding capacity of the blood (20.1 vol% at normal Hct). The bound oxygen content is proportional
to hematocrit:

[O2]bound = SO2[Hb]CHb, (4.2)

where [Hb] is blood hemoglobin concentration and is related to hematocrit (Hct) by

[Hb] = Hct[Hb]RBC, (4.3)

where [Hb]RBC is the average hemoglobin concentration of a single RBC. It is important to recog-
nize the distinction between oxygen content and oxygen saturation:

O2 content in bound form = O2 saturation × O2 -binding capacity. (4.4)

The oxygen dissociation curve is said to have a sigmoid shape, which reflects the cooperative
nature of oxygen binding to Hb. This curve is highly nonlinear in the normal physiological range of
PO2 (i.e., 40–100 mm Hg). The middle portion of the curve (20–80% saturation) is steeper than the
low PO2 and high PO2 segments. The affinity of Hb for oxygen increases steadily as oxygen satura-
tion goes from 0% to 100% for a given oxygen dissociation curve. For different oxygen dissociation
curves, the affinity of Hb for oxygen increases with decreasing P50. A simple quantitative descrip-
tion of the oxygen dissociation is expressed by Hill’s equation [3, 34]:

SO2 = (PO2 /P50)n/[1 + (PO2 /P50)n], (4.5)


Oxygen Transport  29

Figure 2: Oxygen dissociation curve-relating oxygen bound to hemoglobin (oxygen saturation, SO2)
as a function of partial pressure of oxygen (PO2). From CC Michel [80]. Used with permission of the
publisher.

where n is Hill’s coefficient and is about 2.7 for human adult hemoglobin and is related to the
degree of cooperativity of oxygen binding to hemoglobin. The oxygen-binding characteristic of
myoglobin, a related protein with one heme group that reversibly binds oxygen in striated muscle
cells, can also be described by Hill’s equation with n = 1.

4.2.4 Allosteric Effectors of Oxygen Binding to Hemoglobin


Several factors influence the binding of oxygen to hemoglobin: temperature, pH, PCO2 and 2,3 di-
phosphoglycerate (2,3 DPG). Increasing the temperature of Hb lowers its affinity for O2 and shifts
the oxygen dissociation curve to the right, as shown in Figure 3. This has physiological importance
during exercise since the temperature of muscle tissue is higher than 37°C, and oxygen can be un-
loaded from Hb more easily at the higher temperature (lowered oxygen affinity).
As seen in Figure 4, increased H+ activity (decreased pH) also lowers the affinity of Hb for
O2. This was originally noticed by Bohr (Bohr effect) and his colleagues (1904) as an effect of in-
creased PCO2, but it has been shown to be primarily an effect of pH inside the red blood cell. CO2
by itself, at constant pH, also affects the oxygen dissociation curve such that increased PCO2 shifts
the curve to the right (i.e., lowers the affinity of Hb for oxygen).
30  Regulation of  Tissue Oxygenation

Figure 3: Shifts in the oxygen dissociation curve due to changes in temperature. From CC Michel
[80]. Used with permission of the publisher.

2,3 DPG is a glycolytic intermediate produced within the RBC that affects the affinity of
Hb for oxygen. Increases in RBC [H+] cause decreases in [2, 3 DPG], and decreases in RBC [H+]
cause increases in [2, 3 DPG]. 2,3 DPG is a charged ion that cannot permeate the RBC membrane.
Increases in its concentration shift the oxygen dissociation curve to the right. 2,3 DPG is important
during respiratory compensation seen in acclimatization to altitude, whereby the hypoxic hyper-
ventilation of high altitude causes PCO2 and H+ to decrease (left shift of oxygen dissociation curve),
leading to an increase in 2,3 DPG which shifts the curve back to the right. 2,3 DPG binds to the
terminal amino groups of the beta chains and competes with CO2 for binding at those sites.
Shifts in the oxygen dissociation curve can be summarized as follows. A right shift in the
oxygen dissociation curve (� P50 or ¯ Hb-O2 affinity) can be produced by increases in any of the
following: T, PCO2, [H+] (¯ pH) or [2, 3 DPG]. A left shift in the oxygen dissociation curve (¯ P50
or � Hb-O2 affinity) can be produced by decreases in any of the following: T, PCO2, [H+] (� pH)
or [2, 3 DPG]. Ordinarily, the Bohr effect is not important except in exercise. In this situation, the
oxygen dissociation curve is shifted to the right to allow easier unloading of oxygen from Hb in the
tissues. The rightward shift in the oxygen dissociation curve is more important at lower PO2. Al-
though the rightward shift interferes with oxygen loading in the lungs, this never causes a problem
in oxygen transport.
Oxygen Transport  31

Figure 4: Shifts in the oxygen dissociation curve due to changes in pH. From CC Michel [80]. Used
with permission of the publisher.

4.2.5 Overall Oxygen Transport


Figure 5 specifies the normal partial pressures and contents (or concentrations) for oxygen. It
also describes the overall transfer of gas between the lungs and blood and between the blood and
tissue.

4.2.6 Carboxyhemoglobin
Carbon monoxide has a very high affinity for Hb (200–300 times that of oxygen in normal adults).
In situations where there is simultaneously enough oxygen and carbon monoxide to fully saturate
the hemoglobin, these two ligands compete for the same binding sites, and the relative amount of
each bound to Hb is given by Haldane’s first law:

%HbCO / %HbO2 = M PCO /PO2, (4.6)

where M lies in the range 220–270 for normal adult hemoglobin. For example, if PCO = 0.08 mm
Hg and PO2 = 80 mm Hg, about 20% of the Hb is tied up with CO (if M = 250). This represents
20% of the Hb that cannot carry oxygen (i.e., there is one part of HbCO to four parts of HbO2).
Figure 6 illustrates the effect of CO on the oxygen dissociation curve.
32  Regulation of  Tissue Oxygenation

PO2 = 160 mmHg (dry)

PIO2 = 150 mmHg (humidified)

O2 PAO2 = 100 mmHg (alveolar)

PVO2 = 40 mmHg (venous) PaO2 = 100 mmHg (arterial)


[O2]V= 15 vol% [O2]a= 20 vol%
SVO2 = 75% SaO2 = 100%

VO2

5 vol%

Figure 5: Overall oxygen transport. The partial pressure of oxygen is shown in the dry air, humidified
tracheal air (PIO2) and the alveolar compartment (PAO2). After pulmonary gas exchange takes place, the
composition of arterial blood is shown (PaO2, [O2]a, and SaO2). Following uptake of 5 vol% of oxygen
from the arterial blood by the peripheral tissues (resting conditions), the composition of venous blood is
shown (PvO2, [O2]v, and SvO2).

The curve labeled “50% Anaemia” represents a sample of  blood whose oxygen-binding capac-
ity is one-half the normal value, and none of the hemoglobin is combined with CO. Compare this
curve with the one labeled “50% HbCO”  just above it (HbCO = 50%) that also has an oxygen-binding
capacity of one-half the normal value and note that the blood with HbCO has a left-shifted oxygen
dissociation curve.
Carbon monoxide is dangerous for several reasons. When CO binds to one of the binding
sites on hemoglobin, the increased affinity of the other binding sites for oxygen leads to a left shift
of the oxygen dissociation curve and interferes with unloading of oxygen in the tissues. The pres-
ence of CO prevents loading of oxygen due to competition for the same binding sites. Carbon
monoxide binds tightly to hemoglobin (high affinity for hemoglobin), and the cumulative effect of
its binding up to the limit given above by Haldane’s first law shows that very low partial pressures of
CO (<1 mm Hg) can effectively block a large fraction of the heme-binding sites from oxygen. Thus,
the oxygen content of blood in the presence of carbon monoxide is much lower than normal. Blood
remains red because the absorption spectrum for HbCO is similar to that of HbO2, except that it is
shifted slightly to higher wavelengths compared with HbO2. Finally, there are no obvious physical
Oxygen Transport  33

Figure 6: Effect of carbon monoxide on the oxygen dissociation curve. From CC Michel [80]. Used
with permission of the publisher.

signs of carbon monoxide poisoning since carbon monoxide is colorless, odorless and tasteless; it
does not produce respiratory reflexes like coughing or sneezing; there is no increase in ventilation
(thus, PaO2 is normal); and no feeling of difficulty in breathing.

4.3 ARTIFICIAL OXYGEN CARRIERS

In cases where there is a significant (~30–40%) loss of blood volume (i.e., hemorrhage, see Chap-
ter 9), it is important to restore blood volume as soon as possible, so that the capacity of the blood
to carry oxygen to the tissues is not seriously compromised. The natural fluid for such transfusions
is whole blood since it contains all the biologically relevant components normally present in blood.
Because of concerns about the extent and safety of the blood supply, including adverse transfusion
reactions and inadvertent transmission of infectious diseases, there has been a great effort to pro-
duce artificial oxygen carriers that can act as substitutes for whole blood transfusions. Two types of
fluids have been developed for this purpose, hemoglobin-based oxygen carriers and perfluorocarbon
34  Regulation of  Tissue Oxygenation

emulsions. Their characteristics in regard to oxygen transport and regulation of tissue oxygenation
will now be presented.

4.3.1 Hemoglobin-Based Oxygen Carriers


Since plasma cannot carry much oxygen, due to its low solubility for oxygen, and hemoglobin is
the oxygen carrier within RBCs, it is natural to consider hemoglobin when formulating an arti­
ficial oxygen carrier. Several hemoglobin-based oxygen carriers (HBOCs) are in various stages of
development for the purpose of treating hemorrhagic and hypovolemic shock in trauma patients
and other circumstances where there is a compromised oxygen supply [11, 15, 137, 138]. HBOCs
are made from expired human blood or fresh bovine blood which undergoes numerous modifica-
tions to make them safe and effective oxygen carriers [15]. The RBCs are first lysed to release their
hemoglobin, and then the stroma is removed by a variety of methods, including centrifugation,
filtration and chemical extraction [63]. The stroma-free hemoglobin is then purified and undergoes
modifications to cross-link, polymerize or conjugate it to other compounds. Without these modi­
fications, the oxygen affinity of the stroma-free hemoglobin is too great to facilitate oxygen release
in the tissues due to the reduction in 2,3 DPG. When it is outside the RBC, hemoglobin rapidly
dissociates into 32 kDa ab  dimers and 16 kDa a or b monomers, both of which are rapidly filtered
in the kidney and can precipitate in the loop of Henle, resulting in severe renal toxicity [46]. For
this reason, four different types of HBOCs have been considered: cross-linked hemoglobins, cross-
linked and polymerized hemoglobins, hemoglobins conjugated to macromolecules, and encapsu-
lated hemoglobins [15].
The most notable effect following administration of HBOCs is a pressor effect, an increase
of mean arterial blood pressure (MAP) by as much as 10–35% within 15–30 minutes following ad-
ministration [46]. The pressure usually returns to baseline within 2 hours following administration
in most animal studies. The size of the HBOC appears to play a role in this pressor effect. Smaller
HBOCs, such as cross-linked tetramers, produce a larger rise in MAP than do larger polymerized
versions, possibly due to extravasation of smaller components of HBOCs from the microcirculation.
There are several theories as to why the pressor effect occurs [11, 63, 108], but the most favored pos-
sibility has something to do with the interaction of hemoglobin with nitric oxide (NO). Nitric oxide
released from endothelial cells relaxes the smooth muscle in the blood vessel walls. Normally, it is
thought that the RBC membrane prevents NO from interacting with hemoglobin. However, the
cell-free hemoglobin of HBOCs reacts freely with NO to produce HbNO from deoxyhemoglobin
and methemoglobin and nitrate from oxyhemoglobin, resulting in a reduction in NO and leading
to an unopposed vasoconstriction [63]. Another mechanism that has been proposed to account for
the arteriolar vasoconstriction and elevated MAP considers that the HBOC increases the oxygen-
Oxygen Transport  35

carrying capacity of blood and produces an oversupply of oxygen to the tissues, resulting in a com-
pensatory autoregulatory vasoconstriction [108, 137, 138]. However, a pressor effect has been ob-
served in cases where small amounts of HBOC have been infused, too small to have an effect on
oxygen supply.
In the presence of an HBOC, the total oxygen concentration of blood is given by:

[O2 ]TOTAL = [O2 ]D + [O2 ]RBC + [O2 ]HBOC, (4.7)

where [O2]D is the dissolved oxygen given by aO2 PO2, [O2]RBC is SO2RBC Hct [Hb]RBC CHb and
[O2]HBOC is SO2HBOC (1-Hct) [Hb]HBOC CHb. [Hb]RBC is the average concentration of hemoglo-
bin in an RBC, and [Hb]HBOC is the concentration of HBOC in the plasma. The assumption is
made that the oxygen-binding capacity of hemoglobin is the same for the hemoglobin in the RBC
and that making up the HBOC. This assumption can be relaxed if the oxygen-binding capaci-
ties are known to be different. Since the oxygen dissociation curves for RBC hemoglobin and the
HBOC will generally be different (i.e., different P50s and Hill coefficients, see Eq. (4.5)), one would
expect the SO2 values to be different, even assuming that equilibrium exists for PO2 between the
RBCs and the HBOC.

4.3.2 Perfluorocarbon Emulsions


Perfluorocarbon-based emulsions (PFCs) are mixtures of fluorocarbons and emulsifying agents that
differ greatly in structure and mechanism of action from HBOCs. Fluorocarbons vary in shape and
size, but share many of the same general chemical characteristics. These molecules are hydrocarbon
chains that are highly substituted with fluorine atoms. The carbon-fluorine bonds give them their
unique chemical and biological inertness. Fluorocarbon molecules used in potential artificial oxygen
carriers can be linear or cyclic, although it has been shown that linear molecules dissolve greater
amounts of oxygen [104, 108]. Fluorocarbons have a high gas-dissolving capacity and low viscosity
but are highly insoluble in aqueous solutions; they must be emulsified in order to travel through the
circulatory system. Surfactants, or surface-active agents, are used as emulsifiers to form small drop-
lets that are 0.1–0.3 µm in diameter. They are capable of dissolving large quantities of gases, such
as oxygen and carbon dioxide; however, they carry less oxygen than hemoglobin itself [84, 114].
They are capable of delivering oxygen to the tissues passively and can carry an amount of oxygen
proportional to the ambient PO2 without having to rely on the red blood cells [84].
PFCs have a half-life clearance from the body of approximately 2–4 hours and are eliminated
unmetabolized through the lungs after being taken up by the reticuloendothelial system (RES)
[108].
36  Regulation of  Tissue Oxygenation

This short half-life could potentially limit clinical uses in traumatic injury and hemorrhagic
shock. The most critical limitation of  PFCs is the linear relationship between dissolved oxygen con-
centration and PO2. Several side effects have been reported for PFCs [108, 115], and little is known
about the long-term effects of PFC retention.
Unlike hemoglobin molecules, PFCs do not bind oxygen, so that oxygen is carried by PFCs
only in the dissolved form. Thus, the concentration of oxygen carried by PFCs is given by Henry’s
law [104, 116]:

[O2]PFC = aPFCPO2, (4.8)

where aPFC is the solubility of oxygen in the PFC emulsion. The relationship between oxygen con-
centration and PO2 for hemoglobin is sigmoidal in shape, whereas it is linear for PFCs. In contrast
to the oxygen bound to hemoglobin, oxygen dissolved in PFCs is not affected by pH, 2,3 DPG or
other physicochemical factors [116].

• • • •
37

chapter 5

Chemical Regulation of   Respiration

The regulation of tissue oxygenation is based at the start from the ability of the respiratory system
to fully oxygenate the arterial blood which the heart then delivers to the peripheral tissues. The need
for different levels of respiration varies with the physiologic state of the organism (e.g., sleep, excite-
ment, exercise). The respiratory system must try to maintain constant levels of O2, CO2, and H+ in
the arterial blood which then ensures relatively constant levels of these important substances in the
interstitial fluid. For O2, one needs an adequate supply to meet cellular metabolic requirements. For
CO2 and H+, one needs to maintain the acid–base status of the body’s cells. The respiratory system
provides a rapid, but usually incomplete, compensation for acid–base disturbances through altered
PCO2. Changes in the levels of O2, CO2, and H+ in the blood cause compensatory changes in the
level of ventilation [7, 11, 55, 115].

5.1 RESPONSE  TO ALTERED OXYGEN

The response of ventilation to altered alveolar PO2 is displayed in the curves shown in Figure 7.
The following features of these curves are noteworthy. As alveolar PO2 decreases below a
threshold of about 50 mm Hg at normal PCO2, ventilation increases. At a given alveolar PO2, ven-
tilation depends on alveolar PCO2. Ventilation increases with increasing PCO2. Thus, increased CO2
potentiates the response to decreased PO2. For normal alveolar PCO2, no increase in ventilation
is observed until alveolar PO2 falls below about 50 mm Hg. Because arterial blood is still highly
saturated with oxygen (recall the oxygen dissociation curve, SO2 » 80% when PO2 is about 50 mm
Hg), there is no great need for sensitivity to PO2 above about 50 mm Hg. In normal situations, the
hypoxic stimulus is not very important. On ascent to high altitude, however, it takes on considerable
significance. For example, at an elevation of 10,000 ft, barometric pressure is about 550 mm Hg,
yielding an inspired PO2 of about 100 mm Hg and a PAO2 of about 50 mm Hg.
38  Regulation of  Tissue Oxygenation

Figure 7: Ventilatory response to changes in alveolar PO2. Hypoxic response curves. Note that when
the PCO2, is 36 mm Hg, almost no increase in ventilation occurs until the PO2 is reduced to about
50 mm Hg. (Modified from HH Loeschke and KH Gertz: Arch Ges Physiol 267:460, 1958.) From JB
West [133]. Used with permission of the publisher.

