Louisiana State University LSU Digital Commons

LSU Doctoral Dissertations Graduate School

2003 Arterial and venous adaptations to short-term handgrip exercise training Mahmoud Awad Alomari Louisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_dissertations Part of the Kinesiology Commons

Recommended Citation Alomari, Mahmoud Awad, "Arterial and venous adaptations to short-term handgrip exercise training" (2003). LSU Doctoral Dissertations. 188. https://digitalcommons.lsu.edu/gradschool_dissertations/188

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please [email protected]. ARTERIAL AND VENOUS ADAPTATIONS TO SHORT-TERM HANDGRIP EXERCISE TRAINING

A Dissertation Submitted to the Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College in partial fulfillment of the requirements for the degree of Doctor of Philosophy in The Department of Kinesiology

By Mahmoud Alomari B.S., Yarmouk University, Irbid, Jordan, 1990 M.S. Minnesota State University, Mankato, MN, 1995 December, 2003

© Copyright 2003 Mahmoud A. Alomari All right reserved

ii DEDICATION

I dedicate all of my work to my parents, the love of my life. They feel as though they took every exam with me and were as anxious as I was for each defense. Their confidence in me never wavered and helped me to accomplish the dream of my life. Their motivation made me a better person and they continue to show me what service to others really is. Abuwael and Umwael thank you for the endless and exceptional support, sacrifices, inspiration and love.

iii ACKNOWLEDGMENTS

All gratitude goes to “Allah” (God) SW ….. without him my dissertation would not make

it to reality. I must express my deep gratefulness to my committee members for their help and guidance particularly Dr. Wood for his constant positive and encouraging remarks.

Rhonda and Rafael (good luck ……. your next!!!)…… we have always toned on the

same brainwave …….. I really appreciate the bright suggestions, creative ideas, and

fruitful discussions. I also wish to thank Muaz and Nicki for being instrumental in this

work ….. and the rest of my colleagues Arturo, Rania, Andrea, Zeki, and JoeyLynn I will

always remember you. Deep thanks also goes to all my friends and family around the

world for their special encouragement and support.

His teaching, guidance, scholarship, and service to students as well as to the

academic world are models of professional and personal investment. Despite the bad,

ugly, ugly and ugly we have had many many goods…….. I feel extremely lucky to work

under the supervision of Dr. Welsch for the past 6 years ….. Mike has made a change in

my life not just in academia but also has crafted a different person out of me. Words

certainly fail to completely express my appreciation …………….. but I sincerely would

like to say THANK YOU Mike.

iv TABLE OF CONTENTS

DEDICATION...... iii

ACKNOWLEDGMENTS ...... iv

LIST OF TABLES...... vii

LIST OF FIGURES ...... viii

ABSTRACT...... x

CHAPTER 1. INTRODUCTION ...... 1 1.1. REFRENCES ------4

CHAPTER 2: MEASUREMENTS OF VASCULAR FUNCTION USING STRAIN- GAUGE PLETHYSMOGRAPHY: TECHNICAL CONSIDERATIONS, STANDARDIZATION AND PHYSIOLOGICAL FINDINGS ...... 7 2.1. INTRODUCTION------8 2.2. METHODOLOGY------8 2.3. RESULTS ------13 2.4. DISCUSSION------18 2.5. CONCLUSION------26 2.6. REFERENCES ------27

CHAPTER 3. INFLUENCE OF VENOUS FUNCTION ON EXERCISE TOLERANCE IN CHRONIC FAILURE ...... 31 3.1. INTRODUCTION------32 3.2. METHODS------33 3.3. RESULTS ------35 3.4. DISCUSSION------38 3.5. CONCLUSION------42 3.6. REFERENCES ------42

CHAPTER 4. MODIFICATION OF FOREARM VASCULAR FUNCTION FOLLOWING SHORT-TERM HANDGRIP EXERCISE TRAINING ...... 46 4.1. INTRODUCTION------47 4.2. METHODS------48 4.3. RESULTS ------51 4.4. DISCUSSION------54 4.5. REFERENCES ------59

CHAPTER 5. ARTERIAL AND VENOUS ADAPTATIONS TO SHORT-TERM HANDGRIP EXERCISE TRAINING ...... 61 5.1. INTRODUCTION------62 5.2. HYPOTHESES ------63

v 5.3. METHODS------63 5.4. RESULTS ------68 5.5. DISCUSSION------76 5.6. CONCLUSION------86 5.7. REFERENCES ------86

CHAPTER 6. SUMMARY...... 93 6.1. REFERENCES ------95

APPENDIX A: LITERATURE REVIEW...... 97

APPENDIX B: PROPOSAL ...... 280

APPENDIX C: LETTERS OF PERMISSIONS...... 310

VITA...... 315

vi LIST OF TABLES Table 2.1. Subject characteristics...... 13

Table 2.2. Resting vascular indices following varying occlusion pressures...... 14

Table 2.3. Between-days ICCC for forearm arterial and venous indices ...... 17

Table 2.4. Pearson product moment correlation between arterial, venous and physical fitness measures ...... 20

Table 2.5. Differences between the cardiorespiratory groups in forearm arterial and venous indices...... 21

Table 2.6. Differences between the handgrip strength groups in forearm arterial and venous indices...... 21

Table 3.1. Participant characteristics ...... 36

Table 3.2. Independent t-test comparisons between CHF and control groups...... 37

Table 4.1. Subject characteristics...... 52

Table 4.2. Dominant and non-dominant FVR changes in the LO and HI groups at rest, following OCC and following ROCC...... 55

Table 4.3. Literature overview...... 56

Table 5.1. Subject characteristics...... 68

Table 5.2. Cardiovascular hemodynamics and HRV before and after 4-week handgrip exercise training...... 68

Table 5.3. Arm ergometer exercise test, forearm skinfolds, and sum of three skinfolds before and after 4-week handgrip exercise training...... 69

Table 5.4. Resting arterial and venous indices ...... 70

Table 5.5. Post occlusion arterial and venous indices ...... 70

Table 5.6. Pearson product moment correlation between arterial, venous and physical fitness measures ...... 75

vii LIST OF FIGURES Figure 2.1a. Venous capacitance using different pressures...... 15

Figure 2.1b. Venous outflow using different pressures...... 15

Figure 2.2. Venous capacitance and outflow following 5 & 10 minutes of venous occlusion...... 16

Figure 2.3. Venous capacitance over a 10 minute venous occlusion period...... 16

Figure 2.4a. Relationship between estimated VO2 & post arterial occlusion blood inflow...... 17

Figure 2.4b. Relationship between estimated VO2 & resting venous outflow...... 18

Figure 2.5a. Relationship between handgrip strength & resting venous capacitance...... 19

Figure 2.5b. Relationship between handgrip strength & resting venous outflow...... 19

Figure 3.1. Relationship between forearm venous capacitance and maximum walking distance...... 38

Figure 3.2. Relationship between forearm venous outflow and maximum walking distance...... 39

Figure 4.1. FBF changes in the LO group following HG exercise training...... 53

Figure 4.2. FBF changes in the HI group following HG exercise training...... 54

Figure 5.1. Changes in peak forearm blood inflow...... 71

Figure 5.2. Changes in peak forearm vascular resistance...... 71

Figure 5.3. Changes in resting forearm venous compliance...... 72

Figure 5.4. Changes in post occlusion forearm venous capacitance...... 73

Figure 5.5. Changes in post occlusion forearm venous outflow...... 74

Figure 5.6. Relationship between handgrip strength & resting venous capacitance ...... 74

Figure 5.7. Relationship between handgrip strength & resting venous outflow...... 83

Figure 5.8. Relationship between post occlusion blood inflow & venous compliance .... 84

Figure 5.9. Relationship between post occlusion blood inflow & venous outflow ...... 85

viii

Figure 5.10. Relationship between changes in peak blood inflow & resting venous compliance ...... 85

ix ABSTRACT

Four studies on vascular and exercise physiology are presented in this document. The 1st study examined the relationships between measures of fitness and FVF in 55 young [22.6

± 3.5 years] adults. Estimated VO2peak correlated with arterial inflow (Ainf) [r=0.54; p=0.012] and resting venous outflow (Vout) [r=0.56; p=0.016]. Lastly, HG strength was associated with Vcap [r=0.57; p=0.007] and Vout [r=0.67; p=0.001].

The 2nd study examined the relationship between FVF and exercise tolerance

(ExT) in 20 patients with HF [age: 59 ± 13 years] and 10 age-matched controls [age: 51 ±

16 years. The ExT was measured as the maximum walking distance (MWD) in 6 minutes. FVF [Ainf: HF 15.3 ± 6; controls 22 ±6.7; Vcap: HF 1.4 ± 0.5; controls 2.0 ±

0.4; Vout: HF 24.5 ± 9.4; controls 33 ± 10 mL • 100 mL tissue-1 • min –1; and forearm vascular resistance: HF 7.8 ± 3; controls 4.6 ± 1.4U] indices and MWD [HF: 178 ± 65 m; controls: 562 ± 136m, P = .0001] were different between groups. Correlation analysis revealed significant associations between FVF indices and MWD.

The 3rd study examined the effect of 25% (LO) and 75% (HI) of MVC short-term

HG exercise training on FVF in 28 healthy men [Age:23+4.3]. The 4-week program consisted of non-dominant HG exercise performed 5 d/wk for 20-min. Training resulted in increased Ainf in the non-dominant arm in the LO and HI groups by 16.51% and

20.72%, respectively.

The final study examined the time-course FVF adaptations to HG exercise training in 17 men [Age: 22.6 ± 3.5]. The HG exercise was performed in the non- dominant arm 5 d/wk for 20-min at 60% of MVC. The 2 X 5 ANOVA revealed arms X visits interaction for Ainf [p=0.02], while the LSD post-hoc demonstrated unilateral

x increase in Ainf following the 1st week. Additional 2 X 5 split-plot ANOVA tests revealed arms X visits interaction [p=0.04] for venous compliance (Vcomp) with LSD post-hoc demonstrating a decrease in trained arm Vcomp in visit 2 followed by an increase in visit 4 and return to baseline level at visit 5.

xi

CHAPTER 1. INTRODUCTION

The original understanding of vascular physiology has evolved from a passive conduit,

transporting blood to and from body tissues, to a multifaceted system that is interlinked

with many other body systems, regulated by a variety of mechanisms, and adjusts to

different conditions. Our daily poor lifestyle choices, such as high fat meals, cigarette

smoking, inactivity, and work-related stress exposes the vascular system to numerous

assaults. These assaults result in a variety of deteriorating diseases including

atherosclerosis, coronary diseases and hypertension. Many studies have

demonstrated that exercise training can improve sympathetic vascular tone, insulin

sensitivity, and lipid and haemostatic profiles as well as lower blood pressure, obesity,

and stress hormones. Therefore, hand-in-hand with dietary modifications, exercise training appears to be an effective tool to minimize the influence our daily choices and

may results in numerous health benefits.

The current document presents four research projects, designed to examine the acute

and chronic effects of exercise training on vascular function and control in health and

disease. These research projects are introduced in the subsequent four chapters and aimed

at:

1. Addressing technical concerns and determining the reproducibility of vascular

outcomes using strain gauge plethysmography,

2. Examining the ability of vascular measures to differentiate between various

populations, AND

3. Evaluation of the influence of interventions on changing measures of vascular

function.

1 Chapter 2 addresses methodological considerations for mercury strain-gauge

plethysmography (MSGP) and the reliability of measuring arterial and venous function

indices as well as examined the relationships between fitness and vascular indices.

Standard recommendations for venous occlusion pressure and duration have been 50

mmHg and 2 minutes, respectively (7, 11). Others have used, venous pressure “just

below” diastolic blood pressure (22). However, the rationale for these recommendations

are not clear. Additionally, few studies have examined the influence of physical fitness

(cardiovascular and/or muscular) on indices of venous function. The majority of these studies report improvements in venous hemodynamics secondary to changes in blood

volume following endurance training (6, 13, 15). More recently, Wecht et al. attributed

venous modifications, in part, to enhanced vasomotor tone, vessel compliance and vessel

responses to mechanical and neural stimulation (22). Therefore, the major aim of this

study was to examine relationships between measures of physical fitness and arterial and

venous function. It was hypothesized that individuals with the highest levels of fitness

would have greater arterial inflow and venous outflow patterns. However, considering the

lack of standardization and reliability of the protocols used to examine the venous

function, a secondary aim of this study was to examine the effects of different venous

occlusion times and pressures on venous capacitance and outflow as well as the reliability

of MSGP.

Chapter 3 focuses on the effect of vascular function on exercise tolerance in heart

failure (HF). Chronic HF is a complex clinical syndrome characterized by an abnormality

of the heart and by a recognizable pattern of hemodynamic, renal, neural and hormonal

responses (17). Despite of the research describing the effect of HF on blood inflow and

2 the subsequent treatments aimed at improving arterial function, many patients continue to

experience activity-related symptoms (4), possibly suggesting the involvement of other

factors. Interestingly, the role of the venous system on exercise tolerance in patients with

HF has not received much attention, despite evidence of changes in venous structure and

function in patients with HF (12, 18) and in patients with a variety of vascular diseases

(16, 20). Accordingly, the purpose of this study was to examine the relationships between

indices of arterial and venous function and exercise tolerance in chronic HF patients and

age-matched controls. Given the complexity of the HF syndrome, it was hypothesized a

direct relationship between indices of vascular function and exercise tolerance.

Chapter 4 examines the effect of low and high intensity localized exercise training on

arterial function. Exercise training induces arterial adaptations characterized by increased

working muscle perfusion and reduced vascular resistance. The precise mechanisms for

these adaptations are not known however they appear to be exercise training modality

dependent (3, 8, 9). Interestingly, vascular adaptations following small muscle group

exercise training programs using different intensities have not been adequately explored.

A study by Yasuda et al. (23) was the only study where subjects performed 6-weeks of

HG exercise training at intensities of 33 or 50% of maximum voluntary contraction

(MVC). Peak forearm blood flow increased similarly in both groups. However, the

difference between training intensities was only 17% and may have been insufficient to

determine whether the adaptations are indeed intensity-dependent. Accordingly, the

purpose of this study was to determine the effects of low and high intensity short-term

HG exercise training programs on localized vascular function in healthy young men. It

3 was hypothesized that a training exercise intensity of 75% of MVC would promote

greater vasodilatory capacity compared to a training exercise intensity of 25% of MVC.

Finally in chapter 5, the time course changes to 4 week HG exercise training program

on arterial and venous function indices were examined. Structural and functional changes

in the arterial system following exercise training are well documented (2, 5). Recent

studies report training-induced adaptations may be observed following one week of training in animal models (14, 19), and in human single large conduit (i.e.

brachial) (1). Interestingly, the effect of exercise training on venous function has received

considerably less attention despite compelling evidence demonstrating its dynamic role

beyond a typical role as a “passive volume reservoir” (10, 21). Therefore, the purpose of

this study was to examine the influence of 4-weeks of handgrip exercise training on

arterial and venous function. A secondary aim of this study was to determine the time

course of change for both arterial and venous indices associated with the handgrip

exercise training program. It has been hypothesized that handgrip exercise training will

result in changes in arterial and venous function indices following the 1st week.

Appendix A includes an extensive literature review of vascular function and control

at rest and during exercise in health and disease. Appendix B is the proposal of the final

study followed by the author’s vita.

1.1. REFRENCES

1.Allen JD, Geaghan JP, Greenway F and Welsch MA. Time Course of Improved Flow- Mediated Dilation after Short-Term Exercise Training. Med Sci Sports Exerc 35: 847- 853, 2003.

2.Alomari MA, Welsch MA, Prisby RD, Lee CM and Wood RH. Modification of forearm vascular function following short-term handgrip exercise training. Int J Sports Med 22: 361-365., 2001.

4 3.Bond V, Jr., Wang P, Adams RG, Johnson AT, Vaccaro P, Tearney RJ, Millis RM, Franks BD and Bassett DR, Jr. Lower leg high-intensity resistance training and peripheral hemodynamic adaptations. Can J Appl Physiol 21: 209-217, 1996.

4.Clark AL, Poole-Wilson PA and Coats AJ. Exercise limitation in chronic heart failure: central role of the periphery. J Am Coll Cardiol 28: 1092-1102, 1996.

5.Clarkson P, Montgomery HE, Mullen MJ, Donald AE, Powe AJ, Bull T, Jubb M, World M and Deanfield JE. Exercise training enhances endothelial function in young men. J Am Coll Cardiol 33: 1379-1385, 1999.

6.Convertino VA, Mack GW and Nadel ER. Elevated central venous pressure: a consequence of exercise training-induced hypervolemia? Am J Physiol 260: R273-277, 1991.

7.Cramer M, Langlois Y, Beach K, Martin D and Strandness Jr DE. Standardization of venous flow measurements by strain gauge plethysmography in normal subjects. Bruit VII: 33-39, 1983.

8.Delp MD and Laughlin MH. Time course of enhanced -mediated dilation in of trained rats. Med Sci Sports Exerc 29: 1454-1461, 1997.

9.Green DJ, Cable NT, Fox C, Rankin JM and Taylor RR. Modification of forearm resistance vessels by exercise training in young men. J Appl Physiol 77: 1829-1833, 1994.

10.Hester RL and Hammer LW. Venular-arteriolar communication in the regulation of blood flow. Am J Physiol Regul Integr Comp Physiol 282: R1280-1285, 2002.

11.Hokanson DE, Sumner DS and Strandness Jr. DE. An electrically calibrated plethysmograph for direct measurement of limb blood flow. IEEE transactions on bio- medical engineering 22: 21-25, 1975.

12.Ikenouchi H, Iizuka M, Sato H, Momomura S, Serizawa T and Sugimoto T. Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Jpn Heart J 32: 17-34, 1991.

13.Kenney WL and Armstrong CG. The effect of aerobic conditioning on venous pooling in the foot. Med Sci Sports Exerc 19: 474-479, 1987.

14.Laughlin MH. Endothelium-mediated control of coronary vascular tone after chronic exercise training. Med Sci Sports Exerc 27: 1135-1144, 1995.

15.Louisy F, Jouanin JC and Guezennec CY. Filling and emptying characteristics of lower limb venous network in athletes. Study by postural plethysmography. Int J Sports Med 18: 26-29, 1997.

5

16.Monos E, Berczi V and Nadasy G. Local control of : biomechanical, metabolic, and humoral aspects. Physiol Rev 75: 611-666, 1995.

17.Poole-Wilson PA, Buller NP and Lipkin DP. Regional blood flow, muscle strength and skeletal muscle histology in severe congestive heart failure. Am J Cardiol 62: 49E- 52E, 1988.

18.Sato H, Ikenouchi H, Aoyagi T, Matsui H, Mochizuki T, Momomura S, Serizawa T, Iizuka M and Sugimoto T. Pathophysiology and evaluation of severity of congestive heart failure on the basis of venous characteristics. Jpn Circ J 53: 141-145, 1989.

19.Shen W, Zhang X, Zhao G, Wolin MS, Sessa W and Hintze TH. Nitric oxide production and NO synthase gene expression contribute to vascular regulation during exercise. Med Sci Sports Exerc 27: 1125-1134, 1995.

20.Tretbar LL. Chronic venous insufficiency of the legs: pathogenesis of venous ulcers. J Enterostomal Ther 14: 105-108, 1987.

21.Tschakovsky ME, Shoemaker JK and Hughson RL. Vasodilation and muscle pump contribution to immediate exercise hyperemia. Am J Physiol 271: H1697-1701, 1996.

22.Wecht JM, de Meersman RE, Weir JP, Bauman WA and Grimm DR. Effects of autonomic disruption and inactivity on venous vascular function. Am J Physiol Heart Circ Physiol 278: H515-520, 2000.

23.Yasuda Y and Miyamura M. Cross transfer effects of muscular training on blood flow in the ipsilateral and contralateral forearms. Eur J Appl Physiol Occup Physiol 51: 321- 329, 1983.

6 CHAPTER 2: MEASUREMENTS OF VASCULAR FUNCTION USING STRAIN- GAUGE PLETHYSMOGRAPHY: TECHNICAL CONSIDERATIONS, STANDARDIZATION AND PHYSIOLOGICAL FINDINGS

7

2.1. INTRODUCTION

The relationship between exercise performance and arterial inflow is well documented

(25). Surprisingly, the contribution of the venous system to overall cardiovascular function and skeletal muscle activities has received considerably less attention, despite compelling evidence demonstrating its importance beyond the typical role as a “passive volume reservoir” (3, 11, 31)

Additionally, few studies have examined the influence of physical fitness

(cardiovascular and/or muscular) on indices of venous function. The majority of these studies report improvements in venous hemodynamics secondary to changes in blood

volume following endurance training (5, 16, 19). More recently, Wecht et al. attributed

venous modifications, in part, to enhanced vasomotor tone, vessel compliance and vessel

responses to mechanical and neural stimulation (33). Therefore, the major aim of this

study was to examine relationships between measures of physical fitness and arterial and

venous function. It was hypothesized that individuals with the highest levels of fitness

would have greater arterial inflow and venous outflow patterns.

However, considering the lack of standardization and reliability of the protocols

used to examine the venous function, a secondary aim of this study was to consider

several methodological and technical aspects regarding the examination of the venous

system.

2.2. METHODOLOGY

2.2.1. Study Design and Inclusion, Exclusion Criteria

This study consisted of 2 distinct phases. In phase I, 2 separate experiments were conducted to address methodological aspects (Experiment 1) concerning the examination

8 of venous function and to determine the reliability (Experiment 2) of vascular function using mercury in-silastic strain gauge plethysmography (MSGP). In phase II, 2 additional experiments were designed to examine the relationships of vascular function and cardiovascular fitness (Experiment 3) and handgrip strength (Experiment 4). Men and women between the age of 18 and 35 years were recruited to participate. Smokers and individuals with acute medical conditions (e.g. orthopedic injury), active infection and/or on pharmacotherapy with known vascular effects (e.g. anti-inflammatory therapy, cardiovascular medications) were excluded. Following comprehensive explanation of the study, its benefits, inherent risks and expected commitments with regard to time, all participants signed an informed consent approved by the Institutional Review Board of

Louisiana State University.