5.2 CENTRAL AND PERIPHERAL RESPIRATORY


CHEMORECEPTORS

The central chemoreceptor response to hypoxia actually depresses ventilation, presumably by de-
pressing oxidative metabolism in neural tissue. The peripheral chemoreceptors are located in the
carotid (carotid sinus) and aortic bodies (aortic arch). The carotid bodies respond to arterial hypoxia
by increasing the firing rate from the carotid sinus nerve. The carotid bodies are connected to the
respiratory centers in the brainstem, and all of the respiratory response from peripheral chemore-
ception originates in them. The carotid bodies have high blood flow and are not sensitive to CO
or anemia. The aortic bodies are connected to the cardiovascular centers in the brainstem, and they
are responsible for the cardiovascular response to respiratory-linked chemical factors in the arterial
blood. The aortic bodies have lower blood flow than do the carotid bodies, and they are sensitive
to CO and anemia. The carotid bodies are small (1–2 mg) organs with an enormous blood flow
(1–2 l per min per 100 g tissue). Their O2 consumption relative to blood flow is negligible, resulting
Chemical Regulation of Respiration  39

in a small arteriovenous O2 difference. They continuously sample arterial blood. The mechanism
of peripheral chemoreceptor oxygen sensitivity appears to involve PO2 directly rather than [O2] or
SO2. The responses by the central and peripheral respiratory chemoreceptors can be summarized
by stating that the sensed variable is arterial PO2 (PaO2); when it falls below 50 mm Hg, the cen-
tral chemoreceptors give a non-specific metabolic response that depresses ventilation, whereas the
peripheral chemoreceptors stimulate ventilation. Thus, all of the stimulatory response to hypoxia
resides in the peripheral chemoreceptors. Elevation of PaO2 above normal (~100 mm Hg) generally
has no effect on ventilation since the respiratory chemoreceptors appear to be insensitive to changes
in PaO2 above about 50–60 mm Hg.

• • • •
41

chapter 6

Tissue Gas Transport

The ultimate purpose of the cardiorespiratory system is to provide oxygen to tissue cells and to
remove the carbon dioxide produced by them. The classic view of tissue oxygenation is typically
described by one of the following three terms: normoxia—“normal” oxygen levels everywhere (i.e.,
cellular PO2 greater than a critical value—“critical” PO2—needed for generating ATP at a maximum
rate by oxidative phosphorylation in the mitochondria); hypoxia—some regions of the tissue have
less than adequate oxygen levels (i.e., cellular PO2 between zero and “critical” PO2 so that mitochon-
dria are producing ATP at a sub-maximal rate); and anoxia—absence of oxygen everywhere in the
tissue (i.e., mitochondria cease to produce ATP). We will see in the next section that this view is
not consistent with recent reports of the PO2 dependence of cellular respiration (i.e., VO2), both for
tissues in situ, as well as mitochondrial suspensions.
For a given partial pressure (difference), more CO2 than oxygen diffuses through the tissues
and across peripheral capillary walls due to CO2’s greater solubility. Because of this higher solubility,
there is generally no concern about CO2 elimination from tissues since their supply of oxygen will
be compromised earlier.

6.1 UTILIZATION OF OXYGEN BY   TISSUES

6.1.1 Mitochondria
Virtually all the oxygen consumed by tissues (~95%) is due to mitochondrial respiration. The site
within the mitochondria at which oxygen is consumed is cytochrome c oxidase, the terminal elec-
tron acceptor in the electron transport chain. Based on classic studies by Chance and co-workers
[10], the PO2 needed to maintain mitochondrial bioenergetic function (i.e., production of ATP)
is about 0.05 mmHg. This implies that mitochondrial oxidative phosphorylation is essentially an
‘on/off ’ mechanism whereby ATP production is maximal down to a vanishingly small PO2 in the
mitochondria. Other previous studies of mitochondrial and cell suspensions [135] showed that
oxygen consumption remained relatively constant down to a PO2 of about 1 mmHg. These PO2
42  Regulation of  Tissue Oxygenation

values have been used to define anoxia as the absence of oxygen (PO2 = 0 mmHg), hypoxia as the
narrow range of PO2 in cells between 0 and about 1 mmHg, and normoxia as any intracellular
PO2 > 1 mmHg. Thus, as long as the cardiovascular system could supply oxygen to cells to maintain
intracellular PO2 > 1 mmHg, they were considered to be adequately oxygenated.
Based on recent work it has become clear that the classic view expressed in the previous para-
graph should be carefully reexamined, resulting in a redefinition of the terms “hypoxia” and “nor-
moxia.” Using new methods to measure interstitial PO2 and oxygen consumption (VO2), described
in Chapter 10, the PO2 dependence of  VO2 in skeletal muscle in situ has been reported [38]. It was
found that, for a collection of muscle fibers, the empirical sigmoidal relationship between VO2 and
interstitial PO2 was well described by Hill’s equation (similar to Eq. (4.5) describing the equilibrium
oxygen binding to hemoglobin, the oxygen saturation curve):

VO2 = VO2MAX PO2m/ [P50m + PO2m], (6.1)

where VO2 is oxygen consumption at a given interstitial PO2, VO2MAX is the maximum VO2, m
is Hill’s coefficient for mitochondrial respiration (typically, 1<m<1.4) and P50 is the PO2 at half-
maximal VO2. In order to understand this relationship at a cellular level (i.e., the PO2 dependence
of mitochondrial respiration) a model was formulated which included intracellular PO2 gradients
(from PC = PO2 at the center of the cell, the lowest PO2, to PS = PO2 at the surface of the cell) and
pseudo-Michaelis-Menten kinetics for VO2:

v = VO2MAX [ p /(KM + p)], (6.2)

where v is the volume-specific VO2 at a location within the cell with PO2 = p, VO2MAX is the same
maximum volume-specific VO2 as in Eq. (6.1), and KM is the intracellular PO2 for half-maximal
VO2. When this hyperbolic dependence of cellular VO2 on PO2 is integrated over the intracellular
PO2 range from PC to PS, a sigmoidal relationship was obtained in which VO2MAX and KM were
determined. A key finding was that, for an in situ muscle at rest, KM = 10 mmHg, a value consider-
ably higher than expected from previous in vitro studies using mitochondrial and cell suspensions.
Soon after this finding was published, Wilson and co-workers [136] reported that the PO2 for half-
maximal VO2 for mitochondrial suspensions was 12 mmHg when pH was maintained at 7.4 and
changes in cytochrome c were considered.
This close correspondence between KM determined for an in situ perfused muscle and an in
vitro mitochondrial suspension, where care was taken to mimic the intracellular chemical environ-
ment, indicates that the physiologic KM is at least an order of magnitude higher than previously
Tissue Gas Transport  43

thought, an important point to consider for computational modeling of tissue oxygen transport.
These findings also suggest that the definitions for hypoxia and normoxia are in need of revision,
whereas anoxia still means the absence of oxygen, i.e., PO2 = 0. In the context of the pseudo-
Michaelis-Menten model above, the term “critical PO2” refers to the PO2 at the surface of the
cell, PS = PO2CRIT, when the PO2 at the center of the cell, PC, is zero; for resting skeletal muscle,
PO2CRIT = 4-5 mmHg. Hypoxia then refers to the situation when PS < PO2CRIT and normoxia
refers to the opposite situation when PS > PO2CRIT. These definitions are similar to previous ones
for these terms, except that now the range of PO2 over which a cell can be considered hypoxic is
expanded; that is cellular respiration should not be thought of as an “on/off ” process. This implies
that there is a wider range of PO2 at the cellular level where changes in oxygen can have an impact
on cellular function. Typical interstitial PO2 values are in the range of 40-60 mmHg and are well
above the KM and PO2CRIT for resting skeletal muscle.
The circulatory system typically supplies oxygen in the blood at a rate sufficient to maintain
tissue oxygen levels well above the critical value necessary to prevent cellular anoxia. In the context
of “regulation of tissue oxygenation,” it appears necessary only that the PO2 in the vicinity of the mi-
tochondria is greater than a low critical value; there is no need for the oxygen level to be maintained
at some specific baseline PO2. Thus, there does not need to be a single value of “normal” tissue PO2,
but rather tissue PO2 can take on any value over the physiological range from near arterial down to
a few mm Hg above the critical PO2 for normal cellular respiration, so long as the tissue receives an
adequate oxygen supply to maintain normal function.

6.1.2 Role of  Nitric Oxide


The influence of nitric oxide (NO) deserves mention due to its inhibitory action on cytochrome c
oxidase [8, 12, 55]. Thus, NO can be thought of as a “brake” that determines, in part, the maximum
oxygen consumption realized by the mitochondria under a given condition. When NO levels at the
mitochondria are high, oxygen consumption will be low; when NO levels are low, oxygen consump-
tion will be high. Nitric oxide can be produced enzymatically through the action of nitric oxide
synthase (NOS), of which there are three isoforms: n(euronal)NOS or NOS1; i(nducible)NOS or
NOS2; and e(ndothelial)NOS or NOS3. Nitric oxide can also be made in the red blood cell through
the conversion of nitrite (NO2-) to NO by the nitrite reductase activity of hemoglobin. Because
the NO released from the endothelial cells relaxes the smooth muscle of the vascular wall through
activation of soluble guanylate cyclase, NO exerts a dual action on oxygen transport: (1) it can influ-
ence the oxygen supply through its vasodilator action (see description in Chapter 2 about how NO
might be involved in the local regulation of blood flow) and (2) it can influence the oxygen demand
44  Regulation of  Tissue Oxygenation

through its inhibitory action on cytochrome c oxidase. The literature on this important biological
signaling molecule is extensive and continues to expand [55].

6.1.3 Role of  Myoglobin in Striated Muscle


The role of the oxygen-carrying protein myoglobin has been the subject of active study since the
1930s. The oxygen-binding curve for myoglobin is given by Eq. (4.5) with a Hill coefficient of
n = 1 and a P50 of about 5 mm Hg. Since it is composed of only a single polypeptide chain, there
are no allosteric effectors of oxygen binding to myoglobin, and the binding does not exhibit coop-
erativity. Myoglobin occurs primarily in cardiac muscle fibers and oxidative skeletal muscle fibers.
The first role identified for myoglobin was that as a storage reservoir for oxygen, to be used when
there was an immediate need for oxygen, such as in situations when the activity of a muscle would
abruptly increase. In the late 1960s, a second role for myoglobin was proposed, that of a facilita-
tor of oxygen diffusion within muscle fibers [139], especially during muscle contraction. Thus, the
diffusion of oxygen through the cytoplasm of a muscle fiber would be enhanced by the additional
amount that could be carried while bound to myoglobin if the diffusion coefficient of myoglobin
within the muscle cytoplasm was high enough. This facilitated diffusion role gained considerable
support from the work of Honig and co-workers [54] who employed cryomicrospectrophotometry
of myoglobin to measure the oxygenation of myoglobin and used Hill’s equation (see Eq. (4.5)) to
estimate intracellular PO2 from myoglobin oxygen saturation, SO2:

PO2 = P50 [SO2/(1 - SO2)]. (6.3)

The reported spatial resolution of the measurements was quite high—about 1 µm—so that
intracellular gradients in SO2 could be easily measured. Their results showed that in contracting
muscle, myoglobin oxygen saturation and hence intracellular PO2 were both small and uniform, a
predicted result if myoglobin facilitated oxygen diffusion. It was later reported [130] that the spatial
resolution of the technique was actually about 120 µm, considerably less than previously thought,
casting doubt on the use of the myoglobin SO2 data to support the facilitated diffusion role for myo-
globin. Proton magnetic resonance spectroscopy studies in human skeletal muscle have reported
small myoglobin SO2 values consistent with small PO2 gradients [82, 103]. However, given the low
spatial resolution of this approach and questions regarding the accuracy of the in vivo calibrations
[91], the facilitated diffusion role of myoglobin is still not certain. A role for myoglobin as a regula-
tor of NO bioavailabilty has been proposed recently [140]. The widespread distribution of myo-
Tissue Gas Transport  45

globin in cardiac muscle and oxidative fibers of skeletal muscle suggest an important role related to
tissue oxygenation.

6.2 OXYGEN TRANSPORT IN THE MICROCIRCULATION

From the first reports by Krogh [68, 69] on oxygen transport until the seminal work of Duling
and Berne [22] 50 years later, it was thought that most, if not all, oxygen exchange occurs across
the walls of capillaries in the peripheral circulation. However, following the finding of precapillary
oxygen losses [22], it has been recognized that oxygen can be exchanged between any two regions in
which a PO2 difference occurs. Although most (»98%) of the oxygen in blood is reversibly bound to
hemoglobin, the “driving force” for oxygen movement from blood to tissue is the PO2 difference that
exists across the vascular wall; percent oxygen saturation of the hemoglobin is not the driving force.
A description of oxygen exchange in arterioles and capillaries will be presented, along with several
controversial issues associated with oxygen transport in the microcirculation.

6.2.1 Longitudinal (Axial) Profile of Oxygen in Arterioles


Because Krogh [68–70] had emphasized the role of capillaries in the transport of oxygen from blood
to tissue cells, it was surprising to find that PO2 progressively fell along the arteriolar network [22,
126]. The interpretation of this longitudinal gradient in PO2 was that oxygen was diffusing from
the blood as it flowed through the arterioles. As a criticism of this finding, it was argued that PO2
measurements only sampled the dissolved oxygen, a miniscule amount compared to that bound to
hemoglobin in the RBCs, leaving open the possibility that most of the oxygen remained bound to
the hemoglobin until the blood reached the capillaries. With the development of microscopic tech-
niques to measure the hemoglobin oxygen saturation (SO2) of RBCs in situ [94, 95], it was found
that, indeed, SO2 fell in concert with PO2 according to the oxygen dissociation curve [90], indicating
that there was equilibrium between bound and free oxygen in the microcirculation.
This work was followed by a series of studies in which the diffusion of oxygen from arterioles
was quantified by measurement of geometric (diameter and length), hemodynamic (velocity and
hematocrit), and oxygenation (SO2) variables at upstream and downstream sites of unbranched ar-
terioles in four consecutive branching orders [71, 72, 119]. Comparison of the experimental results
with predictions of standard mathematical models of diffusion and consumption of oxygen in the
microcirculation led to the discovery of a substantial ten-fold discrepancy between experiment and
46  Regulation of  Tissue Oxygenation

theory, such that the apparent rate of diffusion of oxygen from arterioles was about ten times higher
than predicted [102]. The diffusion of oxygen from arterioles and the large discrepancy between
theory and experiment raised two related questions: (1) What was the destination of the oxygen?
(2) What was the explanation for the discrepancy in oxygen diffusion?
Three obvious destinations for oxygen which diffuses from arterioles are as follows: (1) the
parenchymal cells near the arteriole, where it is consumed; (2) other nearby microvessels (e.g., capil-
laries, venules), where it is picked up and carried away by the flowing blood; and (3) the arteriolar
wall, where it is consumed. The first possibility is reasonable since cells consuming oxygen would
not care whether the oxygen diffused across the walls of an arteriole or capillary. In dense vascular
networks, the occurrence of close proximity between arterioles and other vessels is commonplace,
so that the second possibility should receive serious consideration. Ellsworth and Pittman [31] re-
ported that capillaries in striated muscle often crossed paths and observed that capillaries acquired
oxygen by diffusion from nearby arterioles. A firm theoretical basis for this experimental finding was
later demonstrated by Secomb and Hsu [109].
The third possibility that a substantial amount of the oxygen could be consumed in the ar-
teriolar wall was initially dismissed by Swain and Pittman [119]. However, Tsai et al. [125] subse-
quently presented data of a large transmural PO2 difference in arterioles (~30 mm Hg), which they
interpreted as evidence of a major sink for oxygen located in the arteriolar wall (oxygen consump-
tion ~100–1000 times higher than that of resting skeletal muscle). This was in contrast to the
1–2 mm Hg transmural PO2 difference reported much earlier by Duling and Berne [22]. Based on an
extensive review of the literature on oxygen consumption by vascular tissue (endothelial cells, smooth
muscle cells, and vascular segments) and theoretical estimates of the maximum oxygen consumption
by mitochondria in the arteriolar wall, Vadapalli et al. [127] concluded that it was highly unlikely
that the arterioles could be responsible for consuming oxygen at much more than the rate for resting
skeletal muscle. Further careful examination of this report led to the conclusion that the particular
implementation of the phosphorescence quenching technique to measure PO2 (see Chapter 10) in
their study was flawed and resulted in an artificially low periarteriolar PO2 that caused the large
transmural PO2 [37, 124]. Thus, the third possibility that a large amount of oxygen is consumed by
the arteriolar wall has been ruled out.
The second question above, dealing with the explanation of the large discrepancy between
theoretical predictions and experimental observations of oxygen disappearance from arterioles, re-
mains a difficult one. The answer might be related to the way in which oxygen diffusion from ar-
terioles was calculated and the way in which the measurements used in the calculation were made.
The principle used to calculate oxygen diffusion from arterioles was a mass balance for the oxygen
entering and leaving an unbranched arteriolar segment of internal radius, R, and length, L. The
Tissue Gas Transport  47

rate at which oxygen diffused across the wall of the arteriolar segment (QO2D) was the difference
between the convective inflow and outflow of oxygen (QO2C at the upstream and downstream mea-
surement sites):

QO2D = QO2C (upstream) - QO2C (downstream). (6.4)

The convective flow of oxygen is given by the product of blood flow (Q = pR2 v) and oxygen con-
centration ([O2] = SO2 [Hb] CHb):

QO2C = Q [O2] = pR2 v SO2 [Hb] CHb, (6.5)

where v is average blood velocity, and the contribution of dissolved oxygen has been neglected. In
addition to the diameter measurement, each of the techniques involved in the measurements (i.e.,
velocity, hemoglobin concentration and oxygen saturation) was calibrated but only under conditions
of uniformity. One would not expect these three variables to be uniform under in vivo conditions in
flowing blood. Blood velocity has been found to be approximately parabolic [30]; the distribution of
RBCs and hence hemoglobin concentration should be non-uniform due to the Fahraeus effect [30];
and oxygen saturation should be non-uniform due to the loss of oxygen along arterioles [71, 72,
89, 90, 119]. The implicit assumptions made in the interpretation of the measurements were that
blood flow was axially symmetric through a vessel of circular cross-section and that the centerline
measurements of velocity, hemoglobin concentration, and oxygen saturation provided appropriate
average values for use in Eq. (6.3). However, it is possible that neglect of the non-uniform radial
profiles of velocity, hemoglobin concentration and oxygen saturation could introduce errors that
resulted in an overestimation of QO2D using Eq. (6.2). An analysis of the data and assumptions asso-
ciated with the calculation of QO2D showed that, in order to explain the 10-fold discrepancy between
predicted and observed oxygen diffusion from arterioles, there must be a systematic difference in the
profiles of oxygen saturation between the upstream and downstream measurement sites [89, 90]. At
the present time, this discrepancy remains unresolved, but it is anticipated that further studies in the
near future will result in a more complete understanding of this puzzle.