2.2.2. Vascular Assessments

Instrumentations and measurements

Forearm arterial inflow, vascular resistance, venous capacitance, and outflow for all experiments were obtained using MSGP (model EC5R, D. E. Hokanson Inc, Bellevue,

WA, 99). Upon arrival, blood pressure cuffs were positioned around the participant’s upper arm and wrist, and a mercury-in-silastic strain gauge (13) placed around the forearm approximately 10 cm distal to the olecranon process while the subject in a supine position. The forearm was extended and slightly supinated and supported by a styrofoam block. Immediately before the blood flow measurements, hand circulation was occluded for 1 min by inflating the cuff at the wrist to 240mmHg. Forearm blood inflow was estimated at rest and following 5 min of upper arm arterial occlusion. Subsequently forearm venous capacitance and outflow were obtained following specific venous

9 occlusion pressures and periods as will be described below in each experiment. Arterial occlusion was achieved by inflating the cuff on the upper arm to 240 mmHg (1, 29).

Throughout the testing procedures, heart rate and blood pressure were obtained at rest, and during and immediately following Occ.

Experiment 1: Effect of varying occlusion pressure and time on vascular indices.

This experiment was conducted on two separate days in eight apparently healthy young men and women (Age: 22.1 ± 1.4). Forearm blood inflow measures were obtained as described above in both days. Day 1: In order to examine the influence of varying upper arm venous occlusion pressures on venous capacitance and outflow, resting forearm venous indices were evaluated following four minutes of upper arm venous occlusion pressure at 7, 14, 21, 28, and 35 mmHg below diastolic blood pressure obtained in the opposite arm. Day 2: In order to examine the influence of upper arm venous occlusion durations, venous capacitance measures were taken every minute for a period of ten minutes, and venous outflow evaluated following 5 and 10 minutes of upper arm venous occlusion pressure using 50 mmHg.

Experiment 2: Reliability of arterial and venous function measurements.

Forearm blood inflow, and venous capacitance and outflow were obtained in 8 apparently healthy young men (Age: 24.0 ± 5.0) on three different occasions within 10 days to determine between-days reliability of MSGP. Vascular indices were evaluated at rest and following 5 min of upper arm arterial occlusion. Venous capacitance and outflow for the two conditions (rest & post occlusion) were evaluated following 6 minutes of upper arm venous occlusion at 7 mmHg below diastolic blood pressure.

10 Experiment 3 & 4: Associations between physical fitness measures and arterial and venous function.

These experiments were designed to examine relationships between forearm arterial and

venous function indices and measures of physical fitness (estimated VO2peak and handgrip strength). Cardiorespiratory fitness (Estimated VO2peak) was determined using the YMCA’s cycle exercise test (2) in twenty-one apparently healthy young (Age: 21.4 ±

1.4) women and expressed in mL•kg-1•min-1. Maximum handgrip strength was

determined, in twenty-two apparently healthy young (Age: 24.5 ± 5.0) men, as the

average of 3 maximum consecutive contraction trials using a hand-dynamometer (model

78010, Lafayette Instrument, Lafayette, IN, 02) and expressed in pound. Vascular indices

were evaluated at rest and following 5 min of upper arm arterial occlusion. Venous

capacitance and outflow for the two conditions (rest & post occlusion) were evaluated following 6 minutes of upper arm venous occlusion at 7 mmHg below diastolic blood pressure.

2.2.3. Data Analysis

Arterial indices

Resting forearm blood inflow was recorded at a paper speed of 5 mm/second and values

were derived from the slope drawn at a best-fit tangent using the first 3 pulses.

Calculations were made as a function of 60 seconds divided by the horizontal distance

(mm) needed for the slope to rise vertically from baseline to the top of the recording paper and multiplied by the full chart range. Peak arterial in-flow following occlusion

was recorded at a paper speed of 25 cm/ second. Analyses were performed using a slope

drawn at a best-fit tangent to the curves of the first 2 pulses of the flow curve post cuff

release. The blood flows were then calculated from 60 sec. multiplied by the paper speed

11 (25 cm/sec) divided by the horizontal distance (mm) needed for the volume slope to increase by 20 mm vertically. Mean arterial pressure (MAP) was calculated using the standard equation: MAP = Diastolic Pressure+[(Systolic Pressure-Diastolic Pressure)/3].

Forearm vascular resistance (FVR) was calculated by dividing mean arterial pressure by forearm blood inflow (FVR = MAP/ forearm blood inflow).

Venous indices

Forearm venous capacitance was measured as the vertical distance (mm) representing the increase in forearm-volume graph after the designated period (as described for each experiment) for venous filling. Analysis of forearm venous outflow was derived from a tangent line that represents the vertical drop in volume-graph from the excursion line and drawn at 0.5 second and 2 seconds after the release of the venous occlusion pressure (7,

12).

2.2.4. Statistical Analysis

All statistical analyses were performed using SPSS statistical software package for windows (version 11.0, Chicago, Ill). Group data were expressed as mean ± sd and alpha was set a priori at P < 0.05.

Experiment 1: Effect of varying occlusion pressure and time on vascular indices.

In order to evaluate the difference in forearm inflow, and venous capacitance and outflow following each venous occlusion pressure, repeated measure ANOVA and Bonferroni pairwise comparison were used. Dependent t-tests were used to compare vascular function measures at 5 and 10 minutes.

12 Experiment 2: Reliability of arterial and venous function measurements.

Interclass correlation coefficient (ICCC) was used to determine the reliability of MSGP.

Measures of the arterial and venous function indices at rest and following occlusion were included.

Experiment 3 & 4: Relationship between physical fitness measures and arterial and venous function indices

Pearson product moment correlation was used to examine the relationship between physical fitness measures (estimated VO2peak & handgrip strength) and arterial and venous function indices as well as the relationship between arterial and venous function indices. Additionally, with participants subdivided into above (HI) and below (LO) average fitness groups according to the YMCA and the Canadian Society of Exercise

Physiology (22), and independent t tests were used to compare fitness measures and vascular indices.

2.3. RESULTS

2.3.1. Participants Characteristics

Table 2.1. Subject characteristics Mean ± sd Range Age (yrs) 22.6 ± 3.5 18-34 Height (cm) 167.5 ± 8.5 150-185 Weight (kg) 69.1 ± 15.2 43-114 Arm length (cm) 25.5 ± 2.2 20-30 Arm circumference (cm) 25.17 ± 3.0 21-35 HR (bpm) 59.0 ± 10.0 45-93 SBP (mmHg) 107.0 ± 10.0 87-142 DBP (mmHg) 62.0 ± 8.0 46-86 MAP (mmHg) 76.0 ± 8.0 60-105

13 A total of 55 individuals (28 men and 27 women) volunteered to participate in these experiments. The participants’ characteristics are presented in Table 2.1.

Experiment 1: Effect of varying venous occlusion pressure and time on vascular indices

The results for experiment 1 are depicted in Table 2.2 as well as Figures 2.1, 2.2 and 2.3.

There was no significant affect in altering the venous occlusion pressure and time on forearm resting or reactive hyperemic blood flow (Table 2.2). The data indicate a stepwise increase in measures of forearm venous capacitance (Figure 2.1a) and outflow

(Figure 2.1b) with increasing venous occlusion pressures. In contrast, forearm inflow was not affected by the pressure maneuvers. Furthermore, forearm venous capacitance and outflow were consistently greater following 10 minutes compared to 5 minutes of venous occlusion (Figure 2.2). Finally, venous capacitance showed a stepwise increase as the venous occlusion time increased (Figure 2.3).

Table 2.2. Resting vascular indices following varying occlusion pressures Press 1 Press 2 Press 3 Press 4 Press 5 (70.05 vs. Press 3; †=p>0.05 vs. Press 4; ‡=p>0.05 vs. Press 5. RBF = Resting blood inflow; FVR = forearm vascular resistance.

Experiment 2: Reliability of MSGP to measure arterial and venous function indices.

Between-day reliability values using the ICCC are presented in Table 2.3. The ICCC values for the peak blood inflow and vascular resistance were 0.81 and 0.75, respectively.

14 Additionally, resting and following upper arm arterial occlusion venous capacitance and outflow ICCC values ranged between 0.70 and 0.94.

6 1 = 07mmHg < DBP 2 = 14mmHg < DBP 5 3 = 21mmHg < DBP 4 = 28mmHg < DBP 5 = 35mmHg < DBP 4 ‡ ‡ ‡ * § 3 * Venous capacitance

(ml·100ml tissue-1·min-1 2

1 12345 Venous occlusion pressures (mmHg)

Figure 2.1a. Venous capacitance using different pressures. Values are in means ± SD. ‡p<0.007 vs. 1; *p<0.01 vs. 2; §p<0.03 vs. 3.

60 1 = 07mmHg < DBP 2 = 14mmHg < DBP 3 = 21mmHg < DBP 50 4 = 28mmHg < DBP 5 = 35mmHg < DBP 40 ‡ ‡ ‡ § 30 * *

Venous outflow †

(ml·100ml tissue-1·min-1 20

10 12345 Venous occlusion pressures (mmHg)

Figure 2.1b. Venous outflow using different pressures. Values are in means ± SD. ‡p<0.02 vs. 1; *p<0.02 vs. 2; §p<0.05 vs. 3; †p<0.06 vs. 4.

15

40 5

)

4 30 ‡

3 outflow Venous 20

Venous capacitanceVenous (ml·100ml tissue-1·min-1 (ml·100ml (ml·100ml tissue-1·min-1) (ml·100ml

2 10 5Minutes 10Minutes 5Minutes 10Minutes

Time Time

Figure 2.2. Venous capacitance and outflow following 5 & 10 minutes of venous occlusion. Values are in means ± SD. ‡p<0.05 vs. 10 minutes.

5 ‡ ‡ ‡ 4 ‡ ‡ ‡ ‡ ‡ 3 ‡

2

Venous capacitance 1 (ml·100ml tissue-1·min-1

0 12345678910 Time (Minutes)

Figure 2.3. Venous capacitance over a 10 minute venous occlusion period. Values are in means ± SD. ‡p<0.05 vs. minute 1.

16

Table 2.3. Between-days ICCC for forearm arterial and venous indices Conditions RBF FVR Venous Venous (ml/100ml/min) (U) Capacitance Outflow (ml/100ml/min) (ml/100ml/min) Rest 0.29 0.30 0.78 0.94 Post occlusion 0.81 0.75 0.70 0.91 ICCC = interclass correlation coefficient; RBF = resting blood inflow; FVR = forearm vascular resistance.

Experiment 3 & 4: Relationship between physical fitness measures and arterial and venous function indices.

The associations between measures of physical fitness and vascular function are presented in Table 2.4 and depicted in Figures 2.4 and 2.5. Estimated VO2peak was significantly correlated with peak forearm blood inflow and vascular resistance, and with venous outflow. Further analyses revealed a significant difference in vascular function between those with higher VO2peak estimates (46.4 ± 6.6 ml/kg/min) compared to the lower fit group (32.3 ± 3.4 ml/kg/min).

65 0.54; p = 0.012

55

45 (mL/kg/min) 2

VO 35

25 10 20 30 Arterial inflow (mL· 100mL tissue-1· min-

Figure 2.4a. Relationship between estimated VO2 & post arterial occlusion blood inflow.

17 60 0.56; p = 0.016

50

40 (mL/kg/min) 2

VO 30

20 25 35 45 55 65 Venous outflow (mL· 100mL tissue-1· min-

Figure 2.4b. Relationship between estimated VO2 & resting venous outflow.

Handgrip strength was significantly correlated with venous capacitance and outflow at

rest and following occlusion (see Figures 2.5 a & b). Moreover, as presented in Table 2.6,

venous function indices were different between the HI (121.1 ± 10.64 lb) vs. LO (82.14 ±

14.9 lb) handgrip strength. No significant associations were found between handgrip

strength and forearm blood inflow at rest or following occlusion (see Table 2.4).

Finally, resting and peak forearm arterial inflow and vascular resistance correlated with resting and post occlusion venous capacitance and outflow (see Table 2.4).

2.4. DISCUSSION

This study was designed to examine relationships between measures of physical fitness and arterial and venous function. In addition, methodological concerns and the reliability of arterial and venous function measures using MSGP were also addressed. Uniquely, the data indicate the magnitude of forearm venous outflow is positively associated with both cardiovascular (estimated VO2peak) and musculoskeletal (handgrip strength) fitness.

Moreover, venous function measures are significantly influenced by the magnitude of the

occlusion pressure and length of occlusion period without changes in arterial inflow.

18 Finally, measures of arterial as well as venous function appear to be reasonably reproducible over a 10-day period.

7 r=0.57; p=0.007

6

5

Venous outflow 4 (mL·100mL tissue-1·min-1)

3 40 60 80 100 120 140 Strength (lb)

Figure 2.5a. Relationship between handgrip strength & resting venous capacitance.

r=0.67; p=0.001 70

60

50

40 Venous outflow 30 (mL·100mL tissue-1·min-1)

20 40 60 80 100 120 140 Strength (lb)

Figure 2.5b. Relationship between handgrip strength & resting venous outflow.

19 Table 2.4. Pearson product moment correlation between arterial, venous and physical fitness measures

Conditions Resting vascular indices Post occlusion vascular indices Fitness measures Blood Vascular Venous Venous Blood Vascular Venous Venous Estimated Handgrip Variables inflow resistance capacitance outflow inflow resistance capacitance outflow VO2peak strength Blood inflow 1.0 (mL/100mL/min) Vascular resistance r=-0.9; 1.0 (U) p=0.001

indices Venous capacitance r=0.5; r=-0.4; 1.0 (mL/100mL/min) p=0.001 p=0.012

Resting vascular vascular Resting Venous outflow r=0.4; r=-0.5; r=0.76; 1.0 (mL/100mL/min) p=0.013 p=0.002 p=0.0001 r Blood inflow r=-0.007; r=-0.18; r=0.18; r=0.38; 1.0 (mL/100mL/min) p=0.9 p=0.26 p=0.27 p=0.018 Vascular resistance r=-0.3; r=0.06; r=0.13; r=-0.2; r=-0.83; 1.0 (U) p=0.13 p=0.7 p=0.45 p=0.19 p=0.0001

indices Venous capacitance r=0.52; r=-0.54; r=0.7; r=0.6; r=0.2; r=0.005; 1.0 (mL/100mL/min) p=0.001 p=0.0001 p=0.0001 p=0.0001 p=0.2 p=0.975 Venous outflow r=0.54; r=-0.63; r=0.77; r=0.85; r=0.4; r=-0.27; r=0.64; 1.0

Post occlusion vascula (mL/100mL/min) p=0.001 p=0.0001 p=0.0001 p=0.0001 p=0.02 p=0.12 p=0.0001

Estimated VO2peak r=0.04; r=-0.3; r=-0.08; r=0.56; r=0.54; r=-0.7; r=0.2; r=0.44; 1.0 xxxx (mL/kg/min) p=0.876 p=0.2 p=0.74 p=0.016 p=0.012 p=0.001 p=0.410 p=0.07

Fitness Handgrip strength r=0.02; r=-0.05; r=0.57; r=0.67; r=0.2; r=0.07; r=0.6; r=0.5; xxxx 1.0 measures (kg) p=0.93 p=0.8 p=0.007 p=0.001 p=0.4 p=0.76 p=0.004 p=0.005

20 Table 2.5. Differences between the cardiorespiratory groups in forearm arterial and venous indices Group Resting Condition Post Occlusion. Condition Binf Vres Binf Vres Arterial LO 1.6 ± 0.7 52.1 ± 18.7 18.5 ± 2.87 4.3 ± 0.7 Indices HI 1.6 ± 0.5 46.7 ± 13.8 22.2 ± 4.9 3.4 ± 0.8 p-value 0.46 0.91 0.05 0.013 Vcap Vout Vcap Vout Venous LO 3.8 ± 0.8 40.7 ± 7.9 2.6 ± 0.7 36.9 ± 8.2 Indices HI 4.0 ± 0.7 49.5 ± 9.7 2.9 ± 0.8 48.9 ± 9.7 P-value 0.54 0.05 0.30 0.014 Binf = blood inflow; Vres = vascular resistance; Vcap = venous capacitance; Vout = venous outflow.

Table 2.6. Differences between the handgrip strength groups in forearm arterial and venous indices Group Resting Condition Post Occlusion. Condition Binf Vres Binf Vres Arterial LO 3.48 ± 1.2 26.3 ± 12.4 19.39 ± 3.7 4.27 ± 1.0 Indices HI 3.53 ± 0.9 24.32 ± 4.9 20.60 ± 4.1 4.24 ± 0.8 p-value 0.94 0.73 0.55 0.94 Vcap Vout Vcap Vout Venous LO 4.9 ± 0.8 45.0 ± 10.7 3.61 ± 0.60 43.8 ± 10.0 Indices HI 5.8 ± 0.6 56.2 ± 5.3 4.47 ± 0.60 54.6 ± 7.2 P-value 0.030 0.038 0.014 0.04 Binf = blood inflow; Vres = vascular resistance; Vcap = venous capacitance; Vout = venous outflow.

2.4.1. Technical Aspects and Reliability of Vascular Measures

In recent years the post occlusion reactive hyperemia model has been developed as a non- invasive surrogate marker of cardiovascular risk (4, 6). Previous work from our laboratory has indicated adequate stability and reproducibility of the post-occlusion technique for the study of arterial reactivity under strictly controlled conditions using

21 both ultrasonography (34) and plethysmography (20). Given recent information regarding

the importance of the venous system on arterial function (3, 11, 31), the present study

was designed to include information regarding venous function. Clearly, if the post-

occlusion model is extended to include information regarding venous function, it is

essential the reproducibility and limitations of the technique be established.

Data from experiment 1 indicates the importance of the selection of an

appropriate venous occlusion pressure in the study of venous capacitance and outflow.

The data reveals a stepwise decrease in venous capacitance and outflow as the magnitude

of venous occlusion pressure from resting diastolic pressure was increased (see Figures

2.1 a & b). The rational for the examination of varying venous occlusion pressures is the lack of consensus regarding the selection of a standard pressure. Two different approaches in the selection of venous pressure appear to exist. The first is the use of a standard (usually 50 mmHg) venous occlusion pressure (7, 12) and the second is the use of a venous occlusion pressure based on the individual’s diastolic pressure (33). The present findings indicate that a venous occlusion pressure of 7 mmHg below diastolic blood pressure appears to yield greater venous capacitance and outflow compared to occlusion pressures further below the diastolic pressure. This is particularly important when considering longitudinal trials in which blood pressure is subject to change.

A second inconsistency in the literature is the duration of venous occlusion.

Standard recommendations suggest venous occlusion periods of 2 minutes to examine venous capacitance and outflow (7, 12). The present data (see Figure 2.2) clearly indicate that venous capacitance and outflow were greater following 10 minutes of venous occlusion at 50 mmHg compared to 5 minutes. The magnitude of the difference was

22 particularly evident when comparing occlusion periods of 1 through 6 minutes, whereas

the difference in the venous measures was less “drastic” with occlusion periods between

7 and 10 minutes (see Figure 2.3).

Thus, when implementing these measures in longitudinal trials careful

consideration should be given regarding the venous occlusion pressure and period.

Interestingly, alterations in venous occlusion pressures and times did not have a

significant effect on the arterial inflow measures.

Experiment 2 examined the reliability of measuring arterial as well as venous function using MSGP at rest and following occlusion over a 10 day period (see Table

2.3). The magnitude of the forearm blood flow values at rest and following occlusion are consistent with data from our laboratory (1) as well as others (29). The reliability measures used in the current study confirm adequate consistency between days for the post occlusion in flow measures (20, 24). In addition, the ICCC results indicate adequate reliability for venous capacitance and outflow at rest and following arterial occlusion.

Resting blood in flow measures appear to vary considerably from visit to visit, despite the fact careful standardization of many external factors, such as meal timing, testing

schedule, testing environment, subject placements and clothing. This appears to be a

consistent inconsistency in the literature and is difficult to rationalize (20, 24, 29). The

values for these measures and the range from day to day are consistent with previously reported values from our (20) as well as other laboratories (24). However, it is important to consider the influence of the “wandering” baseline on the calculation of the blood flow difference between the peak blood flow response following the period of arterial occlusion and resting blood flow. The use of the absolute reactive hyperemic response

23 would therefore appear to be more appropriate when using this model during longitudinal

trials.

2.4.2. Relationship Between Fitness and Vascular Function

The ability to distribute and deliver adequate blood flow, oxygen and nutrients to

contracting skeletal muscle is recognized as a major contributor to exercise performance

(25, 27). Data from the current study are consistent with this idea in that forearm vascular

reactivity defined as the peak blood flow response following arterial occlusion was

significantly related to estimated VO2peak (see Figure 2.4). It is believed the ability for

greater performance is in part secondary to the repeated stimulus of exercise training

which over time contributes to a series of arterial adaptations aimed at increasing

vascular conductance and oxygen and nutrient delivery to working skeletal muscle (18).

The main finding in the present study is the consistent associations between

measures of venous function and cardiovascular and musculoskeletal fitness. These findings further confirm data from our laboratory suggesting that exercise tolerance in

heart failure patients is related to venous function (35). The associations suggest the

contribution of the venous system to overall cardiovascular performance and skeletal

muscle activities is beyond the typical role as a “passive volume reservoir”.

Uniquely, the data of this study reveal an association between venous outflow and

estimated VO2peak; and greater venous outflow in the HI cardiovascular fitness group.

These results suggest that factors associated with venous function might play a role in

exercise performance. Considering the cardiovascular system is a close circuit, as well as

the venous circulation enormous potential to act as a volume reservoir and its contribution to the end-diastolic volume (8), this segment of the vasculature might have a

24 profound impact on exercise performance. We hypothesize improved venous outflow

may contribute to enhanced exercise performance secondary to greater metabolic waste

removal from exercising muscles, increased venous return, and/or enable greater arterial

inflow (3, 8, 11, 31).

Another unique finding in this study is the association between measures of

venous function and handgrip strength suggesting a link between muscular fitness and

venous circulation. Further comparison reveals greater venous function measures in the

HI handgrip strength group suggesting the importance of muscular fitness to venous function. Interestingly, individuals in the HI strength group reported long-term involvement in weight training. Given the numerous benefits of resistive exercise training in cardiovascular performance (9, 21, 37), as well as modification of cardiovascular disease risk profile and overall health (15, 32, 36), the current findings suggest that resistance training might also benefit the venous system. This is particularly relevant considering venous abnormalities in aging (17, 23), and patients with heart failure (14,

28, 35) and venous diseases (10, 30). Therefore, recognizing the contribution of skeletal muscle in venous return and overall cardiovascular performance during exercise, factors associated with muscle strength modification might contribute to increased overall cardiovascular health subsequent to improved venous function, namely venous outflow and capacitance. However, it is recognized that the current study is limited to a cross- sectional comparison and future studies should be designed to directly examine whether resistance training and subsequent strength modifications can indeed alter venous function in both health and disease.