6.2.2 Longitudinal (Axial) Profile of Oxygen in a Capillary


Red blood cells pass through capillaries in single file due to the similar size of RBCs and the cap-
illary caliber. As RBCs flow through capillaries, oxygen is continuously released from the RBC
hemoglobin and eventually diffuses to the mitochondria where it is consumed. The sequential steps
48  Regulation of  Tissue Oxygenation

plasma RBC

HbO2

O2
O2

mitochondrion
(cytochrome oxidase)

cell

Figure 8: Pathway for oxygen from hemoglobin to mitochondria. Oxygen is released from hemoglo-
bin in the RBC, diffuses across the RBC membrane into the plasma, then crosses the microvessel wall
and through the interstitial fluid, eventually entering the mitochondria in the parenchymal cell where it
is consumed when it reacts with cytochrome c oxidase.

in the pathway for oxygen, from its release from oxyhemoglobin to its ultimate consumption in the
mitochondria, are illustrated in Figure 8.
Consider a small volume element of blood containing a red blood cell and adjacent plasma
as it passes through a capillary. Oxygen continuously diffuses from the plasma to the tissue where
it is consumed. Some of this oxygen is replenished by the oxygen which is released from Hb. If one
assumes that all the capillaries perfusing the tissue are identical and that the oxygen consumption of
the tissue is uniform, then the amount of oxygen removed (or extracted) from the volume of blood
per unit length along the capillary is constant. Thus, the amount of oxygen contained in the small vol­­
ume of blood decreases linearly from the arterial to the venous end of the capillary.
Diffusion is the mechanism by which oxygen passes from blood to tissue cells. The familiar
equation (Fick’s first law) for the rate of diffusive transport is:

DN/Dt = DA D(aPO2)/Dx. (6.6)

The amount of oxygen moving across the capillary wall is equal to that consumed by the tissue
supplied by the capillary. Note that changes in the number of perfused capillaries alter two factors
in this equation: A and Dx. The PO2 of blood along the capillary also declines, and the longitudinal
profile is related to the linear decrease in SO2 through the oxygen dissociation curve. Because of the
dissociation curve’s sigmoid shape, the PO2 profile along the capillary is expected to be nonlinear,
a prediction which has recently been confirmed by direct measurement using the phosphorescence
Tissue Gas Transport  49

quenching technique (see description in Chapter 10). The longitudinal SO2 profile along capillaries
has been determined in some tissues [26, 32].

6.2.3 Tissue Oxygen Transport: Krogh Cylinder Model


An accurate description of oxygen transport in tissues requires knowledge of several important
quantities: (1) tissue oxygen consumption—usually assumed to be constant and uniform [52, 69];
(2) tissue oxygen diffusion coefficient—usually assumed to be uniform and constant [52, 68, 69];
and (3) arrangement of capillaries—usually assumed to be uniform in diameter, length and spacing
[69]. This is an oversimplification of the actual circumstances in most cases, but these assumptions
allow one to analyze oxygen transport with relatively simple models [51, 99]. More realistic compu-
tational models of oxygen transport have been formulated to take into account recent information
on detailed three-dimensional reconstructions of microvascular networks [100].
Krogh [69, 70] described the first simplified model of capillary–tissue exchange for oxygen,
and its basic features are illustrated in Figure 9.
Since all capillaries were assumed to be identical and uniformly spaced (as they approximately
are in striated muscle), it was sufficient to consider a single capillary surrounded by a concentric
cylinder of oxygen-consuming tissue. Erlang, a colleague of Krogh, set up and solved the diffusion
equation for steady state oxygen exchange using this geometric layout. The result for the radial de­­
pendence of PO2 in the tissue is:

PO2 (r) = PCO2 + (VO2 / 4K ) [r 2 - RC2 - RT2 ln(r /RC)2], (6.7)

where PCO2 is the capillary PO2, VO2 is the oxygen consumption of the tissue, K is Krogh’s diffu-
sion coefficient for oxygen in the tissue, RC is the radius of the capillary, and RT is the radius of
the tissue cylinder. The lowest PO2 occurs at the edge of the tissue cylinder where r = RT. Based on
values for the variables in Equation 6.5, Krogh concluded that passive diffusion was sufficient to
account for observed oxygen transport and that only a small PO2 difference between blood and tis-
sue was necessary to supply an adequate amount of oxygen to the cells in the tissue cylinder. Since
oxygen is consumed in the tissue, the PO2 falls progressively between adjacent capillaries as shown in
Figure 10.
Taking into account the longitudinal decline in capillary PO2, the lowest tissue PO2 will be
located midway between adjacent capillaries at the venous end of the capillary. This location has
been given the name “lethal corner” because PO2 will be lowest there, and the “lethal corner” will be
the first tissue region to experience the effects of a compromised oxygen supply. If there is chronic
50  Regulation of  Tissue Oxygenation

Figure 9: Schematic of Krogh cylinder model and profile of PO2 along a capillary. PO2 in the blood
progressively falls from the arterial inflow to the venous outflow. At any axial location along the capillary,
PO2 falls in the radial direction in the tissue according to Eq. (6.5). From MM-S Lih [76]. Used with
permission of the publisher.

anoxia in the region of the “lethal corner,” irreversible tissue damage will occur. As would be ex-
pected, the heart and brain, organs with high metabolic rates, are the most sensitive to hypoxia due
to their reliance on oxidative metabolism.
Ellis et al. [27] suggested an alternative approach for describing oxygen transport from capil-
laries to muscle fibers in a contracted muscle based on corrosion casts of capillaries in this circum-
Tissue Gas Transport  51

Figure 10: Profile of PO2 between adjacent capillaries. “Adequate” oxygenation (normoxia) refers
to the case when the tissue PO2 is everywhere greater than the critical value needed for maximal mito-
chondrial ATP production. “Critical” oxygenation refers to the case when the PO2 in some microscopic
and negligible volume of the tissue just reaches the critical value needed for maximal mitochondrial ATP
production. “Inadequate” oxygenation (hypoxia) refers to the case when the PO2 is zero (anoxia) in a
significant volume of the tissue, below the critical value needed for mitochondrial ATP production. From
JB West [133]. Used with permission of the publisher.

stance. In this case, one can consider that the convoluted capillaries wrapped around the fiber are
similar to a uniform oxygen supply around the surface of a cylindrical muscle fiber as modeled by
Hill [52]. Thus, whereas Krogh considered a capillary source of oxygen embedded in a cylinder of
oxygen-consuming tissue, Ellis et al. considered a cylinder of oxygen-consuming tissue surrounded
by a uniform source of oxygen (i.e., the “web” of convoluted capillaries). Both models are worthy of
consideration in their respective situations.
In the context of modeling oxygen transport in capillaries, Krogh and most other authors
considered for simplicity that the blood flowing through capillaries was a uniform fluid, rather than
the realistic particulate suspension of RBCs in plasma that blood is. A consequence of this assump-
tion was that all the resistance to oxygen transport was associated with diffusion of oxygen through
the tissue, while the blood offered no resistance. Hellums [50] recognized that accounting for the
particulate nature of blood might alter the partitioning of resistance to oxygen diffusion between
the blood and tissue. He predicted that at normal hematocrit, about half of the resistance to oxygen
transport actually resided within the capillary, in large part due to the low solubility of oxygen in
the plasma. This effect manifested itself in a low PO2 in the plasma spaces between RBCs, lead-
ing to fluctuations in PO2 at an observation site as alternate regions of plasma (low PO2) and RBCs
(high PO2) passed by the site, so-called erythrocyte-associated transients (EATs) in PO2. Further
theoretical studies [33] confirmed and refined Hellums’ results to show that the separation between
52  Regulation of  Tissue Oxygenation

Figure 11: Erythrocyte Associated Transients (EATs) in PO2 in a capillary due to the passage of suc-
cessive RBCs (higher PO2) and plasma gaps (lower PO2). The existence of EATs is evidence that a large
fraction of the resistance to oxygen transport to tissue resides within the capillary, partitioned between
the intravascular and extravascular space roughly in proportion to the hematocrit.

adjacent RBCs and the low solubility of oxygen in the plasma were the key variables to be consid-
ered. Almost 30 years later, a method to measure PO2 in capillaries with the required spatial (»1
µm) and temporal (»10 ms) resolution was developed to provide experimental confirmation of the
existence of EATs [36]. Figure 11 illustrates the fluctuations in PO2 at the capillary level, based on
the findings of Golub and Pittman [36]. The low PO2 in the plasma space between RBCs effectively
reduces the surface area of the capillary that acts as a source of oxygen for diffusion to the paren-
chymal cells near it, thereby reducing the efficiency of oxygen transport to tissue. In principle, one
should be able, to elevate the plasma PO2 by replacing some of the plasma with a fluid (e.g., artificial
oxygen carrier, see Chapter 4) having a higher capacity than plasma to carry oxygen.

• • • •
53

chapter 7

Oxygen Transport in Normal and


Pathological Situations: Defects
and Compensations

One of the primary functions of the cardiorespiratory system, including the blood, is to ensure that
all tissues are adequately oxygenated at all times, i.e., that the PO2 in the immediate environment
of a cell exceeds the critical PO2 needed for normal mitochondrial oxygen consumption and ATP
production. Deviations from normal values of the key variables of oxygen transport typically lead to
hypoxic tissue environments. It is the role of various regulatory mechanisms in the cardiovascular
system, respiratory system and blood to ensure proper oxygenation of the tissues.

7.1 DESCRIPTION OF OXYGEN TRANSPORT USING


FICK’S PRINCIPLE

Fick (1870) enunciated a principle bearing his name which is a statement that mass is conserved. In
words, it states that “The amount of a substance consumed per unit time in a blood-perfused tissue
is equal to blood flow times the difference in arterial and venous concentrations of the substance.”
Applying Fick’s principle to oxygen transport, one finds that this can be stated succinctly in the
following expression:

VO 2 = Q ([O2 ]a - [O2 ]v ), (7.1)

where VO2 is oxygen consumption (ml O2 min-1 kg-1), Q is blood flow (ml min-1 kg-1) and [O2]a
and [O2]v are arterial and venous oxygen content (ml O2 dl-1), respectively. In words, oxygen con-
sumption equals blood flow times the a–v O2 difference. Fick’s principle is most often applied to
either the whole body, in which case Q represents cardiac output, or to an individual organ, in
54  Regulation of  Tissue Oxygenation

which case Q represents organ blood flow. In any computations using the above equation, one must
maintain awareness that the units “ml” in blood flow and “dl” in oxygen content differ by a factor
of 100.
It is convenient to modify the above equation by multiplying and dividing the right-hand side
by [O2]a / [O2]a and rearrange it to yield:

VO 2 = (Q [O2 ]a ) {([O2 ]a - [O 2 ]v ) /[O 2 ]a }. (7.2)

The term Q [O2]a is the amount of O2 delivered per unit time to the capillaries by arterial
blood flow, often called the “oxygen delivery” or “oxygen supply.” The second term of the product in
braces is the fraction of oxygen removed from the arterial blood during its passage through the cap-
illaries; this is also called “fractional oxygen extraction.” If one assumes that the respiratory system
is providing well-oxygenated arterial blood, then the oxygen delivery is set mainly by the arterioles,
since they primarily determine blood flow, and the oxygen extraction is set mainly by the capillaries,
since they primarily determine the exchange of oxygen between the blood and tissue.
Under normal or baseline conditions, the oxygen-related variables of Fick’s principle have
the following values for a typical adult human: (1) O2 delivery, Q[O2], is 5 1/min ´ 200 ml/l =
1,000 ml/min; (2) O2 extraction is 0.25 (25%), since [O2]a = 20 vol % (corresponding to PaO2 =
100 mm Hg) and [O2]v = 15 vol% (corresponding to PvO2 = 40 mm Hg); and (3) O2 consumption,
VO2, is 250 ml/min. There are then a number of pathologic states of hypoxia that diverge in one or
more ways from these baseline conditions. The types of   hypoxia discussed below are grouped accord­
ing to the primary origin of the defect: cardiovascular system, respiratory system, blood, or tissue cell.

7.2 STAGNANT HYPOXIA (HYPOPERFUSION)

Stagnant hypoxia, as its name implies, refers to situations in which blood flow is abnormally low as
occurs in shock, syncope or other “low-flow” states. In terms of oxygen transport, decreased blood
flow (hypoperfusion) is the primary limitation, and thus, the problem resides with the cardiovascu-
lar system. The defect in blood flow can be local (i.e., ischemic perfusion) or systemic (i.e., reduced
cardiac output, C.O.). Thus, O2 delivery is abnormal since Q is less than normal. In order to meet
the continuing demand for O2, the amount of O2 extracted from the blood must increase. This
means that [O2]v will fall, as will venous PO2. Since PaO2 is normal, this defect is not sensed by the
respiratory chemoreceptors (i.e., carotid bodies). Thus, increasing inspired oxygen fraction is not
Oxygen Transport in Normal and Pathological Situations  55

helpful to correct the problem. Interventions to improve cardiac output or peripheral blood flow
(use of vasodilators) would be appropriate.

7.3 HYPOXIC HYPOXIA

Hypoxic hypoxia occurs when the PO2 of arterial blood falls. This could occur because inspired PO2
is lower than normal (high altitude) or it could be due to a respiratory problem (e.g., hypoventila-
tion, diffusion impairment caused by pulmonary edema, ventilation–perfusion mismatch, or ana-
tomic shunt of blood past the gas exchange region). In terms of O2 transport, decreased arterial
blood oxygenation (hypoxemia) is the primary limitation, and thus, the problem resides with the
respiratory system. Oxygen delivery is abnormal since [O2]a is less than normal. The circulatory
system responds in two ways to improve tissue oxygenation. First, additional capillaries open to
reduce diffusion distances and increase the surface area for oxygen exchange; oxygen extraction sub-
sequently increases. Second, resistance vessels (arterioles) dilate in response to decreased tissue PO2
to increase perfusion and, hence, oxygen delivery. Venous oxygen content, [O2]v and PvO2 will be less
than normal due to the higher oxygen extraction. Since PaO2 is lower than normal (and presumably
lower than the 50 mm Hg threshold for respiratory chemosensory response), this defect is sensed by
the respiratory chemoreceptors (i.e., carotid bodies). Thus, increasing the inspired oxygen fraction
will be helpful except for the case of a pulmonary shunt.

7.4 ANEMIC HYPOXIA

Anemic hypoxia occurs when the oxygen carrying ability of the blood decreases, and thus, this
defect is specifically associated with the blood. This implies that fewer hemoglobin molecules (or
oxygen-binding sites) are available for binding oxygen. There can be several causes of this. The most
common example occurs with decreased hematocrit or true anemia. When the hemoglobin concen­
tration inside RBCs decreases, this also reduces the capacity of the blood to carry oxygen. Another
example is CO poisoning, in which there is virtually irreversible combination of CO with some heme-
binding sites on the hemoglobin molecule. Carbon monoxide binding produces the additional ad-
verse effect of a shift of the oxygen dissociation curve to the left (increased affinity of hemoglobin
for oxygen). Finally, the conversion of some heme-binding sites on hemoglobin to methemoglobin
renders those sites incapable of binding oxygen. This circumstance can occur when nitrites are used
56  Regulation of  Tissue Oxygenation

as vasodilators; iron is oxidized and changes from the ferrous to the ferric state. As with CO binding,
the presence of methemoglobin produces the additional adverse effect of a shift of the oxygen dis-
sociation curve to the left (increased affinity of hemoglobin for oxygen).
After neglecting the small amount of dissolved oxygen (~2%) in blood, the oxygen content
is equal to:

[O 2 ] = SO 2 [Hb]blood CHb = SO 2 Hct i[Hb]RBC CHb. (7.3)

The circulatory adjustments in response to anemia will be similar to those of the preceding
case. In order to maintain tissue oxygen consumption at baseline levels associated with a normal
oxygen carrying capacity of   blood, the reduction in oxygen delivery will lead to an increase in capil-
lary perfusion, and oxygen extraction will increase. Arteriolar dilation and viscosity reduction (for
the case of a reduction in Hct) will cause blood flow and oxygen delivery to increase. Both oxygen
extraction and oxygen delivery will continue to increase until the oxygen requirements of the tis-
sues are met or until the capacity to increase oxygen extraction and delivery has been reached. The
resulting situation is one in which venous oxygen content and PvO2 are less than normal. Since
PaO2 is normal for all the anemic situations considered, this defect is not sensed by the respiratory
chemoreceptors. Thus, increasing the inspired oxygen fraction is not helpful except for the case of
CO poisoning, where high inspired oxygen (e.g., 100% oxygen at ambient barometric pressure or
placement of the subject into a hyperbaric chamber) competes with CO binding at the heme site
(recall Haldane’s first law).