25 The association between arterial and venous measures suggests an

intercommunication link between the arterial and venous circulations. Interestingly,

compelling evidence has emerged illustrating the importance of various aspects of the

venous system beyond its typical role as a “passive volume reservoir”. For example,

Barclay et al. demonstrated that muscle fatigue was delayed as a result of increased blood flow across the muscle bed and subsequent increased metabolic removal independent of oxygen and/or nutrient delivery (3). In a more recent study, Tschakovsky & Hughson reported greater arterial inflow patterns with increased venous emptying following arm elevation (31). The authors linked these findings to the local venoarteriolar sympathetic

axon reflex. The presence of nerve fiber collaterals from the sympathetic arteriolar plexus to adjacent reflex was initially described in 1991 by Rygaard et al. and may serve as the anatomical substrate for the local venoarteriolar sympathetic axon reflex (26).

Finally, evidences of cross-communication between the arterial and venous circulations

indicate that venular endothelium releases a relaxing factor (s) that contributes to the

vasodilatation of adjacent (11). The optimal position of the venous circulation

might provide a monitor of the tissue metabolic status and a potentially valuable feedback

mechanism to control the arterial circulation. Therefore, the relationship between arterial

and venous function indices, suggests that vasoreactivity and muscle perfusion might be

influenced by venous hemodynamics confirming the essential role of venous function in

overall cardiovascular control, muscle contraction and exercise performance (3, 11, 31).

2.5. CONCLUSION

These findings indicate the importance of recognizing the possible influence of venous

occlusion times and pressures on venous capacitance and outflow. Additionally, the

26 ICCC results indicate that MSGP is a reliable technique for measuring venous function on different days under controlled conditions and useful to differentiate between populations. Considering these technical aspects and establishing the reliability of MSGP are particularly relevant during cross-sectional as well as repeated measures designs, especially in cases when blood pressure may be altered. Finally, forearm venous outflow is positively associated with measures of cardiovascular (estimated VO2peak) and musculoskeletal (handgrip strength) fitness. The association between handgrip strength and venous system hemodynamics implies that muscular conditioning might be important for venous health, which warrants future studies examining the effect of muscular strength modifications on venous function.

2.6. REFERENCES

1.Alomari MA, Welsch MA, Prisby RD, Lee CM and Wood RH. Modification of forearm vascular function following short-term handgrip exercise training. Int J Sports Med 22: 361-365., 2001.

2.American College of Sports Medicine., Roitman JL and Herridge M. ACSM's resource manual for Guidelines for exercise testing and prescription. Philadelphia: Lippincott Williams & Wilkins, 2001.

3.Barclay JK. A delivery-independent blood flow effect on skeletal muscle fatigue. J Appl Physiol 61: 1084-1090, 1986.

4.Celermajer DS, Sorensen KE, Gooch VM, Spiegelhalter DJ, Miller OI, Sullivan ID, Lloyd JK and Deanfield JE. Non-invasive detection of endothelial dysfunction in children and adults at risk of atherosclerosis. Lancet 340: 1111-1115, 1992.

5.Convertino VA, Mack GW and Nadel ER. Elevated central venous pressure: a consequence of exercise training-induced hypervolemia? Am J Physiol 260: R273-277, 1991.

6.Corretti MC, Anderson TJ, Benjamin EJ, Celermajer D, Charbonneau F, Creager MA, Deanfield J, Drexler H, Gerhard-Herman M, Herrington D, Vallance P, Vita J and Vogel R. Guidelines for the ultrasound assessment of endothelial-dependent flow-mediated vasodilation of the brachial artery: a report of the International Brachial Artery Reactivity Task Force. J Am Coll Cardiol 39: 257-265, 2002.

27

7.Cramer M, Langlois Y, Beach K, Martin D and Strandness Jr DE. Standardization of venous flow measurements by strain gauge plethysmography in normal subjects. Bruit VII: 33-39, 1983.

8.Guyton AC, Richardson TQ and Langston JB. Regulation of cardiac output and venous return. Clin Anesth 3: 1-34, 1964.

9.Hagerman FC, Walsh SJ, Staron RS, Hikida RS, Gilders RM, Murray TF, Toma K and Ragg KE. Effects of high-intensity resistance training on untrained older men. I. Strength, cardiovascular, and metabolic responses. J Gerontol A Biol Sci Med Sci 55: B336-346, 2000.

10.Hartmann BR, Drews B and Bassenge E. Venous Function in Patients with Venous Disease and Healthy Controls Before and After a Bathing Procedure and Subsequent Cold Stimulus. International Journal of Angiology 7: 252-254, 1998.

11.Hester RL and Hammer LW. Venular-arteriolar communication in the regulation of blood flow. Am J Physiol Regul Integr Comp Physiol 282: R1280-1285, 2002.

12.Hokanson DE, Sumner DS and Strandness Jr. DE. An electrically calibrated plethysmograph for direct measurement of limb blood flow. IEEE transactions on bio- medical engineering 22: 21-25, 1975.

13.Holling HE, Boland HC and Russ E. Investigation of arterial obstruction using a mercury-in-rubber strain gauge. Am Heart J 63: 194-205, 1961.

14.Ikenouchi H, Iizuka M, Sato H, Momomura S, Serizawa T and Sugimoto T. Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Jpn Heart J 32: 17-34, 1991.

15.Kell RT, Bell G and Quinney A. Musculoskeletal fitness, health outcomes and quality of life. Sports Med 31: 863-873, 2001.

16.Kenney WL and Armstrong CG. The effect of aerobic conditioning on venous pooling in the foot. Med Sci Sports Exerc 19: 474-479, 1987.

17.Lanne T and Olsen H. Decreased capacitance response with age in lower limbs of humans--a potential error in the study of cardiovascular reflexes in ageing. Acta Physiol Scand 161: 503-507, 1997.

18.Laughlin MH. Endothelium-mediated control of coronary vascular tone after chronic exercise training. Med Sci Sports Exerc 27: 1135-1144, 1995.

28 19.Louisy F, Jouanin JC and Guezennec CY. Filling and emptying characteristics of lower limb venous network in athletes. Study by postural plethysmography. Int J Sports Med 18: 26-29, 1997.

20.Mahankali B, Allen J, Mann A, Graves A and Welsch MA. Reproducibility of forearm blood flow using mercury-in silastic strain gauge plethysmography. Med Sci Sports Exerc, Suppl. 29: S258(abstract), 1997.

21.Maiorana A, O'Driscoll G, Cheetham C, Dembo L, Stanton K, Goodman C, Taylor R and Green D. The effect of combined aerobic and resistance exercise training on vascular function in type 2 diabetes. J Am Coll Cardiol 38: 860-866, 2001.

22.Nieman DC. Exercise testing and prescription : a health related approach. Boston: McGraw-Hill, 2003.

23.Olsen H and Lanne T. Reduced venous compliance in lower limbs of aging humans and its importance for capacitance function. Am J Physiol 275: H878-886, 1998.

24.Roberts DH, Tsao Y and Breckenridge AM. The reproducibility of limb blood flow measurements in human volunteers at rest and after exercise by using mercury-in-Silastic strain gauge plethysmography under standardized conditions. Clin Sci (Lond) 70: 635- 638, 1986.

25.Rowell LB. Human circulation : regulation during physical stress. New York: Oxford University Press, 1986.

26.Rygaard K, Moller M and Henriksen O. Presence of nerve fiber collaterals from the sympathetic arteriolar plexus the concomitant venules in dog skeletal muscle. Acta Physiol Scand Suppl 603: 115-118, 1991.

27.Saltin B, Radegran G, Koskolou MD and Roach RC. Skeletal muscle blood flow in humans and its regulation during exercise. Acta Physiol Scand 162: 421-436, 1998.

28.Sato H, Ikenouchi H, Aoyagi T, Matsui H, Mochizuki T, Momomura S, Serizawa T, Iizuka M and Sugimoto T. Pathophysiology and evaluation of severity of congestive heart failure on the basis of venous characteristics. Jpn Circ J 53: 141-145, 1989.

29.Sinoway LI, Musch TI, Minotti JR and Zelis R. Enhanced maximal metabolic vasodilatation in the dominant forearms of tennis players. Journal of Applied Physiology 61: 673-678, 1986.

30.Tretbar LL. Chronic venous insufficiency of the legs: pathogenesis of venous ulcers. J Enterostomal Ther 14: 105-108, 1987.

29 31.Tschakovsky ME and Hughson RL. Venous emptying mediates a transient vasodilation in the human forearm. Am J Physiol Heart Circ Physiol 279: H1007-1014, 2000.

32.Warburton DE, Glendhill N and Quinney A. The effects of changes in musculoskeletal fitness on health. Can J Appl Physiol 26: 161-216, 2001.

33.Wecht JM, de Meersman RE, Weir JP, Bauman WA and Grimm DR. Effects of autonomic disruption and inactivity on venous vascular function. Am J Physiol Heart Circ Physiol 278: H515-520, 2000.

34.Welsch MA, Allen JD and Geaghan JP. Stability and reproducibility of brachial artery flow-mediated dilation. Med Sci Sports Exerc 34: 960-965, 2002.

35.Welsch MA, Alomari M, Parish TR, Wood RH and Kalb D. Influence of venous function on exercise tolerance in chronic heart failure. J Cardiopulm Rehabil 22: 321- 326., 2002.

36.Winett RA and Carpinelli RN. Potential health-related benefits of resistance training. Prev Med 33: 503-513, 2001.

37.Wood RH, Reyes R, Welsch MA, Favaloro-Sabatier J, Sabatier M, Matthew Lee C, Johnson LG and Hooper PF. Concurrent cardiovascular and resistance training in healthy older adults. Med Sci Sports Exerc 33: 1751-1758, 2001.

30 CHAPTER 3. INFLUENCE OF VENOUS FUNCTION ON EXERCISE TOLERANCE IN CHRONIC HEART FAILURE*

*Reprinted by permission of Lippincott Williams & Wilkins publisher (See appendix C)

31 3.1. INTRODUCTION

Chronic heart failure (HF) is a complex clinical syndrome characterized by an

abnormality of the heart and by a recognizable pattern of hemodynamic, renal, neural and

hormonal responses (24). The manifestations of the profound multi-organ changes often become most evident during acute bouts of physical activity as many patients experience shortness of breath and early fatigue compared to age-matched controls. Indeed, much

research has focused on the role of vascular and skeletal muscle alterations on exercise

tolerance. Blood flow to exercising skeletal muscle is reduced in chronic HF patients

(32). The proposed mechanisms for the reduction in blood flow appear to be quite

complex but include cardiac insufficiency, neurohumoral hyperactivity (eg, sympathetic

nervous system, angiotensin and endothelin mediated), and neurohumoral hypoactivity

(eg, endothelium-derived-relaxing factors), and vascular remodeling (3). In addition, decreased skeletal muscle mass, alterations in fiber type composition and metabolic milieu further modify the regional distribution of blood flow (3). As a result of these findings, additional treatment strategies (eg, angiotensin- converting enzyme inhibitors

and exercise training) have been developed and implemented in the care of the chronic

HF patient. Yet, despite these interventions many patients continue to experience activity-related symptoms, possibly suggesting the involvement of other factors.

Interestingly, the role of the venous system on exercise tolerance in patients with HF has not received much attention, despite evidence of changes in venous structure and

function in patients with HF (14, 27) and in patients with a variety of vascular diseases

(20, 29). Considering that the cardiovascular system is a closed circuit, cardiac

insufficiency also may have a profound effect on venous function. Accordingly, the

32 purpose of this study was to examine the relationships between indices of arterial and venous function and exercise tolerance in chronic HF patients and age-matched controls.