7.5 HISTOTOXIC HYPOXIA

Histotoxic hypoxia refers to a reduction in ATP production by the mitochondria due to a defect
in the cellular usage of oxygen. An example of histotoxic hypoxia is cyanide poisoning. There is a
profound drop in tissue oxygen consumption since the reaction of oxygen with cytochrome c oxidase
is blocked by the presence of cyanide. There are other chemicals that interrupt the mitochondrial
electron transport chain (e.g., rotenone, antimycin A) and produce effects on tissue oxygenation
similar to that of cyanide. Oxygen extraction decreases in parallel with the lower oxygen consump-
tion, with a resulting increase in venous oxygen content and PvO2. Although cyanide stimulates the
peripheral respiratory chemoreceptors, increasing the inspired oxygen fraction is not helpful, since
there is already an adequate amount of oxygen which the poisoned cells cannot use.
Oxygen Transport in Normal and Pathological Situations  57

7.6 SUMMARY OF HYPOXIC CONDITIONS AND RESPONSES

Table 4 summarizes the state of blood oxygenation and blood flow in the different situations dis-
cussed above (a hyphen means little or no change from normal values). The primary change is in­­
dicated by a double arrow (¯¯).
Note that “Stagnant” and “Anemic” hypoxias are not sensed by the respiratory chemore-
ceptors since PaO2 is normal in both cases. “Hypoxic” and “Histotoxic” (e.g., cyanide poisoning)
hypoxias are sensed. Is increased inspired PO2 helpful in reversing the hypoxia? It depends on the
cause of the hypoxia. For stagnant hypoxia, the answer is “no” since this type of hypoxia has its
origin in the cardiovascular system; PaO2 and [O2]a are not the problem. For hypoxic hypoxia, the
answer is “yes” for most cases (except pulmonary shunt) since this type of hypoxia has its origin in
the respiratory system. In the cases of low PIO2 (high altitude), decreased alveolar ventilation (VA),
diffusion limitation (e.g., emphysema, fibrosis) and VA/Q mismatch, increased FIO2 can elevate PaO2
and thereby compensate for the oxygen transport limitation in the respiratory system. In the case
of a pulmonary shunt, the answer is “no” since the problem is venous admixture, and the blood that
participates in gas exchange is already well oxygenated. For anemic hypoxia, the answer is “no” in the
case of reduced hematocrit or MetHb since the available binding sites for oxygen on hemoglobin are
already fully oxygenated. However, the answer is “yes” in the case of CO poisoning since high PO2
can displace CO from the hemoglobin. For histotoxic hypoxia, the answer is “no” since the problem
generally is due to a disturbance of oxygen usage by the cells and not oxygen supply.

Table 4:  Values and changes in variables for Fick’s principle analysis of   hypoxia

Type of Hypoxia [O2]a [O2]v C.O. VO2


(vol %) (vol %) (l/min) (ml/min)

Normoxia 20 15 5 250

Stagnant Normal ¯ ¯¯ Normal

Hypoxic ¯¯ ¯ � Normal

Anemic ¯¯ ¯ Normal or � Normal

Histotoxic Normal � Normal ¯¯


58  Regulation of  Tissue Oxygenation

Figure 12: Causes of   hypoxia from the perspective of the oxygen dissociation curve. The arterial and
venous oxygenation of the blood, as well as the arteriovenous differences, is shown for the normal situ-
ation, as well as for the four examples of hypoxia. From RL Riley [105]. Used with permission of the
publisher.

It is also instructive to look at these pathologies in terms of the oxygen dissociation curve.
Appropriate oxygen dissociation curves which correspond to the different causes of hypoxia are
displayed in Figure 12 [105]. Combined with an analysis of oxygen transport using Fick’s principle,
Eq (7.1) above, the results shown in Figure 12 are obtained.

• • • •
59

chapter 8

Matching Oxygen Supply to


Oxygen Demand

When the demand for oxygen by the cells in a tissue changes, it is important to be able to have a
corresponding change in the oxygen supply so that there is an efficient distribution of oxygen within
the organism. The overall regulation of tissue oxygenation starts with appropriate oxygen uptake
in the lungs, transfer of that oxygen by diffusion to the blood and then delivery of the oxygenated
blood to the various organs by the pumping action of the heart and bulk flow of oxygenated blood
through the complicated branching networks of blood vessels. Once the blood has made its way to
the microcirculation, transfer of oxygen to its final destination in the parenchymal cells takes place.
Despite the overall importance of the major systems involved in oxygen transport to the tissues—
the lungs, heart, blood, and major vessels connected to the organs—the final steps in regulation of
tissue oxygenation take place in the confined spaces occupied by the tissue cells and their associated
network of tiny microvessels, the microcirculation.

8.1 FICK’S PRINCIPLE

Fick’s principle was introduced in Chapter 6, and it can be used to advantage in the discussion sur-
rounding the matching of oxygen supply to oxygen demand. Fick’s principle, applied to an oxygen-
consuming, blood-perfused tissue, can be quantified by the following equation:

VO2 = Q {[O2]in - [O2]out}, (8.1)

where VO2 is oxygen consumption (or demand), Q is blood flow, [O2]in is inlet oxygen content or
concentration and [O2]out is outlet oxygen content or concentration. [O2] is determined by O2
bound to the hemoglobin in RBCs (»98% of total) since the dissolved oxygen in plasma and RBCs
is negligible.
It is important to recognize that Fick’s principle can be applied at the level of the whole or-
ganism, in which VO2 is whole body oxygen consumption, Q becomes cardiac output and [O2]in and
60  Regulation of  Tissue Oxygenation

[O2]out are arterial and venous oxygen contents, respectively. At the level of a single organ, VO2 is the
oxygen consumption of the organ, Q is the blood flow through the organ and the oxygen contents
are those of arterial and venous blood (recall the consequences of parallel vascular architecture of
the circulatory system). Since it is not difficult to define the extent of the tissue involved in these
two applications of Fick’s principle (i.e., whole organism or single organ), the interpretation of
data on oxygen transport from these perspectives is relatively straightforward. Application of   Fick’s
principle to collections of microvascular networks, a single network or a single microvessel, requires
careful consideration of the tissue volume supplied with oxygen by the network(s) or single vessel.
However, when appropriate boundaries are defined to match the tissue volume to the microvessels
supplying it with oxygen, Fick’s principle must work since it is based on the fundamental principle
of conservation of mass.

8.2 CONVECTIVE VS DIFFUSIVE OXYGEN TRANSPORT

Rearranging Fick’s principle, as was done in Chapter 6, leads to the following expression:

VO2 = Q [O2]in {[O2]in - [O2]out}/[O2]in. (8.2)

The term Q [O2]in represents the oxygen delivery or supply and is the convective transport of oxy-
gen determined mainly by arterioles. The term {[O2]in - [O2]out} / [O2]in represents oxygen extrac-
tion and is the diffusive transport determined mainly by capillaries. How does the cardiovascular
system match oxygen supply to changes in oxygen demand? The convective component of oxygen
transport can be modified by changing oxygen delivery or supply (Q [O2]in) through altered blood
flow brought about by changes in arteriolar tone [47]. The diffusive component of oxygen transport
can be modified by changing oxygen extraction ({[O2]in - [O2]out} / [O2]in) through altered oxygen
extraction in capillaries [47] brought about by changes in the so-called functional capillary density
or the surface area of capillaries in contact with RBCs.

8.3 MATCHING OXYGEN SUPPLY TO OXYGEN DEMAND:


ROLE OF ARTERIOLES AND CAPILLARIES

The transition from rest to exercise is a classic example of an increase in oxygen demand by the tis­­sue.
The arterioles are involved since there is an immediate need for increased blood flow to provide the
additional oxygen needed to satisfy the increased ATP hydrolysis and to prevent the accumula­tion
Matching Oxygen Supply to Oxygen Demand  61

of products of increased metabolism. The increased blood flow arises from the dilation of ar­­terioles,
the microvessels that control blood flow. The chemical signals responsible for arteriolar va­­sodilation
are derived from several sources: increased production of vasodilators from the active muscle tissue;
increased production of   NO due to increased shear stress communicated to the endothelium; and
spread of vasodilation to upstream sites due to the propagated response. The capillaries are involved
because there is an immediate need to extract more oxygen from the incoming blood. In the past,
recruitment of previously unperfused capillaries has been thought to play an important role; how-
ever, about 80% of the anatomically present capillaries are already perfused with RBCs [98]. The
current view is that recruitment of more surface area for oxygen exchange in the capillaries is the key
factor. There is a need to extract more oxygen from the incoming blood to the capillaries. Since most
(»98%) of the oxygen in blood is bound to the hemoglobin in RBCs, the surface area in contact
with RBCs is important. It has been observed that the RBC content (hematocrit) in capillaries in-
creases during muscle contraction, so that the capillary surface area in contact with RBCs increases.
This is another way of saying that the functional capillary density (FCD) increases during exercise.
Although a number of approaches have been used to quantify FCD, not all of them are aimed at
quantifying the surface area in contact with RBCs.

8.4 OXYGEN PROFILE ALONG A CAPILLARY: MASS BALANCE

It is instructive to consider the diffusion of oxygen from a small cylindrical volume element that
extends along a capillary of   length L from position x to x + dx [76,81]. Oxygen enters and exits the
volume element by convection (i.e., bulk flow) and also diffuses between the blood and interstitial
fluid (ISF) across the surface of the volume element. By considering the mass balance of oxygen
entering and leaving the volume element, the following expression is obtained:

Q {[O2(x + dx)] - [O2(x)]} = - aD (2pR / Dx) (PO2(x) - PISFO2 ) dx, (8.3)

where Q is blood flow, [O2] is oxygen content at the specified positions along the capillary, a is the
solubility of oxygen, D is the diffusion coefficient for oxygen, 2pR/Dx is the surface area per unit
length of capillary divided by the capillary wall thickness, and PO2 is the partial pressure of oxygen
at the specified locations (position x along the capillary and the ISF). It is assumed here that PO2
in the ISF can be considered uniform. Note that the left-hand side of this equation is the differ-
ence in convective flow of oxygen into and out of the volume element, whereas by mass balance, the
right-hand side is the rate at which oxygen diffuses between the blood and ISF (i.e., Fick’s first law
of diffusion).
62  Regulation of  Tissue Oxygenation

Figure 13: Approximation of the oxygen dissociation curve by a straight line through the origin and
the P50 point with slope equal to 0.5/P50. The oxygen dissociation curve was calculated based on Hill’s
equation (Eq. (4.5)) with P50 = 26.6 mmHg and Hill coefficient, n = 2.56. This linear approximation is
a good representation of the curve between SO2 of 20–80% (difference between straight line and curve
is < 5%).

In order to solve the above equation and determine the longitudinal oxygen profile along the
capillary, it is convenient to express [O2] in terms of PO2. Neglecting dissolved oxygen and using
the quasi-linear portion of the oxygen dissociation curve (i.e., the central portion above and below
P50), one can approximate [O2] as

[O2] » b¢ [Hb] CHb PO2, (8.4)

where b  ¢ is the slope of the oxygen dissociation curve (dSO2/dPO2 in mm Hg-1) in the quasi-linear
region, [Hb] is the blood hemoglobin concentration and CHb is the oxygen carrying capacity of the
hemoglobin. Figure 13 illustrates the linear approximation involved by expressing b  ¢ as 0.5 P50.
Matching Oxygen Supply to Oxygen Demand  63

Substituting for [O2] in the above mass balance Eq. (8.3) yields:

Qb [Hb] {PO2 (x + dx) - PO2(x)} = -aDg (PO2(x) - PISFO2) dx, (8.5)

where we define b as b  ¢ CHb and g  as 2pR/Dx. Using the definition of the first derivative of PO2
allows rewriting Eq. (8.5) as:

dPO2(x)/dx = -(aDg/Qb  [Hb]) (PO2(x) - PISFO2) (8.6)

and then as:

dPO2(x)/(PO2(x) - PISFO2) = -(aDg/Qb  [Hb]) dx. (8.7)

Integrating both sides of   Eq. (8.7) along the capillary from x = 0 (inlet) to some intermediate
point between the inlet and L (outlet) yields the longitudinal PO2 profile along the capillary:

ln{(PO2(x) - PISFO2) / (PO2(0) - PISFO2)} = -(aD g/Qb  [Hb])x = -(aDG/Qb  [Hb]) (x/L),


(8.8)

where G (= 2pR L/Dx) is the surface area of the capillary divided by its wall thickness. Finally, the
PO2 at the outlet (x = L) is given by:

PO2(L) = PISFO2 + {PO2(0) - PISFO2)} exp(-aDG/Qb [Hb]). (8.9)

These results allow one to express the Oxygen Extraction in terms of PO2. Since the oxygen
extraction is {[O2]in - [O2]out} / [O2]in, by substituting [O2] » b [Hb] PO2, we have

Oxygen Extraction = {(PinO2 - PISFO2) / PinO2}{1 - exp(-aDG/Qb [Hb])}. (8.10)

What factors can change to increase oxygen extraction and hence elevate oxygen consump-
tion? By examining the variables in the previous equation—PISFO2, a, D, G, Q, b, and [Hb]—for a
given value of inlet PO2, oxygen extraction will increase when PISFO2, Q, b, or [Hb] decreases; or a,
D, or G increases. Consider the case of exercise and ask what changes are known to take place in
these variables. PISFO2 initially decreases due to the increase in VO2; G and [Hb] increase due to the
increase of capillary surface area in contact with RBCs (the local hematocrit increases in capillaries
64  Regulation of  Tissue Oxygenation

during muscle contraction); Q increases due to arteriolar vasodilation to bring more oxygen and
wash out metabolic products; and b decreases due to the decreased slope of the oxygen dissociation
curve (elevated CO2 and H+ shift the dissociation curve to the right, the Bohr effect). Thus, the
decreased PISFO2 and b  will lead to increased oxygen extraction, while the increased Q will have
the opposite effect on oxygen extraction. Since G and [Hb] are both proportional to hematocrit and
thus their ratio will not change, to the approximation being used in this description, these variables
will not have an effect on oxygen extraction during muscle contraction.
It is of interest to consider two extremes of oxygen transport in the context of the present
description. Those extremes correspond to the ratio aDG/Qb [Hb] used to characterize oxygen
transport as being very large or very small. In the case where aDG/Qb [Hb] « 1, Fick’s principle
takes the form:

VO2 » (PinO2 - PISFO2) aDG (8.11)

so that oxygen uptake is independent of blood flow and is proportional to aDG, a measure of oxygen
diffusion. This limit is known as “diffusion-limited oxygen exchange” since more oxygen could be
delivered to the tissue if only diffusion (higher aDG ) could be increased. In the case where aDG/
Qb [Hb] » 1, Fick’s principle takes the form:

VO2 » Qb [Hb] (PinO2 - PISFO2) (8.12)

so that oxygen uptake is independent of diffusion (through aDG ) and is proportional to Qb [Hb], a
measure of oxygen delivery to the capillaries. This limit is known as “flow-limited oxygen exchange”
since more oxygen could be delivered to the tissue if only flow (higher Q) or [Hb] (more oxygen car-
riers) could be increased. Under physiological conditions, oxygen exchange in the peripheral tissues
is typically considered to be flow limited.

8.5 HETEROGENEITY OF BLOOD FLOW AND


OXYGEN DELIVERY

Most models of   blood flow and convective oxygen delivery in tissues make the simplifying assump-
tion that blood flow and the oxygen supply are uniform. Experimental observations of blood flow,
however, have shown that it is a variable with considerable spatial heterogeneity, as well as temporal
heterogeneity [61,79]. Ellsworth et al. [32] demonstrated heterogeneity of  blood flow and oxygen
Matching Oxygen Supply to Oxygen Demand  65

delivery in capillaries and used computational modeling to predict the consequences of such het-
erogeneity on tissue oxygenation. Not surprisingly, the greater the heterogeneity in oxygen supply,
the greater the heterogeneity in tissue oxygenation. Piiper and Scheid [86] compared homogeneous
and heterogeneous models of oxygen supply and assessed the consequences of the diffusion limita-
tion on tissue oxygenation. An interesting situation arises when one considers that both the oxygen
supply by the circulatory system and the oxygen demand by the tissue are heterogeneous. Is local
blood flow regulated in such a way as to match local oxygen supply and demand? Walley [131] pre-
sented an analysis of this circumstance and suggested that the answer is in the affirmative. Alders
et al. [1] provided experimental evidence that in the heart, oxygen con­­sumption is heterogeneously
distributed in parallel to heterogeneous oxygen delivery. If such coordination of the convective
oxygen supply to local oxygen demand occurs in general, then there would be a tendency for tissue
oxygenation to remain constant and uniform, an appearance of reg­­ulation of tissue oxygenation.

• • • •
67

chapter 9

Exercise and Hemorrhage

Exercise and hemorrhage are examples of high blood flow and low blood flow states, respectively,
which illustrate several features of the regulatory systems for maintaining tissue oxygenation. These
two situations will be analyzed in terms of Fick’s principle discussed in previous chapters.

9.1 EXERCISE

9.1.1 Fick’s Principle in Exercise


The problem the organism is faced with is to maintain a sufficient flow of substrates and oxygen
for metabolism to supply energy at a rate equal to the rate of its utilization. Since oxygen is often
the rate-limiting substance for tissue metabolism, the above requirement generally means providing
sufficient oxygen to maintain oxidative phosphorylation at a rate equal to ATP hydrolysis by muscle
(i.e., skeletal and cardiac muscle). Thus, when muscle contraction begins, the ATP is hydrolyzed to
form ADP and inorganic phosphate, Pi. In the presence of an adequate oxygen supply, the ADP is
rephosphorylated to replace the ATP that was utilized.
Considering the contracting muscle in isolation, the rate at which oxygen is delivered by bulk
blood flow is given by

QO2 = Q[O2]a, (9.1)

where QO2 is convective oxygen delivery, Q is blood flow, and [O2]a is arterial oxygen content (recall
that tissues are connected in parallel to yield blood of the same chemical composition at the inflow
of each organ or tissue). We will assume that [O2]a is normal and remains so during the period of
exercise. Thus, the only way in which oxygen delivery can increase is for Q to increase.
The oxygen consumption of the contracting muscle, VO2, is related to blood flow, Q, arte-
rial oxygen content, [O2]a, and venous oxygen content, [O2]v, by Fick’s principle as we have shown
previously:

VO2 = Q {[O2]a - [O2]v}. (9.2)


68  Regulation of  Tissue Oxygenation

This can be rewritten as:

VO2 = Q [O2]a {([O2]a - [O2]v)/[O2]a}. (9.3)

The quantity to the left of { } is the oxygen delivery and the term in { } is the fractional oxy-
gen extraction (i.e., the fraction of the incoming oxygen content that is utilized by the contracting
muscle). When the demand for oxygen increases, it can be met either by increased oxygen delivery,
increased oxygen extraction or both. Oxygen delivery depends on blood flow, which is primarily
determined by the contractile state of the small arteries and arterioles. Oxygen extraction depends
on how much oxygen can diffuse across the walls of the capillaries. Thus, oxygen extraction should
depend on the state of capillary perfusion and such factors as: (1) surface area for exchange (related
to the number of RBC-perfused capillaries); (2) maximum diffusion distance of oxygen (related to
average intercapillary distance); (3) PO2 difference between the capillary blood and the mitochon-
dria; and (4) amount of time blood is in the capillaries (inversely proportional to the velocity of red
blood cells). Oxygen transport during muscle contraction has been analyzed according to Fick’s
principle with the results interpreted in terms of a microvascular perspective [53].

9.1.2 Temporal Phases of Exercise


The temporal phases of the response to exercise are (note that in Figure 14 below containing the
intersecting cardiac function and vascular function curves that intersection point A corresponds to
the pre-exercise resting state): (1) Phase 1 (mechanical response); (2) Phase 2 (neural response); and
(3) Phase 3 (local response).
During Phase 1, the muscles, particularly of the abdomen, are tensed. The effect of the me-
chanical phase is to mobilize blood in the venous system, increasing venous return to the heart and
causing cardiac output to increase (intersection point B in Figure 14). During Phase 2, all parts of
the sympathetic neural response are activated: (1) vagal stimulation of the sinoatrial (SA) node is
decreased, producing increased heart rate; (2) sympathetic stimulation of the SA node is increased,
further increasing heart rate, and sympathetic stimulation of ventricular muscle is increased, pro-
ducing increased cardiac contractility and stroke volume; (3) sympathetic stimulation of arteries
and arterioles is increased, producing an increase in total peripheral resistance (TPR); and (4) sym-
pathetic stimulation of venules and veins is increased, further mobilizing blood to increase venous
return and hence cardiac output. Therefore, performance of both the cardiac and vascular systems
are enhanced, leading to increased cardiac output, most of which is needed to supply the increased
blood flow requirements of the heart and contracting muscles (intersection point C in Figure 14).
Exercise and Hemorrhage  69

Figure 14: Changes in cardiac and vascular function during different stages of exercise. The labeled
intersection points of the cardiac and vascular function curves show how cardiac output and right atrial
pressure change as the stage of exercise progresses from rest (A), to mechanical mobilization of blood
volume (B), to full sympathetic stimulation of the heart and vasculature (C), and finally to vasodilation in
the contracting muscles (D). From AC Guyton et al. [49]. Used with permission of the publisher.

The classic explanation of the local response of Phase 3, is that the increased frequency of muscle
stimulation leads to a loss of muscle K+ to the interstitium surrounding nearby resistance vessels. In
addition, the increased level of muscle work associated with exercise increases the rate of   breakdown
of ATP, some of which is lost from the cell as adenosine (ATP ® ADP ® AMP ® adenosine).
The increased level of muscle work can also lead to the breakdown of glycogen and the formation
of protons (i.e., H+) in association with the formation of lactate. The loss of K+, adenosine and H+
from muscle cells and their accumulation around arterioles cause relaxation of their smooth muscle
and increase the flow of blood through the capillaries, according to the classic hypoxic vasodilator
mechanism of blood flow regulation which leads to active hyperemia. These vasodilators also cause
the opening of unperfused portions of the capillary bed, thus increasing the number of RBC-
perfused capillaries (however, compare the more recent arguments against capillary recruitment in
exercise as described in Chapter 2). The effect of the latter on oxygen diffusion is to increase surface
area and decrease diffusion distance, both of which tend to increase oxygen extraction. Relaxation
70  Regulation of  Tissue Oxygenation

of the vascular smooth muscle of the arterioles also decreases TPR significantly, leading to a fur-
ther increase in cardiac output (intersection point D in Figure 14). The new NO/O2- radical pair
hypothesis of blood flow regulation described in Chapter 2 has been investigated during muscle
contraction using the fluorescent NO indicator diaminofluorescein-FM (DAF-FM) confined to
the interstitial space. It was shown that the kinetics of NO changes during muscle contraction were
consistent with the NO/O2- radical pair hypothesis, marking the first time that the kinetics for
any proposed vasodilator for active hyperemia has been measured in real time [44]. Note that the
increase in vasodilator concentration in the interstitial space overcomes the sympathetic constrictor
effect on the arterioles, but not on the venous vessels. Thus, there is vasodilation in the arterioles
of the active tissues (contracting muscles and heart), while the veins remain constricted and help to
maintain the elevated venous return.

9.1.3 Microvascular Approach to Oxygen Transport during Muscle Contraction


Measurements of the microcirculation during muscle contraction are typically problematic due to
the movement of the contracting muscle—fragile microelectrodes can be broken, and it is difficult
to maintain microvessels in good focus. Lash and Bohlen [74] used strengthened microelectrodes to
record perivascular and tissue PO2 (see Chapter 10) and found that tissue PO2 declined during con-
tractions. Smith et al. [113] used phosphorescence quenching microscopy (PQM, see Chapter 10)
to measure arteriolar, venular and tissue PO2 soon after contractions ceased and investigated the
kinetics of recovery of arteriolar diameter, RBC velocity and oxygenation following muscle con-
traction. Nugent et al. [85] followed up these findings by measuring both interstitial PO2 and local
VO2 before, during and after a bout of muscle contraction using PQM. As would be expected, PO2
rapidly decreased in the transition from rest to contraction and VO2 increased with a similar time
course. Poole and co-workers [4, 5, 64] also employed phosphorescence quenching measurements
of intravascular oxygenation, along with direct intravital microscopic determinations of capillary
perfusion, to assess the kinetics of oxygen transport at the onset of, during and following muscle
contractions. These innovative and challenging studies hold the promise of enhancing our under-
standing of the various factors involved in matching oxygen supply to increased oxygen demand.

9.1.4 Limited Oxygen Release from Red Blood Cells—Effect of  Transit Time
The velocity of RBC flow through capillaries is usually the lowest of all the vessels because the
capillaries offer the largest cross-sectional area of all the vessels in the circulatory network. Thus,
the rate at which oxygen is released from oxyhemoglobin in the RBCs as they traverse the typi-
cal capillary is almost never a limiting factor for tissue oxygenation. The average time needed for
Exercise and Hemorrhage  71

oxygen to be released from oxyhemoglobin is about 100 ms. So, as long as the RBCs spend at least
this much time in a capillary, oxygen release should not be a limiting factor. For a typical capillary
length of about 500 µm, the maximum velocity of an RBC should be about 0.5 mm/0.1 s or 5 mm/s
to ensure that oxygen unloading time from hemoglobin is not a problem. Gutierrez [48] predicted
that, under conditions of high RBC velocity as occur primarily in exercise, the oxygen release kinet-
ics from hemoglobin could limit oxygen diffusion to the tissue. Lash and Bohlen [75], using mea­­
surements of SO2 in venules of spontaneously hypertensive rats, found what they described as “excess
oxygen delivery” during muscle contractions. This interpretation arose from the unexpected finding
that venular SO2 did not fall much following muscle contraction, when one would expect the increased
energy demands of contraction to utilize a great deal of oxygen. A later report by Smith and co-
workers [113], in a similar study of muscle contraction in spontaneously hypertensive rats, described
the delayed return of venular PO2 to baseline values following contraction. While the results of these
two studies appear to be contradictory, they can be reconciled by considering that the RBC velocity
through the capillaries was too high (i.e., transit time was too short) during contractions to allow for
adequate oxygen release from hemoglobin. Thus, SO2 remained elevated (little oxygen release from
RBCs) while PO2 was depressed, due to its diffusion from the plasma and insufficient replacement
from RBC oxygen—an example of disequilibrium between hemoglobin SO2 and plasma PO2.

9.2 HEMORRHAGE

Hemorrhage refers to a significant loss of blood volume, enough so that arterial blood pressure falls
below normal baseline levels. The immediate effects of a significant reduction in blood volume are a
reduction in venous return to the right side of the heart, thereby reducing cardiac output and mean
arterial pressure, MAP (»CO ´ TPR). The decrease in cardiac output causes a generalized reduc-
tion in blood flow to all tissues, leading to a concomitant fall in the delivery of oxygen to all tissues.
If the loss of blood is severe enough, the delivery of oxygen might be too low to support oxidative
phosphorylation in the tissues, leading to a reliance on glycolysis to supply needed ATP. The in-
creased production of lactic acid and lower blood flow will lead to the buildup of this metabolite
and could compromise cell function and overall viability.

9.2.1 Fick’s Principle in Hemorrhage


Fick’s principle can be applied to the low-flow state of hemorrhage to gain some insight into the
alterations in tissue oxygenation that would be expected to take place. Equation (9.3) is also opera-
tive under conditions of hemorrhage. In this case and before any compensations take place, Q would
72  Regulation of  Tissue Oxygenation

be reduced from baseline, and [O2]a should be normal. If enough oxygen can be extracted from the
arterial blood, then baseline oxygen consumption could be supported, although this is typically not
the case. Thus, survival and maintenance of tissue oxygenation depend critically on the ability of
the organism to compensate for the adverse consequences of significant (~40% of blood volume or
greater) blood loss.

9.2.2 Compensatory Mechanisms in Hemorrhage


There are several compensatory mechanisms in response to hemorrhage that are classified by their
time course of action: fast, intermediate and slow. The rapidly acting compensatory mechanism (sec-
onds–minutes) involves the autonomic nervous system. The decrease in arterial pressure is sensed
by the baroreceptors in the carotid sinus region, where there is a decrease in the firing rate of the
carotid sinus nerve. This reduced firing rate decreases the activation of the vasomotor inhibitory cen­­
ter in the brainstem and leads to the activation of the vasomotor center and inhibition of vagal acti­­v­­
ity to the sinoatrial (SA) node, thereby raising heart rate. The increased activity of the sympathetic
nerves produces the following effects: (1) increased heart rate and cardiac contractility and (2) in-
creased arterial and venous contraction which cause increased TPR and venous return. The arterio-
lar constriction and the fall in arterial and venous pressures subsequent to blood loss cause a decrease
in capillary hydraulic pressure and, therefore, decreased fluid filtration (or increased fluid absorp-
tion). The fall in capillary hydraulic pressure will cause increased net fluid absorption (i.e., fluid flow
from the interstitium to the vascular space) because the driving pressure for hydraulic water flow out
of the capillary falls below the osmotic driving force for water flow into the capillary. This fluid shift
from the interstitium to the vascular space helps restore blood volume toward normal. However, the
combined effects of increased fluid absorption, increased cardiac output, and increased TPR cannot
return arterial pressure to normal.
The intermediate speed (hours to days) compensatory mechanisms rely on activation of the
renin–angiotensin–aldosterone system that involves the kidneys, blood, and lungs. The failure of
arterial pressure to return to normal will cause renal perfusion to remain below normal and, thus, in­­
crease renin secretion. Activation of this system will increase the retention of sodium, and indirectly
water, by the kidneys and expand the extracellular fluid volume which includes blood volume. This
will increase venous return and, thus, cardiac output and arterial blood pressure. Blood pressure is
generally returned close to normal by this part of the control system for restoring arterial pressure.
The slow compensatory (days to weeks) mechanisms involve restoration of red blood cell and
plasma protein concentrations. The expansion of   blood volume which occurs due to pressure-related
fluid shifts and the action of the renin–angiotensin–aldosterone system produces blood which is low
Exercise and Hemorrhage  73

in red blood cells (reduced hematocrit, hemodilution) and plasma proteins. These deficiencies of re­­
duced [O2]a and oxygen delivery are corrected by activation of the hematopoietic system and plasma
protein synthesis. The combination of low renal blood flow and low blood hematocrit causes the
release of erythropoietin. It is carried by the vascular system to the bone marrow where it stimu-
lates red blood cell maturation and their release into the blood until hematocrit returns to normal,
thereby restoring [O2]a. The perfusion of the liver by blood containing low concentrations of plasma
proteins activates protein synthesis in the liver. This increased protein formation returns the con-
centration of plasma proteins to normal.

9.2.3 Circulatory Shock and Resuscitation


Circulatory shock is a life-threatening condition, usually associated with an inadequacy of tissue
oxygen supply and demand. The microcirculation is directly involved, and the decreased vascular
blood flow may lead to global tissue hypoxia, finally resulting in multiple-organ failure or death. A
frequent cause of shock is acute hemorrhage, which is associated with high rates of morbidity and
mortality, especially when therapeutic treatment is delayed. The mechanisms by which perfusion
of peripheral tissues remains below pre-hemorrhage levels are complex and not well understood.
Tissue oxygenation depends on the balance between the diffusion of oxygen supplied by the red
blood cells and tissue metabolic demand. Moreover, cellular oxygen uptake is limited by the rate
of diffusion of oxygen to the mitochondria, as well as by the convective delivery of oxygen by mi-
crovessels. In situations such as hemorrhage, major systemic changes occur, and blood flow can be
diverted away from some tissues, such as the mesentery, in favor of others, such as the brain. Under
these conditions, a tissue with lowered blood flow may also show changes in its oxygen tension (PO2)
distribution. A better comprehension of the pathophysiology of hemorrhagic hypotension and of
the potential mechanisms behind the beneficial role of resuscitation fluids (see Artificial Oxygen
Carriers in Chapter 4) requires knowledge of the complex interactions between microhemodynam-
ics and tissue oxygenation under reduced blood pressure and flow.
In regard to regulation of tissue oxygenation in severe hemorrhage, serious and irreversible
damage to the tissues and the organism as a whole can occur before there is time to let the intrinsic
regulatory mechanisms perform their compensations. Thus, it is standard procedure to administer
resuscitation fluids to restore blood volume and hence arterial pressure and tissue perfusion more
rapidly. The composition of these fluids can have a significant impact on the rate and degree of
restoration of normal function. In addition to restoring blood volume and arterial pressure, restora-
tion of the oxygen-carrying function of the blood is often critical. Thus, the use of whole blood as a
resuscitation fluid is indicated when the blood type of the subject is known. Otherwise, crystalloid
74  Regulation of  Tissue Oxygenation

(i.e., electrolytes) or colloid (i.e., plasma or artificial plasma expanders) resuscitation fluids are used.
However, the oxygen-carrying capacity of these last two types of fluids happens to be quite low
(recall the low solubility of oxygen in aqueous media), and these fluids cannot be expected to re-
store tissue oxygenation. The use of artificial oxygen carriers, such as perfluorocarbon emulsions or
hemoglobin-based oxygen carriers described in Chapter 4, has the ability to restore both blood vol-
ume and the ability of the hemodiluted blood to provide adequate oxygen to the tissues.

• • • •
75

chapter 1 0

Measurement of Oxygen

Because of   the importance of blood and tissue oxygenation in health and disease, an accurate knowl-
edge of oxygen levels at these locations can play a critical role in the diagnosis of pathologies related
to an inadequate oxygen supply and in the progression of treatment. A number of different methods
have been developed over the years to make quantitative assessments of oxygen levels in the blood
and tissues. These methods can be used to quantify the oxygen of in vitro samples, but their applica-
tion for oxygen measurements under in vivo conditions allows one to gain important information
about the regulation of tissue oxygenation. The methods described below have been implemented
for measurements of oxygen levels in the microcirculation, but they can also be used in larger blood
vessels with appropriate modification.