Given the complexity of the HF syndrome, we hypothesized a direct relationship between indices of vascular function and exercise tolerance.

3.2. METHODS

3.2.1. Subjects Class I and III (New York Heart Association Functional Classification) HF patients and age-matched controls were recruited to participate in the study. Individuals recovering from recent hospitalization, with unstable myocardial ischemia and/or angina, uncontrolled diabetes mellitus and/or hypertension, anemia, disease, or renal failure were excluded from participation. The patients continued to take the dose of medications as prescribed by their physicians. After comprehensive explanation of the proposed study, it benefits, inherent risks, and expected commitments, all participants signed an informed consent approved by the Institutional Review Board of Louisiana State

University.

3.2.1. Assessment of Vascular Function Forearm vascular indices were obtained using mercury in-silastic strain-gauge plethysmography (EC-5R system, Hokanson, Wash). Prior to the experiment, blood pressure cuffs were positioned around the participant’s upper arm and wrist, and a mercury-in-silastic strain gauge placed around the forearm approximately 10 cm distal to the olecranon process (1). Resting vascular indices were made in duplicate, after 15 minutes of supine rest, with the measures separated by 5 minutes. Immediately before obtaining measurements, hand circulation was occluded for 1 minute by inflating the cuff at the wrist to 240mmHg (1). Forearm arterial inflow was obtained using an upper arm

33 venous collecting pressure of 50mmHg (1). Forearm venous capacitance and venous

outflow were measured after an additional 2 minutes of venous occlusion (1).

Subsequently, forearm vasoreactivity was examined after 5 minutes of forearm arterial

occlusion achieved by inflating the cuff on the upper arm to 240mmHg (1). Forearm

vascular indices were then determined as described above. Throughout the experiment,

heart rate and blood pressure were obtained at rest, during and immediately after forearm

occlusion.

3.2.3. Assessment of Exercise Tolerance Exercise tolerance was measured on a motor driven treadmill as the maximum walking

distance achieved on a 6-minute walk test (9). Participants were instructed to walk as far

as possible in 6 minutes, avoiding chest pain, shortness of breath, or other symptoms.

During the test the participants controlled the speed of the treadmill and allowed to

change the speed at any time. The participants was not coached during the test, but made

aware of time remaining to completion. Throughout the test, blood pressure and heart rate

were monitored continuously by an exercise physiologist.

3.2.4. Data Analysis Resting forearm arterial inflow was recorded at a paper speed of 5 mm/sec. and values

were derived from the slope drawn at a best-fit tangent using the first 3 pulses.

Calculations were made as a function of 60 seconds divided by the horizontal distance

(mm) needed for the slope to rise vertically from baseline to the top of the recording paper and multiplied by the full chart range (4). Forearm arterial inflow responses after occlusion were recorded at a paper speed of 25 mm/sec. Analyses was performed using a slope drawn at a best-fit tangent to the curves of the first 3-pulse flows post cuff release.

The blood flows were then calculated from 60 seconds multiplied by the paper speed (25

34 mm/sec.) divided by the horizontal distance (mm) needed for the volume slope to

increase by 20 mm vertically (24). Forearm vascular resistance was calculated by

dividing mean arterial pressure by forearm arterial inflow. Forearm venous indices at rest

and following occlusion were recorded at a paper speed of 25mm/sec. Forearm venous

capacitance was calculated as the vertical distance (mm) representing the increase in

forearm-volume graph after 2 minutes of venous filling (4). Analysis of forearm venous

outflow was derived from a tangent line that represents the vertical drop in the volume-

graph from the excursion line and drawn at 0.5 seconds after the release of the venous

occlusion pressure (4). All vascular indices are expressed in mL • tissue-1 • min-1.

3.2.5. Statistical Analysis Statistical analysis was performed using SPSS for Windows (version 10.0,Chicago, Ill).

Group values are expressed as mean ± SD. Independent t tests were used to compare

vascular indices and maximum walking distance between the HF and control groups.

Pearson product moment correlation was used to examine relationships between vascular

indices and maximum walking distance. Alpha was set a priori at p<.05.

3.3. RESULTS

3.3.1. Patient Characteristics and Compliance Twenty HF patients (age 57 ± 15) and age 10 age-matched controls (age 54 ± 17 years)

volunteered to participate in the study (Table 1). Seventy-five percent of the patients were

black and 55% were women. All patients were recruited from lower socioeconomic areas.

The major etiologies of HF in this population included coronary artery disease (5 of 20),

hypertensive heart disease (15 of 20), and idiopathic cardiomyopathy (5 of 20). Due to

35 equipment failure, forearm venous capacitance and venous capacitance and venous outflow after occlusion were not obtained in one and three HF subjects, respectively.

Table 3.1. Participant characteristics CHF (n=20) Control (n=9) Mean ± sd Range Mean ± sd Range p-value Age (yr.) 59 ± 13 33-80 51 ± 16 32-74 0.18

Height (cm) 170 ± 8 155-180 169 ± 10 152-180 0.78

Weight (kg) 97 ± 27 58-165 74 ± 12 57-91 0.03

Resting Heart Rate 72 ± 17 45-98 64 ± 8 55-82 0.24 (beats/min) Resting Systolic Blood Pressure 139 ± 12 117-164 131 ± 13 114-155 0.14 (mmHg) Resting Diastolic Blood Pressure 79 ± 9 60-95 75 ± 6 69-86 0.19 (mmHg) Maximum Walking Distance 178 ± 65 75-321 563 ± 136 346-756 0.0001 (m) Etiology of CHF CAD 5/20 N/A HTN 15/20 ICM 5/20 Other 4/20 Systolic Failure 18/20

Left Ventricular Ejection 38 ± 18 10-70 N/A Fraction (%) CAD= Coronary Artery Disease; HTN=Hypertension; ICM=Idiopathic Cardiomyopathy.

3.3.2. Exercise Tolerance The maximum walking distance achieved on the 6-min walk test was significantly lower in the patients with HF compared to the control group (HF: 171 ± 75 m versus controls:

541.1 ± 145 m; p=0.0001).

36 3.3.3. Vascular Function As depicted in Table 2, forearm vascular indices at rest were not different between

Table 3.2. Independent t-test comparisons between CHF and control groups. Heart Failure Control t p (n=20) (n=9) Forearm arterial inflow (ml•100ml tissue-1 •min-1) 3.20 ± 1.20 3.46 ± 1.17 -0.63 0.53 Rest 14.8 ± 6.31 22.20 ± 6.30 -2.94 0.007 Post Occlusion Forearm vascular resistance (U) 36.00 ± 14.87 29.44 ± 10.26 1.2 0.22 Rest 8.40 ± 4.50 4.50 ± 1.34 2.69 0.012 Post Occlusion Forearm venous capacitance (ml•100ml tissue-1 •min-1) 1.90 ± 0.36 2.25 ± 0.13 -2.20 0.031 Rest 1.40 ± 0.54 1.88 ± 0.40 -3.43 0.022 Post Occlusion Forearm venous outflow (ml•100ml tissue-1 •min-1) 24.30 ± 8.00 27.25 ± 4.40 -1.10 0.3 Rest 29.50 ± 11.70 42.30 ± 10.37 -2.80 0.009 Post Occlusion groups. However, forearm arterial inflow, forearm venous capacitance, and forearm venous outflow after occlusion were significantly lower in the patients with HF.

Additionally, forearm vascular resistance was significantly higher in the patients with

HF.

3.3.4. Relationship Between Forearm Vascular Indices and Exercise Tolerance Correlation analysis revealed significant associations between all indices of vasoreactivity and exercise tolerance (forearm arterial inflow: r = 0.58, P = .001; forearm vascular resistance: r = -0.52, P = .004). Lastly, as shown in Figure 1, indices of venous function after occlusion were significantly related to maximum walking distance (forearm venous capacitance: r = 0.57, P = .001). Forearm venous outflow is shown in Figure 2 (r

= 0.67, P = .0001).

37 3.4. DISCUSSION

The unique finding of this study is the association between forearm venous capacitance

and venous outflow and the maximum walking distance achieved on a 6-minute walk

test. These data suggest factors that contribute to exercise impairment in patients with HF

may extend to the venous system. Given the important contributions of the venous

circulation to metabolic waste removal and end-diastolic volume, changes in this segment

of the vasculature could have a profound impact on the exercise response.

800 700 r =0.57; p =001 600 500 400

MWD (m) 300 200 100

.25 .75 1.25 1.75 2.25 2.75 Vcap (ml·100ml tissue-1·min-1)

Figure 3.1. Relationship between forearm venous capacitance and maximum walking distance.

As expected, the maximum walking distance achieved on the 6-minute walking test was significantly lower in the HF patients, suggesting compromised exercise tolerance.

The reduced maximum walking distance reported in the current study is consistent with previous reports (18, 23, 30) and have been attributed to pathophysiological abnormalities

that accompany the disease (28).

38 800 700 r =0.67; p =0001 600 500 400

MWD (m) 300 200 100

10 15 20 25 30 35 40 45 50 55 60 Vout(ml·100ml tissue-1·min-1)

Figure 3.2. Relationship between forearm venous outflow and maximum walking distance.

The use of the 6-minute walking test to examine exercise tolerance in patients with HF has received a lot of attention recently, due to the relative ease of its administration. The test is also thought to reflect a realistic effort as performed in daily life, and highly expectable by patients (9). Furthermore, the test appears to have adequate construct,

criterion, and predicted validity, as well as have a high degree of reproduceability.16

Construct validity has been shown by several investigators as the six-minute walk test

appears to discriminate between various populations (8, 15, 16, 23). Criterion validity

stems from studies showing adequate associations between exercise capacity (as defined

by VO2peak) and maximum walking distance on the six-minute walk test (11, 26). Yet, as

these findings are not consistent from one study to another, further work is needed to

verify the validity of the test. Finally, several studies have reported the prognostic

significance 6-min walking test in patients with HF indicating its predictive validity (16,

19).

39 A number of studies have reported adequate test, retest reliability for the maximum

walking distance achieved on the 6-min walking test (9, 11, 21). In the present study the

average distance walked increased by 5% from the first to the second test. However, the

interclass correlation coefficient of 0.97 and the CV of 8.9% indicate adequate stability of

maximum walking distance over a one week period. These data support research by

Guyatt et al, who recommend 2 walks be performed, and if the difference between the

first and second walk is greater than 10% a third walk should be undertaken and used as

the baseline measure (9).

The reduction in forearm arterial inflow and elevation in vascular resistance observed

in the current study are similar to previous reports (31, 33). Although the precise

mechanisms for the alterations in arterial vasoreactivity is not entirely understood, proposed mechanism for the reduction in peripheral blood flow include cardiac

insufficiency, neurohuromal hyper- (e.g. sympathetic nervous system, angiotensin and endothelin mediated vasoconstriction) and hypo-activity (e.g. endothelium- derived- relaxing factors), and vascular remodeling. In addition, decreased skeletal muscle mass,

alterations in fiber type composition and metabolic milieu further modify the regional

distribution of blood flow.

Previous investigations have related diminished venous function to respiratory-

cardiovascular hemodynamics, plasma norepinephrine (14) and severity of disease in

patients with HF (14, 27). Data from the current study confirm differences in venous

function following forearm occlusion between HF patients aged-matched controls. The unique finding of the current study is the demonstration that venous function are

40 associated with a measure of exercise tolerance, suggesting factors that contribute to

exercise impairment in chronic HF patients may extend to the venous system.

Examination of the literature regarding the reproducibility and validity of strain gauge

plethysmography to investigate arterial and venous function is difficult due to variations

in study aims, protocol, equipment, and interpretation of results. However in general the

literature appears to indicate adequate reproducibility for measures of arterial

vasoreactivity. In fact, unpublished data from our laboratory for within and between day

reliability indicate interclass correlation coefficients between 0.82 and 0.95 for forearm

arterial inflow and vascular resistance at rest and following forearm occlusion. The

stability of the venous function measurements is currently not clear. However, strain

gauge plethysmography is routinely used to detect venous dysfunction (12, 22, 25). The

technique appears to have adequate construct validity as it is able to differentiate between

patients with venous insufficiency and age-matched controls (7), and between patients of

different age groups (13). Moreover, high sensitivity, specificity, and predictive values were reported when strain gauged plethysmography technique was tested against

venography, the “gold dash standard”, to detect deep venous thrombosis (5). Preliminary

data from our laboratory revealed less than 7% difference for within and between day

measures at rest and following forearm occlusion.

Given the important contributions of the venous circulation to metabolic waste

removal and end-diastolic volume, changes in this segment of the vasculature could have

a profound impact on the exercise response. Evidence indicates that many HF patients are hyperadrenergic (34) and have elevated levels of endothelin-1 (17) and angiotensin II

(6) suggesting a possible neurohumoral influence on venous function (20). Considering

41 the cardiovascular system is a closed circuit, we speculate cardiac insufficiency,

decreased arterial reactivity, reduced muscle blood flow, venous alterations and elevated

cardiac pressures may contribute to a reduction in the driving pressure across metabolic

tissue. Consequently, decreased driving pressure could effect metabolic waste removal

(2), alter cardiac filling kinetics (10) and decrease exercise tolerance.

3.5. CONCLUSION

This study shows both arterial and venous indices are altered in HF patients and are associated with impaired exercise tolerance. Given the important contributions of the venous circulation to metabolic waste removal and end-diastolic volume, and the association of venous function with exercise tolerance among patients with HF demonstrated herein, future investigations should address venous health in the management of HF.

3.6. REFERENCES

1.Allen JD, Welsch M, Aucoin N, Wood R, Lee M and LeBlanc KE. Forearm vasoreactivity in type 1diabetic subjects. Can J Appl Physiol 26: 34-43, 2001.

2.Barclay JK. A delivery-independent blood flow effect on skeletal muscle fatigue. J Appl Physiol 61: 1084-1090, 1986.

3.Clark AL, Poole-Wilson PA and Coats AJ. Exercise limitation in chronic heart failure: central role of the periphery. J Am Coll Cardiol 28: 1092-1102, 1996.

4.Cramer M, Langlois Y, Beach K, Martin D and Strandness DEJ. Standardization of venous outflow measurements by strain gauge plethysmography in normal subjects. Bruit VII: 33-39, 1983.

5.Croal S, Birkmyre J, McNally M, Hamilton C and Mollan R. Strain gauge plethysmography for the detection of deep venous thrombosis. J Biomed Eng 15: 135- 139, 1993.

6.Curtiss C, Cohn JN, Vrobel T and Franciosa JA. Role of the -angiotensin system in the systemic vasoconstriction of chronic congestive heart failure. Circulation 58: 763- 770, 1978.

42

7.Doko S, Katsumura T, Fujiwara T, Nogami A and Masaki H. Maximum venous outflow and development of deep thrombosis. Jpn Circ J 55: 185-193, 1991.

8.Fitts SS and Guthrie MR. Six-minute walk by people with chronic renal failure. Assessment of effort by perceived exertion. Am J Phys Med Rehabil 74: 54-58, 1995.

9.Guyatt GH, Sullivan MJ, Thompson PJ, Fallen EL, Pugsley SO, Taylor DW and Berman LB. The 6-minute walk: a new measure of exercise capacity in patients with chronic heart failure. Can Med Assoc J 132: 919-923, 1985.

10.Guyton AC, Richardson TQ and Langston JB. Regulation of cardiac output and venous return. Clin Anesth 3: 1-34, 1964.

11.Hamilton DM and Haennel RG. Validity and reliability of the 6-minute walk test in a cardiac rehabilitation population. J Cardiopulm Rehabil 20: 156-164, 2000.

12.Hirai M, Yoshinaga M and Nakayama R. Assessment of venous insufficiency using photoplethysmography: a comparison to strain gauge plethysmography. Angiology 36: 795-801, 1985.

13.Hohn L, Chiarenza S, Schmied E and Bounameaux H. Age-related rather than ulcer- related impairment of venous function tests in patients with venous ulceration. Dermatologica 180: 73-75, 1990.

14.Ikenouchi H, Iizuka M, Sato H, Momomura S, Serizawa T and Sugimoto T. Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Jpn Heart J 32: 17-34, 1991.

15.King S, Wessel J, Bhambhani Y, Maikala R, Sholter D and Maksymowych W. Validity and reliability of the 6 minute walk in persons with fibromyalgia. J Rheumatol 26: 2233-2237, 1999.

16.Lipkin DP, Scriven AJ, Crake T and Poole-Wilson PA. Six minute walking test for assessing exercise capacity in chronic heart failure. Br Med J (Clin Res Ed) 292: 653- 655, 1986.

17.McMurray JJ, Ray SG, Abdullah I, Dargie HJ and Morton JJ. Plasma endothelin in chronic heart failure. Circulation 85: 1374-1379, 1992.

18.Meyer K, Gornandt L, Schwaibold M, Westbrook S, Hajric R, Peters K, Beneke R, Schnellbacher K and Roskamm H. Predictors of response to exercise training in severe chronic congestive heart failure. Am J Cardiol 80: 56-60, 1997.

19.Milligan NP, Havey J and Dossa A. Using a 6-minute walk test to predict outcomes in patients with left ventricular dysfunction. Rehabil Nurs 22: 177-181, 1997.

43

20.Monos E, Berczi V and Nadasy G. Local control of veins: biomechanical, metabolic, and humoral aspects. Physiol Rev 75: 611-666, 1995.

21.O'Keeffe ST, Lye M, Donnellan C and Carmichael DN. Reproducibility and responsiveness of quality of life assessment and six minute walk test in elderly heart failure patients. Heart 80: 377-382, 1998.

22.Olsen H and Lanne T. Reduced venous compliance in lower limbs of aging humans and its importance for capacitance function. Am J Physiol 275: H878-886, 1998.

23.Peeters P and Mets T. The 6-minute walk as an appropriate exercise test in elderly patients with chronic heart failure. J Gerontol A Biol Sci Med Sci 51: M147-151, 1996.

24.Poole-Wilson PA, Buller NP and Lipkin DP. Regional blood flow, muscle strength and skeletal muscle histology in severe congestive heart failure. Am J Cardiol 62: 49E- 52E, 1988.

25.Rooke TW, Heser JL and Osmundson PJ. Exercise strain-gauge venous plethysmography: evaluation of a "new" device for assessing lower limb venous incompetence. Angiology 43: 219-228, 1992.

26.Roul G, Germain P and Bareiss P. Does the 6-minute walk test predict the prognosis in patients with NYHA class II or III chronic heart failure? Am Heart J 136: 449-457, 1998.

27.Sato H, Ikenouchi H, Aoyagi T, Matsui H, Mochizuki T, Momomura S, Serizawa T, Iizuka M and Sugimoto T. Pathophysiology and evaluation of severity of congestive heart failure on the basis of venous characteristics. Jpn Circ J 53: 141-145, 1989.

28.Sullivan MJ and Hawthorne MH. Exercise intolerance in patients with chronic heart failure. Prog Cardiovasc Dis 38: 1-22, 1995.

29.Tretbar LL. Chronic venous insufficiency of the legs: pathogenesis of venous ulcers. J Enterostomal Ther 14: 105-108, 1987.

30.Troosters T, Gosselink R and Decramer M. Six minute walking distance in healthy elderly subjects. Eur Respir J 14: 270-274, 1999.

31.Zelis R and Flaim SF. Alterations in vasomotor tone in congestive heart failure. Prog Cardiovasc Dis 24: 437-459, 1982.

32.Zelis R, Longhurst J, Capone RJ and Mason DT. A comparison of regional blood flow and oxygen utilization during dynamic forearm exercise in normal subjects and patients with congestive heart failure. Circulation 50: 137-143, 1974.

44 33.Zelis R, Mason DT and Braunwald E. A comparison of the effects of vasodilator stimuli on peripheral resistance vessels in normal subjects and in patients with congestive heart failure. J Clin Invest 47: 960-970, 1968.

34.Zucker IH, Wang W, Brandle M, Schultz HD and Patel KP. Neural regulation of sympathetic nerve activity in heart failure. Prog Cardiovasc Dis 37: 397-414, 1995.

45 CHAPTER 4. MODIFICATION OF FOREARM VASCULAR FUNCTION FOLLOWING SHORT-TERM HANDGRIP EXERCISE TRAINING*

*Reprinted by permission of Georg Thieme Verlag Stuttgart Publisher (See appendix C)

46 4.1. INTRODUCTION

Endurance exercise training promotes adaptations of vascular control mechanisms that contribute to an increase in vasodilatory capacity thereby facilitating the delivery of blood to exercising muscle. The precise mechanism for vascular adaptations following exercise training are not fully understood, however, changes in the responsiveness of vascular endothelial and/or smooth muscle may, in part, enhance the vasodilatory capacity. For example, Delp & Laughlin (3) reported enhanced endothelium-dependent dilation of the rat aorta by the fourth week of endurance training and concluded that the adaptation was in part mediated through the nitric oxide (NO) synthase pathway.

Interestingly, the enhanced forearm peak vasodilatory capacity observed following four weeks of handgrip (HG) exercise training occurred without changes in the activity of the

NO system as reported by Green et al. (4). This suggests that the exercise stimulus to induce changes in vasoresponsiveness may, in part, be dependent on the exercise training modality. Indeed the above-mentioned studies appear to indicate that exercise modality has a differential effect on vascular adaptations (2, 3). More recently, Hepple et al. reported a dissociation of peak vascular conductance and VO2max in body builders and

endurance runners compared to healthy sedentary control, suggesting training-specific

effects of the cardiovascular system to local skeletal muscle metabolic demand (6).

An alternate explanation for the difference in vasoresponsiveness between

training protocols is that these adaptations are, in part, intensity-dependent. Interestingly,

vascular adaptations following small muscle group exercise training programs using different intensities have not been adequately explored. We are aware of only one study by Yasuda et al. (15) where subjects performed 6-weeks of HG exercise training at

47 intensities of 33 or 50% of maximum voluntary contraction (MVC). Peak forearm blood

flow adaptations increased significantly, but similarly in both groups. However, the

difference between training intensities was only 17% and may have been insufficient to

determine whether the adaptations are indeed intensity-dependent. Accordingly, the

purpose of this study was to determine the effects of low (LO) and high (HI) intensity short-term HG exercise training programs on localized vascular function in healthy young men. It was hypothesized that a training exercise intensity of 75% of MVC would promote greater vasodilatory capacity compared to a training exercise intensity of 25% of

MVC.