10.1 OXYGEN TENSION (PO2)

10.1.1 Polarographic Electrodes


Determination of PO2 in microvessels and tissue is a fundamental measurement to provide impor-
tant information about the state of oxygenation at these sites. Although the currently preferred
method to measure PO2 is the phosphorescence quenching technique, described in detail below, for
historical purposes, it is useful to mention the use of polarographic electrodes to perform this mea-
surement [9, 22, 23, 56, 74, 111]. The recessed polarographic microcathode as described by Whalen
et al. [134] has been used in numerous microcirculatory studies where measurements of PO2 were
made. Schneiderman and Goldstick [107] carried out an extensive theoretical analysis of the design
characteristics of the recessed-tip microcathode (<5 mm in diameter), which confirmed the empirical
data collected by others, and concluded that the presence of a recessed tip whose length was about
10 times the diameter of the tip region gave the electrode the property that it minimally disturbed
the local PO2 at the measurement site. The catalytic surface of the electrode is polarized at -0.7 V
for the measurement of oxygen. An electrode is calibrated by immersing it in a solution similar in
composition to the chemical and thermal environment of the actual measurement site and which
has been equilibrated with a sequence of calibrated gas mixtures. The relationship between PO2 and
current through the electrode should be linear and stable over the course of the measurements.
76  Regulation of  Tissue Oxygenation

10.1.2 Phosphorescence Quenching Microscopy


Phosphorescence quenching microscopy (PQM) is a relatively new method for measuring PO2
non-invasively with a high degree of spatial and temporal resolution [36, 37, 121, 129]. A phosphor
molecule can be excited from the ground state by application of a brief flash of light (i.e., time-
domain method) or a continuous oscillatory illumination (i.e., frequency-domain method) of the
appropriate wavelength. The excited phosphor can return to the ground state by emitting a photon
(i.e., phosphorescence emission) or by transferring energy to a nearby oxygen molecule (i.e., colli-
sional quenching) with a lifetime on the order of hundreds of milliseconds. Since oxygen is the prin-
cipal quenching molecule in biological tissues, the following relationship between phosphorescence
lifetime and tissue PO2 can be readily established. The oxygen dependence of the phosphorescence
decay is described by the Stern–Volmer equation:

1 /t = 1 /t0 + kqPO2, (10.1)

where t is the lifetime at the measured PO2, t0 is the phosphorescence lifetime in the absence of
oxygen and kq is the quenching coefficient. This equation can also be written in terms of decay rate,
k, rather than lifetime, where k = 1/t and k0 = 1/t0. Under conditions where PO2 is uniform, the time
course of phosphorescence decay is monoexponential in time so that the intensity of the phospho-
rescence light emission, IPHOS(t) is given by:

I PHOS (t) = I PHOS (0) exp( -kt ), (10.2)

where IPHOS(0) is the initial phosphorescence intensity (at t = 0) and k is the phosphorescence decay
rate, related to PO2 by the Stern–Volmer equation, expressed in terms of phosphorescence decay
rates:

k = k0 + kqPO2. (10.3)

So, determining PO2 from k is a matter of measuring the decay rate and using the values k0
and kq determined independently in calibration experiments.
This procedure is valid for situations in which the PO2 is homogeneous, as in a uniform
calibration medium; however, under most in vivo conditions, there are gradients in PO2 (e.g., in­
terstitium, luminal blood) that must be taken into account. This is particularly true for skeletal
muscle during changes in metabolic demand when a dynamic gradient in PO2 exists between the
vascular supply and respiring tissue. This heterogeneity can cause significant underestimates of PO2
when calculations are made using a monoexponential time course. A simple approach to deal with
Measurement of Oxygen  77

this problem was proposed by Golub et al. [43] in which a Rectangular distribution model of PO2
was used to approximate the distribution of PO2 within the measured region. Since the actual dis-
tribution of PO2 is generally not known, the assumption that all PO2 values between a minimum,
PMIN, and a maximum value, PMAX, are equally represented within the measurement region is the
approximation that allows one to express the time course of the phosphorescence decay in terms of
the average PO2, M = (PMAX + PMIN) / 2, and the half-width of the Rectangular distribution, d =
(PMAX - PMIN) / 2. This approach mitigates the systematic error in PO2 that would otherwise occur
by accounting for regional variability in PO2 within the measurement region. The mean PO2, M, can
be calculated by fitting the phosphorescence decay curve to the following equation and determining
the best-fit value for M:

IPHOS(t) = IPHOS(0) exp[-(k0 + kqM )t] sinh(kqdt) / kqdt, (10.4)

where sinh is the hyperbolic sine function. For a full description of the derivation of this equation,
see [43].
An example of a phosphorescent probe used to determine PO2 is Pd-meso-tetra-(4-
carboxyphenyl) porphyrin, which is typically bound to albumin to minimize its passage between the
intravascular and interstitial spaces. For intravascular measurements of PO2, the probe can be infused
into the circulatory system where it resides for several hours. Because the vascular wall in some cir­­
culatory beds (e.g., intestine, skin, skeletal muscle) has a finite permeability to even a molecule as large
as albumin, it is expected that there could be some extravasation of the probe into the interstitium, so
that the phosphorescence signal might arise from both intravascular and extravascular compartments.
For interstitial measurements of PO2, the probe can be topically applied to the tissue of interest. Fol­­
lowing an approximately 30-minute incubation period, the probe can penetrate 200–500 mm through
the interstitium into the tissue so that PO2 in the upper layer of a thick tissue or full thickness of a
thin tissue can be determined. It has been established that the probe concentration does not affect the
phosphorescence lifetime, so that measurements of PO2 are insensitive to minor local variations in
probe concentration and signal amplitude [129]. The probe has light absorption peaks at 415 nm and
524 nm (the Soret and Q bands, respectively) and an emission peak at 690 nm [24], so that the probe
can be excited at either or both absorption bands and the phosphorescence emission conveniently de­­
tected at a wavelength region far from the excitation wavelength(s).
An important feature of the phosphorescence quenching technique to measure PO2 is that,
when the excited probe interacts with oxygen, the highly reactive oxygen species singlet oxygen
is produced [96, 129]. The singlet oxygen can then react with nearby organic molecules, effec-
tively consuming the original oxygen molecule and lowering the PO2. The magnitude of the oxygen
consumption for each excitation flash is roughly proportional to the probe concentration and the
78  Regulation of  Tissue Oxygenation

intensity of the excitation flash. If a sufficient time is not allowed for oxygen to diffuse into the
excited region of radius, R (t  DIFF » R2/2D) [14], before the next excitation flash (t  EXCIT = 1/F,
where F is the excitation frequency), then the PO2 in the excitation region will decrease. Measures
need to be put in place to account for the oxygen consumption and steps taken to minimize it (e.g.,
by reducing probe concentration, flash intensity, flash frequency or radius of excitation region);
otherwise, PO2 at the measurement site will be artifactually reduced. This issue of photoconsump-
tion of oxygen is addressed by Golub and Pittman [37, 124] using comparisons between different
implementations of the method [112, 125] and describing the consequences of inappropriate mea-
surement parameters. Photoconsumption of oxygen by the method is particularly troublesome for
repeated measurements of a stationary volume of tissue (e.g., interstitial fluid) and can also lead to
PO2 measurement error in cases of intravascular measurements when the blood velocity falls below
a critical value.
A working microscopic setup used to perform phosphorescence quenching measurements
of PO2 is diagrammed schematically in Figure 15. In this implementation of PQM, the size of the
excitation region is small compared with that of the detection region, providing good signal-to-
noise ratio. In addition, this PQM setup has a scanning mode so that successive measurements of
PO2 occur at different sites thereby minimizing and practically eliminating the impact of the photo-
consumption artifact.
A recent application of phosphorescence quenching microscopy is the localized measurement
of tissue oxygen consumption [45]. The phosphorescent oxygen probe is topically applied to the
tissue as described above, and blood flow is abruptly arrested (within »0.1 s) by inflating a transpar-
ent bag attached to the end of the microscope objective at the measurement site. At the time when
the oxygen supply has been removed by flow arrest and extrusion of RBCs from the compressed
microvessels, oxygen continues to be consumed by mitochondria in the underlying tissue, and the
linear decrease in PO2 at the measurement site is related to oxygen consumption by

VO2 = aO2 dPO2/dt, (10.5)

where aO2 is the oxygen solubility of the interstitial fluid and dPO2/dt is the rate of change of PO2 or
slope of the oxygen disappearance curve at the measurement site. Because only a brief arrest (»5 s)
of flow, followed by a 15-s recovery period, is needed to obtain an accurate estimate of the oxygen dis­
appearance rate, it is possible to make several measurements of VO2 per minute (3 measurements per
min in this scenario), allowing an investigation of the oxygen kinetics in response to interventions,
such as muscle contraction or ischemia/reperfusion. Shortening the flow arrest/recovery cycle will
lead to an improvement in the ability of this approach to assess kinetic changes in VO2 with better time
resolution. Also, by arresting flow over a longer time interval and making rapid PO2 measurements
Measurement of Oxygen  79

DISABLE
SIGNAL COMPUTER
CLOCK With
11 Hz Lab VIEW

AMPLIFIER
PMT

BARRIER
LASER

FILTER

SWITCH
MIRROR VIDEO
TV MONITOR

DICHROIC
DOVE
MIRROR
PRISM

DVR

OBJECTIVE
SCANNING
MIRROR
10 Hz

VESSEL

TRANSILLUMINATION

Figure 15: Setup used for phosphorescence quenching microscopy (PQM). The phosphor is excited
by pulsed laser illumination through the microscope objective. The phosphorescent emission is filtered
and then detected by the photomultiplier tube (PMT). The amplified signal from the PMT is then
acquired, displayed and analyzed by the computer. From AS Golub et al. [42]. Used with permission of
the publisher.

during this time, VO2 can be calculated according to Eq. (10.5), and the pairs of PO2 and VO2 values
can be used to determine the PO2 dependence of VO2 [38].

10.2 HEMOGLOBIN OXYGEN SATURATION (SO2)

10.2.1 Spectrophotometry of   Hemoglobin


The color of blood changes dramatically from the bright red of well-oxygenated blood to the deep
red, almost purple hue of deoxygenated blood. These color changes are a manifestation of differences
80  Regulation of  Tissue Oxygenation

in the absorption spectra of oxy- and deoxy-hemoglobin in the visible spectrum. These differences
can be utilized to quantify the relative amounts of oxy- and deoxy-hemoglobin, as well as other spe-
cies normally present in blood (e.g., COHb and metHb). The basis for the spectral determination
of hemoglobin concentration is the Beer–Lambert law of light absorption:

I = I 0 10 -e cl , (10.6)

where I is the intensity of light transmitted through the solution of hemoglobin of concen­tration, c,
and path length, l, I0 is the incident light intensity and e is the extinction coefficient of hemoglobin
at the measuring wavelength, a measure of the intrinsic absorption of light by hemoglobin.
The variation of e with wavelength, l, is considered to be the absorption spectrum of the
particular species of hemoglobin, as illustrated in Figure 16. Tables and graphs of e vs. l for the vari-
ous hemoglobin species are available in the literature [128].
It is customary to define the optical density, OD, as:

OD = log (I0  /I ) = e cl, (10.7)

where log is the common logarithm (i.e., to base 10). For a mixture of oxy- (HbO2) and deoxy-
(Hb) hemoglobin being observed at wavelength, l, Beer–Lambert’s law takes the form:

ODl = el1 [HbO2]l + el0 [Hb]l, (10.8)

where el1 (“1” for fractional SO2 of 1) is the extinction coefficient for HbO2 and el0 is the extinction
coefficient for Hb (“0” for fractional SO2 of 0). Since the total hemoglobin concentration, [Hb]T,
is the sum of the two components [HbO2] and [Hb], and SO2 is defined as [HbO2]/[Hb]T, this
expression for the optical density can be written as:

ODl = el1 SO2 [Hb]T l + el0 (1 - SO2)[Hb]T l. (10.9)

Inspection of the absorption spectra of oxy- and deoxy-hemoglobin over the visible spectrum
(i.e., l » 400–700 nm) reveals that there are a number of wavelengths at which the extinction co-
efficients for oxy- and deoxy-hemoglobin are equal; such wavelengths are referred to as isosbestic
wavelengths. The determination of SO2 and [Hb]T for a hemoglobin sample requires OD measure-
ments at a minimum of two wavelengths, at least one of which must be non-isosbestic. A common
implementation of such measurements is to have one of the wavelengths an isosbestic one, lI, and
Measurement of Oxygen  81

Figure 16: Optical absorption spectra for oxy- and deoxy-hemoglobin. Relationship between ex-
tinction coefficient (Beer-Lambert law) and wavelength is shown for oxyhemoglobin (solid curve) and
deoxyhemoglobin (dashed curve). From Pittman RN [87]. Used with permission of the publisher.

the other a (“measuring”) wavelength at which there is a large difference between the extinction
coefficients for oxy- and deoxy-hemoglobin, lM. For this case, the Beer–Lambert law becomes for
the two wavelengths:

ODI = e I1 SO2 [Hb]T l + e I0 (1 - SO2)[Hb]T l = e I [Hb]T l (10.10)

since by definition e I1 = e I0 = e I and

ODM = e M1 SO2 [Hb]T l + e M0 (1 - SO2)[Hb]T l. (10.11)

The OD at the isosbestic wavelength can be used to determine [Hb]T as

[Hb]T = ODI/e I l (10.12)


82  Regulation of  Tissue Oxygenation

and SO2 can be determined from the ratio of the two ODs as

SO2 = [e I/(e M1 - e M0)] (ODM/ODI) - e M0/ (e M1 - e M0), (10.13)

where the total hemoglobin concentration, [Hb]T, and path length, l, cancel out in the calculation
of SO2. This approach to determining SO2 is independent of both [Hb]T and l.
The choice of wavelengths to use in determining SO2 depends on the product of hemoglobin
concentration and path length, [Hb]T l [87, 88]. The specific relationship can be derived first by
considering the sensitivity of ODM to changes in SO2 as given by:

dODM/dSO2 = (e M1 - e M0) [Hb]T l. (10.14)

To compare the sensitivity index, dODM/dSO2, at different wavelengths, the path length, l,
should be chosen to make the midpoint of the oxygen saturation scale (SO2 = 0.5) the same for all
wavelengths. Using the value of 0.434 for OD that gives minimum error [87] and SO2 = 0.5 in the
equation for ODM above, we find

l = 0.868 / (e M1 + e M0)[Hb]T. (10.15)

Substituting this value into the sensitivity equation yields

dODM/dSO2 = 0.868(e M1 - e M0)/ (e M1 + e M0). (10.16)

As e M takes on values from 400–700 nm, the sensitivity index is zero at all isosbestic wave-
lengths since the numerator of this fraction is zero at those wavelengths. In between the neighbor-
ing isosbestic wavelengths, the sensitivity index increases to local maxima and identifies usually
narrow wavelength ranges that are quite sensitive to changes in SO2, as illustrated in Figure 17.
An additional and important factor which must be taken into account is the range of values
for [Hb]Tl expected to be encountered in an experimental setting. By setting [Hb]T to an expected
value for blood at a normal hematocrit for the species and conditions under consideration (e.g., Hct
of 50%; [Hb]T = 10 mM, heme basis), one can determine the “optimum” path length as a function
of wavelength, as illustrated in Figure 18 [87]. This analysis indicates that determination of SO2 for
vessels in the 5–20 mm range (capillaries, smaller arterioles and venules) should employ wavelengths
in the 400–450 nm range (blue light); for vessels in the 20–100 mm range (larger arterioles and ve­­
nules), wavelengths in the 500–600 nm range (green light) should be used.
Measurement of Oxygen  83

Figure 17: Spectral sensitivity to changes in hemoglobin oxygen saturation. See Eq. (10.16) and the
associated text for a description of the derivation of the spectral sensitivity index. The wavelengths at
which the sensitivity index falls to zero correspond to isosbestic wavelengths in the hemoglobin absorp-
tion spectra. From Pittman RN [87]. Used with permission of the publisher.

The considerations about light absorption by hemoglobin apply to the contents of red blood
cells, but because there is a mismatch between the refractive index of the RBC hemoglobin and the
plasma in which they are suspended, light is scattered by the RBCs. Light scattering decreases the
amount of light collected by the photodetector (e.g., video camera, photomultiplier, photodiode),
and this manifests itself in a higher optical density than would otherwise be expected for a hemo-
globin solution. It has been found that light scattering by RBCs is independent of wavelength in
the 500–600 nm range [88, 94], so that the contribution of light scattering can be included in the
measured OD by adding a term B which depends on hematocrit. For measurements made at a given
hematocrit, B should be constant and can be determined by making an OD measurement at an ad-
ditional wavelength. This is the basis for the three-wavelength method to measure SO2 as described
84  Regulation of  Tissue Oxygenation

Figure 18: Optimum pathlength for spectrophotometric measurement of hemoglobin for a hemo-
globin concentration of 10 mM and oxygen saturation of 50%. See the description in the text associated
with the derivation of Equation 10.15. This graph shows that wavelengths in the blue region of the
spectrum are most appropriate for vessels with diameter less than about 20 mm, whereas wavelengths in
the green region of the spectrum are most appropriate for vessels in the diameter range of 20 to 100 mm.
From Pittman RN [87]. Used with permission of the publisher.

by Pittman and Duling [94, 95]. This approach has also been used to determine hemoglobin con-
centration in microvessels [77]. Capturing and analysis of video images at appropriate wavelengths
has been employed to provide a more large-scale view of oxygen saturation in small networks of
microvessels and to yield information on the potential transfer of oxygen among nearby microves-
sels [66, 120]. With recent advances in technology, it is possible to use rapid, multi-wavelength de-
terminations of light transmission through blood vessels and fit the resulting multi-point spectra to
linear combinations of known absorption spectra of the various hemoglobin species [118], thereby
providing more accurate assessments of blood oxygenation.
Measurement of Oxygen  85

The oxygenation of blood flowing through capillaries is a special case since RBCs pass
through them in a single file, and the SO2 of individual RBCs can be determined. It has been found
that light scattering by single RBCs is not as problematic as for larger vessels where a light beam
traverses multiple RBCs between the light source and the detector. Thus, only two wavelengths
in the blue region of the spectrum are needed to calculate SO2. This two-wavelength method to
measure SO2 of individual RBCs in capillaries has been implemented by using a dual-camera video-
computer method [25]. More recently, an integrated image analysis system has been developed,
based on this principle, which determines SO2 of single RBCs from the same images used for he-
modynamic measurements [58, 59].

10.2.2 Resonance Raman Spectroscopy of   Hemoglobin


A recent addition to the techniques for measuring blood oxygenation in microvessels uses Raman
microspectroscopy of hemoglobin. The technique offers several advantages over other currently used
techniques by providing noninvasive and reliable in vivo determinations of SO2 in thin tissues, as
well as in solid organs and tissues, which are unsuitable for techniques requiring transillumination
[122, 123]. The spectroscopic basis for this technique is the resonant Raman enhancement of hemo­
globin in the violet/ultraviolet region, allowing simultaneous identification of oxy- and deoxyhe-
moglobin with the same excitation wavelength. Laser light is focused onto small areas (~20 µm in
diameter) containing hemoglobin, and Raman spectra are obtained in backscattering geometry us-
ing a microscope coupled to a spectrometer and a cooled detector. Calibration of the method is per­­
formed in vitro using hemoglobin solutions and red blood cell suspensions at several hemoglobin
concentrations, equilibrated at various PO2. SO2 can be calculated from the Raman band intensities
at 1360 cm-1 and 1375 cm-1 (the oxidation state marker band n4). The method allows SO2 determi­
nations over the range of arterioles, venules, and capillaries. Taking advantage of the resonant Raman
enhancement of hemoglobin in the Q-band region, a system for measuring SO2, using 532-nm ex­­
citation, has also been implemented [123]. The SO2 in microvessels can be estimated by measuring
the intensity of Raman signals of band n4, as well as from the n19 and n10 bands (1500–1650 cm-1
range).
An advantage of the optical techniques described here for the determination of PO2 or SO2
is that they can be combined with other optically based techniques to measure various key variables
in oxygen transport, such as diameter, erythrocyte velocity and hemoglobin concentration.

• • • •
87

chapter 1 1

Summary

Cells in an organism depend on a continuous supply of oxygen to produce the energy needed to
perform their myriad functions, so that the regulation of tissue oxygenation is a key activity of any
organism. This regulation involves the coordinated action of the cardiovascular and respiratory sys-
tems, as well as the blood which carries the oxygen to the vicinity of cells. The utilization of oxygen
is a local phenomenon, with each cell having its own specific needs which can change over time
according to the current activity of the cell. Because of the local nature of oxygen utilization, inte-
gration and matching of oxygen supply to oxygen demand take place in the microcirculation. The
mechanism(s) by which parenchymal cell activity, and hence oxygen needs, are communicated to
nearby vessels is still not well understood, despite over a century of intense study. The recent emer-
gence of the NO/O2- radical pair mechanism provides a simple and novel explanation for carrying
out this important regulatory function. An overall summary of the diffusional interactions among
different vessels in the microcirculation is shown in Figure 19.
Altered demand for oxygen is sensed in various ways and is communicated to the terminal
vasculature which responds by changing blood flow in a way that satisfies the oxygen demand. An im-
balance between oxygen supply and oxygen demand is a hallmark of a number of (patho)physiological
conditions, such as aging, exercise, hemorrhage, sepsis, diabetes, anemia, and heart failure. As our un­­
derstanding of oxygen transport increases, based on modern techniques of measuring oxygen, it can
be anticipated that new therapeutic interventions that increase the oxygen supply (e.g., artificial oxy­­
gen carriers) or reduce oxygen demand (e.g., inhibition of tissue metabolism) will become available
to aid in the regulation of tissue oxygenation.
88  Regulation of  Tissue Oxygenation

Figure 19: Pathways for oxygen diffusion in microvascular networks. Whereas Krogh considered
oxygen diffusion from capillaries to the tissue (lower left of figure), current knowledge about oxygen
diffusion in the microcirculation shows that oxygen can diffuse between any regions in close proximity
where PO2s are different. From Ellsworth ML et al. [29]. Used with permission of the publisher.

• • • •
89

References

[1] Alders DJ, Groeneveld AB, de Kanter FJ, van Beek JH. Myocardial O2 consumption in
porcine left ventricle is heterogeneously distributed in parallel to heterogeneous O2 delivery.
Am J Physiol Heart Circ Physiol 287: H1353–61, 2004. doi:10.1152/ajpheart.00338.2003
[2] Bagher P, Segal SS. Regulation of blood flow in the microcirculation: Role of conducted
vasodilation. Acta Physiol (Oxf ). 2010 doi: 10.1111/j.1748-1716.2010.02244.x.
[3] Barcroft J, Hill AV. The nature of oxyhaemoglobin, with a note on its molecular weight.
J Physiol 39: 411–28, 1910. doi:10.1113/jphysiol.1910.sp001350
[4] Behnke BJ, Barstow TJ, Kindig CA, McDonough P, Musch TI, Poole DC. Dynamics of
oxygen uptake following exercise onset in rat skeletal muscle. Respir Physiol Neurobiol 133:
229–39, 2002. doi:10.1016/S1569-9048(02)00183-0
[5] Behnke BJ, Kindig CA, Musch TI, Koga S, Poole DC. Dynamics of microvascular oxygen
pressure across the rest-exercise transition in rat skeletal muscle. Respir Physiol 126: 53–63,
2001. doi:10.1016/S0034-5687(01)00195-5
[6] Bernard, C. Lecons sur les phenomenes de la vie communs aux animaux et aux vegetaux.
Cours de Physiologie generale du Museum d’Historie Naturelle. Paris, Baillere et fils, 1878.
doi:10.5962/bhl.title.44802
[7] Boron WF, Boulpaep, EL. Medical physiology, 2nd ed., Philadelphia: Saunders, 2008.
[8] Brown GC. Nitric oxide and mitochondrial respiration. Biochim Biophys Acta 1411: 351–69,
1999. doi:10.1016/S0005-2728(99)00025-0
[9] Buerk DG. Measuring tissue PO2 with microelectrodes. Meth Enzymol 381: 665–90, 2004.
doi:10.1016/S0076-6879(04)81043-7
[10] Chance B, Williams GR. Respiratory enzymes in oxidative phosphorylation. I. Kinetics of
oxygen utilization. J Biol Chem 217: 383-393, 1955.
[11] Chang TMS. Future Prospects for Artificial Blood. Trends Biotechnol 17: 61–67, 1999.
doi:10.1016/S0167-7799(98)01242-6
[12] Cooper CE, Giulivi C. Nitric oxide regulation of mitochondrial oxygen consumption. II.
Molecular mechanism and tissue physiology. Am J Physiol Cell Physiol 292: C1993-C2003,
2007. doi:10.1152/ajpcell.00310.2006
90  Regulation of  Tissue Oxygenation

[13] Costanzo LS. Physiology, 4th ed. Philadelphia: Saunders Elsevier, 2010.
[14] Crank J. The mathematics of diffusion. Oxford: Clarendon Press, 1956.
[15] Creteur J, Sibbald W, Vincent JL. Hemoglobin solutions—not just red blood cell substi-
tutes. Crit Care Medi 28: 3025–34, 2000. doi:10.1097/00003246-200008000-00058
[16] Davis MJ. Perspective: Physiological role(s) of the vascular myogenic response. Microcircu­
lation 19: 99-114, 2012. doi:10.1111/j.1549-8719.2011.00131.x
[17] Davis MJ, Hill MA, Kuo L. Local regulation of microvascular perfusion. In: Handbook
of physiology: microcirculation, 2nd ed., eds. Tuma RF, Duran WN, Ley K. San Diego, CA:
Elsevier, pp 161–284, 2008. doi:10.1016/B978-0-12-374530-9.00006-1
[18] Dickerson RE, Geis I. Hemoglobin. Menlo Park, CA: Benjamin/Cummings Publishing
Company, 1983.
[19] Duling BR. Changes in microvascular diameter and oxygen tension induced by carbon
dioxide. Circ Res 32: 370–76, 1973. doi:10.1161/01.RES.32.3.370
[20] Duling BR. Oxygen sensitivity of vascular smooth muscle. II. In vivo studies. Am J Physiol
227: 42–49, 1974.
[21] Duling BR. Oxygen, metabolism and microcirculatory control. In: Microcirculation Vol. II,
ed. Kaley G, Altura BM. Baltimore: University Park Press, pp 401–29, 1978.
[22] Duling BR, Berne RM. Longitudinal gradients in periarteriolar oxygen tension. A possible
mechanism for the participation of oxygen in local regulation of blood flow. Circ Res 27:
669–78, 1970. doi:10.1161/01.RES.27.5.669
[23] Duling BR, Kuschinsky W, Wahl M. Measurements of the perivascular PO2 in the vicinity
of the pial vessels of the cat. Pflugers Arch 383: 29–34, 1979. doi:10.1007/BF00584471
[24] Dunphy I, Vinogradov SA, Wilson DF. Oxyphor R2 and G2: phosphors for measuring
oxygen by oxygen-dependent quenching of phosphorescence. Anal Biochem 310: 191–98,
2002. doi:10.1016/S0003-2697(02)00384-6
[25] Ellis CG, Ellsworth ML, Pittman RN. Determination of red blood cell oxygenation in vivo
by dual video densitometric image analysis. Am J Physiol 258: H1216–23, 1990.
[26] Ellis CG, Goldman D, Hanson M, Stephenson AH, Milkovich S, Benlamri A, Ellsworth
ML, Sprague RS. Defects in oxygen supply to skeletal muscle of prediabetic ZDF rats. Am
J Physiol Heart Circ Physiol 298: H1661–70, 2010. doi:10.1152/ajpheart.01239.2009
[27] Ellis CG, Potter RF, Groom AC. The Krogh cylinder geometry is not appropriate for mod-
eling O2 transport in contracted skeletal muscle. Adv Exp Med Biol 159: 253–68, 1983.
[28] Ellsworth ML, Ellis CG, Goldman D, Stephenson AH, Dietrich HH, Sprague RS. Eryth-
rocytes: oxygen sensors and modulators of vascular tone. Physiology (Bethesda) 24: 107–16,
2009. doi:10.1152/physiol.00038.2008
[29] Ellsworth ML, Ellis CG, Popel AS, Pittman RN. Role of microvessels in oxygen supply to
tissue. News Physiol Sci 9: 119–23, 1994.
References  91

[30] Ellsworth ML, Pittman RN. Evaluation of photometric methods for quantifying convec-
tive mass transport in microvessels. Am J Physiol 251: H869–79, 1986.
[31] Ellsworth ML, Pittman RN. Arterioles supply oxygen to capillaries by diffusion as well as
by convection. Am J Physiol 258: H1240–43, 1990.
[32] Ellsworth ML, Popel AS, Pittman RN. Assessment and impact of heterogeneities of con-
vective oxygen transport parameters in capillaries of striated muscle: experimental and theo-
retical. Microvasc Res 35: 341–62, 1988. doi:10.1016/0026-2862(88)90089-1
[33] Federspiel WJ, Popel AS. A theoretical analysis of the effect of the particulate nature of
blood on oxygen release in capillaries. Microvasc Res 32: 164–89, 1986. doi:10.1016/0026-
2862(86)90052-X
[34] Garby L, Meldon J. The respiratory functions of blood. New York, NY: Plenum, 1977.
doi:10.1007/978-1-4684-2313-6
[35] Gladwin MT, Raat NH, Shiva S, Dezfulian C, Hogg N, Kim-Shapiro DB, Patel RP. Ni-
trite as a vascular endocrine nitric oxide reservoir that contributes to hypoxic signaling,
cytoprotection, and vasodilation. Am J Physiol Heart Circ Physiol 291: H2026-H2035, 2006.
doi:10.1152/ajpheart.00407.2006
[36] Golub AS, Pittman RN. Erythrocyte-associated transients in PO2 revealed in capillar-
ies of rat mesentery. Am J Physiol Heart Circ Physiol 288: H2735–43, 2005. doi:10.1152/
ajpheart.00711.2004
[37] Golub AS, Pittman RN. PO2 measurements in the microcirculation using phosphores-
cence quenching microscopy at high magnification. Am J Physiol Heart Circ Physiol 294:
H2905–16, 2008. doi:10.1152/ajpheart.01347.2007
[38] Golub AS, Pittman RN. Oxygen dependence of respiration in rat spinotrapezius muscle
in situ. Am J Physiol Heart Circ Physiol 303: H47-H56, 2012. doi:10.1152/ajpheart.00131
.2012
[39] Golub AS, Pittman RN. Bang-bang model for regulation of local blood flow. Microcircula­
tion 20: 455–483, 2013. doi:10.1111/micc.12051
[40] Golub AS, Pittman RN. A paradigm shift for local blood flow regulation. J Appl Physiol
116: 703–705, 2014. doi:10.1152/japplphysiol.00964.2013
[41] Golub AS, Pittman RN. Letter to the Editor – Last Word on Viewpoint: A paradigm shift
for local blood flow regulation. J Appl Physiol 116: 708, 2014.
[42] Golub AS, Barker MC, Pittman RN. PO2 profiles near arterioles and tissue oxygen
consumption in rat mesentery. Am J Physiol Heart Circ Physiol 293: H1097–1106, 2007.
doi:10.1152/ajpheart.00077.2007
[43] Golub AS, Popel AS, Zheng L, Pittman RN. Analysis of phosphorescence in heteroge-
neous systems using distributions of quencher concentration. Biophys J 73: 452–65, 1997.
doi:10.1016/S0006-3495(97)78084-6
92  Regulation of  Tissue Oxygenation

[44] Golub AS, Song BK, Pittman RN. Muscle contraction increases interstitial nitric oxide as
predicted by a new model of local blood flow regulation. J Physiol 592.6: 1225–1235, 2014.
doi:10.1113/jphysiol.2013.267302
[45] Golub AS, Tevald MA, Pittman RN. Phosphorescence quenching micro-respirometry of
skeletal muscle in situ. Am J Physiol Heart Circ Physiol 300: H135–43, 2011.
[46] Goodnough LT, Scott MG, Monk TG. Oxygen carriers as blood substitutes: Past, present,
and future. Clin Orthopaed Relat Res 357: 89–100, 1998. doi:10.1097/00003086-199812000
-00013
[47] Granger HJ, Shepherd AP Jr. Intrinsic microvascular control of tissue oxygen delivery. Mi­
crovasc Res 5: 49–72, 1973. doi:10.1016/S0026-2862(73)80006-8
[48] Gutierrez G. The rate of oxygen release and its effect on capillary O2 tension: a mathemati-
cal analysis. Respir Physiol 63: 79–96, 1986. doi:10.1016/0034-5687(86)90032-0
[49] Guyton AC, Jones CE, Coleman TG. Circulatory physiology: cardiac output and its regula­
tion, 2nd ed. Philadelphia: Saunders, 1973. doi:10.1097/00000441-196501000-00051
[50] Hellums JD. The resistance to oxygen transport in the capillaries relative to that in the sur-
rounding tissue. Microvasc Res 13: 131–36, 1977. doi:10.1016/0026-2862(77)90122-4
[51] Hellums JD, Nair PK, Huang NS, Ohshima N. Simulation of intraluminal gas transport
processes in the microcirculation. Ann Biomed Eng 24: 1–24, 1996. doi:10.1007/BF027
70991
[52] Hill AV. The diffusion of oxygen and lactic acid through tissues. Proc Roy Soc Ser B 104:
39–96, 1928. doi:10.1098/rspb.1928.0064
[53] Hogan MC, Bebout DE, Wagner PD. Effect of hemoglobin concentration on maximal O2
uptake in canine gastrocnemius muscle in situ. J Appl Physiol 70: 1105–12, 1991.
[54] Honig CR, Gayeski TE. Comparison of intracellular PO2 and conditions for blood-tissue
O2 transport in heart and working red skeletal muscle. Adv Exp Med Biol  215: 309–21, 1987.
doi:10.1007/978-1-4684-7433-6_36
[55] Ignarro LJ. Nitric oxide: biology and pathobiology. San Diego, CA: Academic Press, 2000.
doi:10.1016/B978-0-12-801238-3.00245-2
[56] Ivanov KP, Derry AN, Vovenko EP, Samoilov MO, Semionov DG. Direct measurements
of oxygen tension at the surface of arterioles, capillaries and venules of the cerebral cortex.
Pflugers Arch 393: 118–20, 1982. doi:10.1007/BF00582403
[57] Jacquez JA. Respiratory physiology. New York: McGraw-Hill, 1979.
[58] Japee SA, Pittman RN, Ellis CG. A new video image analysis system to study red blood
cell dynamics and oxygenation in capillary networks. Microcirculation 12: 489–506, 2005.
doi:10.1080/10739680591003332
[59] Japee SA, Pittman RN, Ellis CG. Automated method for tracking individual red blood cells
References  93

within capillaries to compute velocity and oxygen saturation. Microcirculation 12: 507–15,
2005. doi:10.1080/10739680591003341
[60] Jagger JE, Bateman RM, Ellsworth ML, Ellis CG. Role of erythrocyte in regulating lo-
cal O2 delivery mediated by hemoglobin oxygenation. Am J Physiol Heart Circ Physiol 280:
H2833-H2839, 2001.
[61] Jespersen SN, Ostergaard L. The roles of cerebral blood flow, capillary transit time hetero-
geneity, and oxygen tension in brain oxygenation and metabolism. J Cereb Blood Flow Metab
32: 264–277, 2012. doi:10.1038/jcbfm.2011.153
[62] Joyner MJ, Wilkins BW. Exercise hyperaemia: is anything obligatory but the hyperaemia?
J Physiol 583: 855–860, 2007. doi:10.1113/jphysiol.2007.135889
[63] Ketcham EM, Cairns CB. Hemoglobin-based oxygen carriers: Development and clinical
potential. Ann Emerg Med 33: 326–37, 1999. doi:10.1016/S0196-0644(99)70370-7
[64] Kindig CA, Richardson TE, Poole DC. Skeletal muscle capillary hemodynamics from
rest to contractions: implications for oxygen transfer. J Appl Physiol 92: 2513–20, 2002.
doi:10.1152/japplphysiol.01222.2001
[65] Klitzman B, Damon DN, Gorczynski RJ, Duling BR. Augmented tissue oxygen sup-
ply during striated muscle contraction in the hamster. Relative contributions of cap-
illary recruitment, functional dilation, and reduced tissue PO2. Circ Res 51: 711–21,
1982. doi:10.1161/01.RES.51.6.711
[66] Kobayashi H, Takizawa N. Imaging of oxygen transfer among microvessels of rat cremaster
muscle. Circulation 105: 1713–19, 2002. doi:10.1161/01.CIR.0000013783.63773.8F
[67] Koeppen BM, Stanton BA. Berne & Levy physiology, 6th ed. Philadelphia: Mosby Else­
vier, 2008.
[68] Krogh A. The rate of diffusion of gases through animal tissues with some remarks on the
coefficient of invasion. J Physiol 52: 391–408, 1919. doi:10.1113/jphysiol.1919.sp001838
[69] Krogh A. The number and distribution of capillaries in muscles with calculations of
the oxygen pressure head necessary for supplying the tissue. J Physiol 52: 409–15, 1919.
doi:10.1113/jphysiol.1919.sp001839
[70] Krogh A. The anatomy and physiology of capillaries. New York: Hafner Publishing, 1959.
[71] Kuo L, Pittman RN. Effect of hemodilution on oxygen transport in arteriolar networks of
hamster striated muscle. Am J Physiol 254: H331–39, 1988.
[72] Kuo L, Pittman RN. Influence of hemoconcentration on arteriolar oxygen transport in
hamster striated muscle. Am J Physiol 259: H1694–1702, 1990.
[73] Lamb IR, Murrant CL. Potassium inhibits nitric oxide and adenosine arteriolar vasodi-
latation via KIR and Na+/K+ ATPase: implications for redundancy in active hyperaemia.
J Physiol 593.23: 5111-5126, 2015.
94  Regulation of  Tissue Oxygenation

[74] Lash JM, Bohlen HG. Perivascular and tissue PO2 in contracting rat spinotrapezius muscle.
Am J Physiol 252: H1192–1202, 1987.
[75] Lash JM, Bohlen HG. Excess oxygen delivery during muscle contractions in spontaneously
hypertensive rats. J Appl Physiol 78: 101–11, 1995.
[76] Lih MM-S. Transport phenomena in medicine and biology. New York, NY: John Wiley &
Sons, 1975.
[77] Lipowsky HH, Usami S, Chien S, Pittman RN. Hematocrit determination in small bore
tubes by differential spectrophotometry. Microvasc Res 24: 42–55, 1982. doi:10.1016/0026-
2862(82)90041-3
[78] Loeschke HH, Gertz KH. Einfluss des O2-Druckes in der Einatmungsluft auf die Atem-
tatigkeit des Menschen, gepruft unter Konstanthaltung des alveolaren CO2-Druckes. Arch
Ges Physiol 267: 460–477, 1958.
[79] Marconi C, Heisler N, Meyer M, Weitz H, Pendergast DR, Cerretelli P, Piiper J. Blood
flow distribution and its temporal variability in stimulated dog gastrocnemius muscle. Respir
Physiol 74: 1–13, 1988. doi:10.1016/0034-5687(88)90135-1
[80] Michel CC. The transport of oxygen and carbon dioxide by the blood. In Respiratory
physiology, ed Guyton AC, Widdicombe JG, Baltimore: University Park Press, pp 67–104,
1974.
[81] Middleman S. Transport phenomena in the cardiovascular system. New York, NY: John
Wiley & Sons, 1972.
[82] Molé PA, Chung Y, Tran TK, Sailasuta N, Hurd R, Jue T. Myoglobin desaturation with
exercise intensity in human gastrocnemius muscle. Am J Physiol Reg Integ Comp Physiol
277: R173–80, 1999.
[83] Murrant CL, Sarelius IH. Local control of  blood flow during active hyperaemia: what kinds
of integration are important? J Physiol 593.21: 4699-4711, 2015. doi:10.1113/JP270205
[84] Nucci ML, Abuchowski A. The search for blood substitutes. Sci Am 278: 72–77, 1998.
doi:10.1038/scientificamerican0298-72
[85] Nugent WH, Song BK, Pittman RN, Golub AS. Simultaneous sampling of tissue ox-
ygenation and oxygen consumption in skeletal muscle. Microvasc Res 105: 15–22, 2016.
doi:10.1016/j.mvr.2015.12.007
[86] Piiper J, Scheid P. Diffusion limitation of O2 supply to tissue in homogeneous and hetero-
geneous models. Respir Physiol 85: 127–36, 1991. doi:10.1016/0034-5687(91)90011-7
[87] Pittman RN. In vivo photometric analysis of hemoglobin. Ann Biomed Eng 14: 119–37,
1986. doi:10.1007/BF02584263
[88] Pittman RN. Microvessel blood oxygen measurement techniques. In Microvascular technol­
ogy, eds. CH Baker and WL Nastuk. Orlando, FL: Academic Press, 367–89, 1986.
References  95

[89] Pittman RN. Influence of microvascular architecture on oxygen exchange in skeletal muscle.
Microcirculation 2: 1–18, 1995. doi:10.3109/10739689509146755
[90] Pittman RN. Oxygen transport and exchange in the microcirculation. Microcirculation 12:
59–70, 2005. doi:10.1080/10739680590895064
[91] Pittman, RN. Oxygen gradients in the microcirculation. Acta Physiologica (Oxf ) 2011 doi:
10.1111/j.1748-1716.2010.02232.x.
[92] Pittman RN. Oxygen transport in the microcirculation and its regulation. Microcirculation
20: 117–137, 2013. doi:10.1111/micc.12017
[93] Pittman RN, Duling BR. Oxygen sensitivity of vascular smooth muscle. I. In vitro studies.
Microvasc Res 6: 202–11, 1973. doi:10.1016/0026-2862(73)90020-4
[94] Pittman RN, Duling BR. A new method for the measurement of percent oxyhemoglobin.
J Appl Physiol 38: 315–20, 1975.
[95] Pittman RN, Duling BR. Measurement of percent oxyhemoglobin in the microvasculature.
J Appl Physiol 38: 321–27, 1975.
[96] Pittman RN, Golub AS, Popel AS, Zheng L. Interpretation of phosphorescence quenching
measurements made in the presence of oxygen gradients. Adv Exp Med Biol 454: 375–83, 1998.
doi:10.1007/978-1-4615-4863-8_45
[97] Poole DC, Brown MD, Hudlicka O. Counterpoint: there is not capillary recruitment in ac-
tive skeletal muscle during exercise. J Appl Physiol 104: 891-893, 2008. Discussion 893–894.
doi:10.1152/japplphysiol.00779.2007a
[98] Poole DC, Copp SW, Hirai DM, Musch TI. Dynamics of muscle microcirculatory
and blood-myocyte O2 flux during contractions. Acta Physiol (Oxf ) 2010 doi: 10.1111/
j.1748-1716.2010.02246.x.
[99] Popel AS. Theory of oxygen transport to tissue. Crit Rev Biomed Eng 17: 257–321, 1989.
[100] Popel AS, Goldman D, Vadapalli A. Modeling of oxygen diffusion from the blood vessels
to intracellular organelles. Adv Exp Med Biol 530: 485–95, 2003. doi:10.1007/978-1-4615-
0075-9_46
[101] Popel AS, Pittman RN. The microcirculation physiome. In: Handbook of biomedical engi­
neering, 4th ed., ed. by JD Bronzino and DR Peterson, Boca Raton, FL: Taylor & Francis/
CRC Press, Chapter 20:20.1-20.17, 2015.
[102] Popel AS, Pittman RN, Ellsworth ML. Rate of oxygen loss from arterioles is an order of
magnitude higher than expected. Am J Physiol 256: H921–24, 1989. doi:10.1016/0883-
9441(89)90050-6
[103] Richardson RS, Noyszewski EA, Kendrick KF, Leigh JS, Wagner PD. Myoglobin O2 de-
saturation during exercise. Evidence of limited O2 transport. J Clin Invest 96: 1916–26,
1995. doi:10.1172/JCI118237
96  Regulation of  Tissue Oxygenation

[104] Riess JG, LeBlanc M. Preparation of perfluorocarbon emulsions for biomedical use: prin-
ciples, materials, and methods. In KC Lowe (ed), Blood substitutes: preparation, physiology,
and medical applications, pp 94–129, Chichester, UK: Ellis Horwood Ltd., 1988.
[105] Riley RL. Gas exchange and transportation. In TC Ruch, HD Patton (ed), Physiology and
biophysics, 19th ed, Philadelphia: Saunders, pp 761–87, 1965.
[106] Roughton FJW. Respiratory functions of blood. In Handbook of respiratory physiology.
pp 55–102, Randolph Field, TX: USAF School of Aviation Medicine, 1954.
[107] Schneiderman G, Goldstick TK. Oxygen electrode design criteria and performance char­
acteristics: recessed cathode. J Appl Physiol 45: 145–54, 1978.
[108] Scott MG, Kucki DF, Goodnough LT, Monk TG. Blood substitutes: Evolution and future
applications. Clin Chem, 43: 1724–31, 1997.
[109] Secomb TW, Hsu R. Simulation of O2 transport in skeletal muscle: diffusive exchange
between arterioles and capillaries. Am J Physiol 267: H1214–21, 1994.
[110] Segal SS. Regulation of blood flow in the microcirculation. Microcirculation 12: 33–45,
2005. doi:10.1080/10739680590895028
[111] Sharan M, Vovenko EP, Vadapalli A, Popel AS, Pittman RN. Experimental and theo-
retical studies of oxygen gradients in rat pial microvessels. J Cereb Blood Flow Metab 28:
1597–1604, 2008. doi:10.1038/jcbfm.2008.51
[112] Shibata M, Ichioka S, Ando J, Kamiya A. Microvascular and interstitial PO2 measurements
in rat skeletal muscle by phosphorescence quenching. J Appl Physiol 91: 321–27, 2001.
[113] Smith LM, Barbee RW, Ward KR, Pittman RN. Prolonged tissue PO2 reduction after
contraction in spinotrapezius muscle of spontaneously hypertensive rats. Am J Physiol Heart
Circ Physiol 287: H401–07, 2004. doi:10.1152/ajpheart.00980.2002
[114] Spahn D, Leone B, Reves J, Pasch T. Cardiovascular and coronary physiology of acute
isovolemic hemodilution: A review of nonoxygen-carrying and oxygen-carrying solutions.
Anesth Analg 78: 1000–021, 1994. doi:10.1213/00000539-199405000-00029
[115] Spahn DR, Pasch T. Physiological properties of  blood substitutes. News Physiolog Sci 16:
38–41, 2001.
[116] Spence RK. Perfluorocarbons. In: EC Rossi, TL Simon, GS Moss, SA Gould (eds), Prin­
ciples of transfusion medicine, pp 189–96, Baltimore: Williams & Wilkins, 1996.
[117] Stamler JS, Jia L, Eu JP, McMahon TJ, Demchenko IT, Bonaventura J, Gernert K, Pianta-
dosi CA. Blood flow regulation by S-nitrosohemoglobin in the physiological oxygen gradi-
ent. Science 276: 2034–2037, 1997. doi:10.1126/science.276.5321.2034
[118] Styp-Rekowska B, Disassa NM, Reglin B, Ulm L, Kuppe H, Secomb TW, Pries AR. An
imaging spectroscopy approach for measurement of oxygen saturation and hematocrit during
intravital microscopy. Microcirculation 14: 207–21, 2007. doi:10.1080/10739680601139302
References  97

[119] Swain DP, Pittman RN. Oxygen exchange in the microcirculation of hamster retractor
muscle. Am J Physiol 256: H247–55, 1989.
[120] Tateishi N, Suzuki Y, Tanaka J, Maeda N. Imaging of oxygen saturation and distribution of
erythrocytes in microvessels. Microcirculation 4: 403–12, 1997. doi:10.3109/107396897091
46804
[121] Torres Filho IP, Intaglietta M. Microvessel PO2 measurements by phosphorescence decay
method. Am J Physiol Heart Circ Physiol 265: H1434–38, 1993.
[122] Torres Filho IP, Terner J, Pittman RN, Somera LG 3rd, Ward KR. Hemoglobin oxygen
saturation measurements using resonance Raman intravital microscopy. Am J Physiol Heart
Circ Physiol 289: H488–95, 2005. doi:10.1152/ajpheart.01171.2004
[123] Torres Filho IP, Terner J, Pittman RN, Proffitt E, Ward KR. Measurement of hemoglobin
oxygen saturation using Raman microspectroscopy and 532-nm excitation. J Appl Physiol
104: 1809–17, 2008. doi:10.1152/japplphysiol.00025.2008
[124] Tsai AG, Cabrales P, Johnson PC, Intaglietta M, Golub AS, Pittman RN. Effect of oxygen
consumption by measuring method on PO2 transients associated with the passage of eryth-
rocytes in capillaries of rat mesentery. Am J Physiol Heart Circ Physiol 289: H1777–79, 2005.
doi:10.1152/ajpheart.00503.2005
[125] Tsai AG, Friesenecker B, Mazzoni MC, Kerger H, Buerk DG, Johnson PC, Intaglietta M.
Microvascular and tissue oxygen gradients in the rat mesentery. Proc Natl Acad Sci USA 95:
6590–95, 1998. doi:10.1073/pnas.95.12.6590
[126] Tsai AG, Johnson PC, Intaglietta M. Oxygen gradients in the microcirculation. Physiol Rev
83: 933–63, 2003. doi:10.1152/physrev.00034.2002
[127] Vadapalli A, Pittman RN, Popel AS. Estimating oxygen transport resistance of the micro-
vascular wall. Am J Physiol Heart Circ Physiol 279: H657–71, 2000.
[128] van Assendelft OW. Spectrophotometry of haemoglobin derivatives. Assen, The Netherlands:
Charles C Thomas, 1970.
[129] Vanderkooi JM, Maniara G, Green TJ, Wilson DF. An optical method for measurement
of dioxygen concentration based upon quenching of phosphorescence. J Biol Chem 262:
5476–82, 1987.
[130] Voter WA, Gayeski TE. Determination of myoglobin saturation of frozen specimens using
a reflecting cryospectrophotometer. Am J Physiol 269: H1328–41, 1995.
[131] Walley KR. Heterogeneity of oxygen delivery impairs oxygen extraction by peripheral tis-
sues: theory. J Appl Physiol 81: 885–94, 1996.
[132] Weibel ER. The pathway for oxygen. Cambridge: Harvard University Press, 1984.
[133] West JB. Respiratory physiology: the essentials, 8th ed., Baltimore: Lippincott Williams &
Wilkins, 2008. doi:10.1016/B978-0-12-801238-3.00214-2
98  Regulation of  Tissue Oxygenation

[134] Whalen WJ, Riley J, Nair P. A microelectrode for measuring intracellular PO2.  J Appl Physiol
23: 798–801, 1967.
[135] Wilson DF, Erecin´ska M, Drown C, Silver IA. The oxygen dependence of cellular energy
metabolism. Arch Biochem Biophys 195: 485–93, 1979. doi:10.1016/0003-9861(79)90375-8
[136] Wilson DF, Harrison DK, Vinogradov, SA. Oxygen, pH, and mitochondrial oxidative
phosphorylation. J Appl Physiol 113: 1838-1845, 2012. doi:10.1152/japplphysiol.01160
.2012
[137] Winslow RM. Blood substitutes. Adv Drug Deliv Rev 40: 131–42, 1999. doi:10.1016/
S0169-409X(99)00045-9
[138] Winslow RM. Blood substitutes. Curr Opin Hematol  9: 146–51, 2002. doi:10.1097/00062
752-200203000-00011
[139] Wittenberg JB. Myoglobin-facilitated oxygen diffusion: role of myoglobin in oxygen entry
into muscle. Physiol Rev 50: 559–636, 1970.
[140] Wittenberg JB, Wittenberg BA. Myoglobin function reassessed. J Exp Biol 206: 2011–20,
2003. doi:10.1242/jeb.00243
99

Author Biography

Roland N. Pittman is a professor of  Physiology and Biophysics at the Medical College of  Virginia
Campus of Virginia Commonwealth University in Richmond, Virginia. He received his S.B. in
Physics from the Massachusetts Institute of   Technology in 1966, his M.A. in Physics from the State
University of New York at Stony Brook in 1968, and his Ph.D. in Physics from the State Univer-
sity of New York at Stony Brook in 1971. He then received post­doctoral training in microcircula-
tion in the laboratory of Dr. Brian R. Duling in the Department of Physiology at the University
of Virginia. Dr. Pittman also holds joint faculty appointments in the Department of Biomedical
Engineering and the Department of Emergency Medicine at Virginia Commonwealth University.
His research has focused on the transport of oxygen and its regulation from the perspective of the
microcirculation, and he has worked with several different colleagues to develop intravital micro-
scopic methods to measure blood and tissue oxygenation. He has published more than 150 articles
on oxygen transport and related topics.

You might also like