4.2. METHODS

4.2.1. Experimental Groups and Screening Measures Twenty-eight males (Age: 23±4.3yrs) were recruited and randomly assigned to either a

LO (n=15) or HI training group (n=13). Smokers and individuals with acute medical

conditions (e.g. orthopedic injury), active infection and/or on pharmacotherapy with

known vascular effects (e.g. anti-inflammatory therapy, cardiovascular medications) were

excluded. Following comprehensive explanation of the proposed study, its benefits,

inherent risks and expected commitments with regard to time, all participants signed an

informed consent approved by the Institutional Review Board of Louisiana State

University. In addition, each subject filled out a 24-hour medical and activity history questionnaire.

4.2.2. Experimental Design The study was a randomized, prospective design consisting of two experimental sessions

separated by a 4-week intervention period. The experimental sessions involved the

measurement of forearm blood flow (FBF) at rest (RBF), following 5 minutes of forearm

48 occlusion (OCC) and OCC combined with 3 minutes of HG exercise (ROCC). The

subjects exercised 5 days per week for 4 consecutive weeks at the exercise physiology

laboratory of Louisiana State University. Each training session consisted of 20 minutes of non-dominant HG exercise at a cadence of one contraction every 4 seconds.

Participants in the LO group trained at 25% of MVC, participants in the HI group trained

at 75% of MVC.

4.2.3. Experimental Measurements Upon arrival to the laboratory, the subject’s height and weight were obtained. In addition,

MVC was determined for both arms as the average of three maximum consecutive

contraction trials using a Stoelting hand-dynamometer. Thereafter, each subject was placed in a supine position for 20 minutes prior to evaluation of forearm vascular blood flow. Forearm blood flow measurements were obtained from both arms (the non- dominant first) using an EC-5R plethysmography system (Hokanson, 1995). Prior to each experiment, blood pressure cuffs were positioned around the participant’s upper arm and wrist, and a mercury-in-silastic strain gauge (7) placed around the forearm approximately

10 cm distal to the olecranon process (13). The forearm was extended and slightly supinated and supported by a Styrofoam block. Immediately before the blood flow measurements, hand circulation was occluded for 1 min by inflating the cuff at the wrist to 240mmHg (10). All blood flow measures were made with an upper arm venous collecting pressure of 50mmHg (11) and reported in ml per min per 100 ml of tissue (13).

Reactive hyperemia following OCC was achieved by inflating the cuff on the upper arm to 240mmHg (13). Blood flow measures were then determined as described above.

Finally, vasoresponsiveness following ROCC was obtained. Handgrip exercise for the

ROCC condition was performed at an intensity of 25% MVC and a cadence of 1

49 contraction per 4 sec during the 2nd, 3rd, and 4th minutes of occlusion. ROCC

measurements were obtained following OCC and were separated by 15 minutes.

Throughout the testing procedures, heart rate and blood pressure were obtained at the

following time points: rest, during and immediately after OCC and ROCC. Throughout

the testing procedures, which lasted approximately 120 minutes the subject remained in

the supine position.

4.2.4. Data Analysis Resting blood flow was recorded at a paper speed of 5 mm/sec. and values were derived

from the slope drawn at a best-fit tangent using the first 3 pulses. Calculations were made

as a function of 60 seconds divided by the horizontal distance (mm) needed for the slope

to rise vertically from baseline to the top of the recording paper and multiplied by the full

chart range. Blood flow responses following OCC and ROCC were recorded at a paper

speed of 25 cm/sec. Analyses were performed using a slope drawn at a best-fit tangent to the curves of the first 3-pulse flows post cuff release. The blood flows were then calculated from 60 sec. multiplied by the paper speed (25 cm/sec) divided by the horizontal distance (mm) needed for the volume slope to increase by 20 mm vertically.

Mean arterial pressure was calculated from the difference between SBP and DBP divided by three and added to DBP {MAP = DBP+[(SBP-DBP)/3]}. Forearm vascular resistance was calculated by dividing mean arterial pressure by the forearm blood flow (FVR =

MAP/FBF).

50 4.2.5. Statistical Analysis

Grip strength Dependent t-tests were used to compare the grip strengths of dominant and non-dominant

arms observed prior to and following the training period. Separate t-tests were conducted

for the LO and HI intensity training groups.

Comparison of dominant to non-dominant arms Two 3x2 ANOVA’s were used to compare the FBF and FVR during each condition (rest vs. OCC vs. ROCC) in the dominant and non-dominant arms. The pre-training and post- training data were treated separately.

Effect of training on vascular responses. A 3x2x2 ANOVA was used to examine the effects of the training period on FBF and

FVR as measured under the three conditions (rest vs. OCC vs. ROCC), and across the

two training groups (LO vs. HI intensity).

Alpha was set a-priori at p<0.05 for all statistical analyses. The LSD method of post-hoc

testing was employed where indicated.

4.3. RESULTS

4.3.1. Participant Characteristics The initial number of subjects was thirty, however one participant was excluded because

of noncompliance and one participant for violation of exclusion criteria (initiated

smoking). As shown in Table 1 there were no statistically significant differences between

the LO and the HI groups in the pre-training measurements for age, height, weight,

resting heart rate, and blood pressure.

51 Table 4.1. Subject characteristics LO HI (n=15) (n=13) Age (yrs) 22 ± 2.5 24 ± 5.8 Weight (kg) 86 ± 7.5 70 ± 5.4 Height (cm) 175 ± 15.4 174 ± 10

SBPrest (mmHg) 106.5 ± 18.78 106.9 ± 9

DBPrest (mmHg) 68.3 ± 12.3 68.3 ± 7.4

MAPrest (mmHg) 81 ± 10 81.5 ± 6.3 -1 HRrest (bts•min ) 62.1 ± 6.3 64.5 ± 7 Values are mean ± SD. LO, low training intensity = 25% of maximum voluntary contraction (MVC); HI, high training intensity = 75% of MVC; SBPrest, systolic blood pressure at rest; DBPrest, diastolic blood pressure at rest; MAPrest, mean arterial blood pressure at rest; HRrest, heart rate at rest.

4.3.2. Grip Strength All participants completed at least 18 sessions of HG exercise training over the 4-week

period. The result of the t-test revealed no change in maximal grip strength in the trained

and non-trained arms of both groups. However, there was trend towards an improvement

in the trained arm strength in the HI group (35.3±8.8 to 38.3±8.8 kg; 8%; p=0.095).

4.3.3. Comparison of Dominant to Non-Dominant Arm As could be depicted from figure 1 and 2 and table 2, the results of the ANOVA on the

pre-training values of FBF revealed an interaction for test condition (p<0.0001) and arm

(p=0.002). This interaction was such that OCC and ROCC resulted in a graded FBF

response, with the greatest hyperemic FBF occurring following ROCC. Furthermore FBF

following OCC and ROCC were greater in the dominant arm compared to the non-

dominant arm. There was also a main effect of test condition (p<0.0001) on FVR. FVR

was lower following OCC and ROCC compared to rest, but not different from each other.

However, FVR in the dominant arm was not different from that of the non-dominant arm.

52 ‡ † 45 § * § § 40 * § * ‡ † * 35 * * * 30 *

25

20 Pre Post 15

10

Forearm Blood Flow (ml.100ml-1.min-1) 5

0 Rest OCC ROCC Rest OCC ROCC Dominant Arm Non-dominant Arm

Figure 4.1. FBF changes in the LO group following HG exercise training. Values are means ± SD. *p<0.05 vs. rest; §p<0.05 vs. OCC ‡p<0.05 vs. Pre Non-dominant; †p<0.05 vs.

The results of the ANOVA on the post-training FBF values revealed a main effect of test

condition (p<0.0001) but not for arm (p=0.718). This main effect was such that the OCC

and ROCC once again resulted in a graded response. However, following training the

response of the non-dominant arm was not different from that of the dominant arm. FVR

responses following OCC and ROCC were similar to pre-training values for both arms.

4.3.3. Effect of Training on Vascular Responses. The results of the ANOVA on the effects of training on FBF in the non-dominant arm revealed a main effect for training period (p=0.005). However, the ANOVA did not

reveal a group by training interaction (p=0.632). The main effect was such that FBF values following OCC and ROCC were higher after the training period as compared to pre-training values. However, there was no interaction with training intensity, suggesting

53 that both training groups responded similarly. The results of the ANOVA on FVR

revealed only a main effect of the test condition (i.e. rest, OCC, ROCC). These main

effects are the same as described above, and did not change significantly as a result of the

training period in either the LO or HI intensity group.

† 45 ‡ § § § § * 40 ‡ * * † * 35 * * * 30 *

25

20 Pre 15 Post 10 Forearm Blood Flow (ml.100ml-1.min-1) 5

0 Rest OCC ROCC Rest OCC ROCC Dominant Arm Non-dominant Arm

Figure 4.2. FBF changes in the HI group following HG exercise training. Values are means ± SD. *p<0.05 vs. rest; §p<0.05 vs. OCC ‡p<0.05 vs. Pre Non-dominant; †p<0.05 vs. Pre

4.4. DISCUSSION

This study examined the role of a short-term HG exercise program on FBF and FVR. Our findings indicate that the degree of vascular responsiveness is directly related to the intensity of the vasodilatory stimulus, given that the blood flow measures were significantly different for all 3 conditions (ROCC > OCC > RBF). Furthermore, vasodilatory responsiveness prior to 4 weeks of HG exercise was significantly greater in the dominant compared to the non-dominant arm. Finally, following 4 weeks of HG exercise

54 performed 5 days per week for 20 minutes vasodilatory responsiveness improved unilaterally but similarly for the LO and HI intensity groups.

Table 4.2. Dominant and non-dominant FVR changes in the LO and HI groups at rest, following OCC and following ROCC. Dominant Non-dominant Pre Post Pre Post

LO group (15) FVRrest 47.85 ± 13 42.2 ± 13.4 42.5 ± 15.4 44.6 ±17.0 FVROCC 3.3 ± 1* 3.5 ± 0.5* 3.5 ± 0.6* 3.5 ± 0.8* FVRROCC 2.5 ± 0.5* 2.8 ± 0.4* 3 ± 0.4* 2.9 ± 0.7* HI group (13) FVRrest 44.5 ± 19 45.7 ± 22.7 39.4 ± 15 34.4 ± 8.8 FVROCC 3.1 ± 0.5∗ 3.1 ± 0.5∗ 3.7 ± 0.6∗ 3.3 ±0.7∗ FVRROCC 2.7 ± 0.3∗ 2.77 ± 0.4∗ 3.3 ± 0.7∗ 2.5 ± 0.7∗ Values are means ± SD. LO, training intensity = 25% of maximum voluntary contraction (MVC); HI, training intensity = 75% of MVC; FVRrest, forearm vascular resistance at rest; FVROCC, forearm vascular resistance following five minute of occlusion; FVRROCC, forearm vascular resistance following OCC coupled with 3-min. of HG of exercise. ∗p < 0.05 vs. rest.

Previous studies have reported improved unilateral vasodilatory responsiveness following small muscle group conditioning programs (see Table 3) (4, 9, 14). The improvements in peak FBF following OCC in this study were similar for the HI group

(20.6%) to those reported by Green et al. (23%) but lower than Sinoway et al. (30%). The discrepancy between our findings and Sinoway et al. could be the result of a difference in total minutes of exercise performed or methods of assessing forearm blood flow. For example, the training regime used by Sinoway et al. included four, 30-minute sessions per week (120 minutes per week) versus our five, 20-minute sessions per week (100 minutes per week) for 4 weeks. In addition, Sinoway et al. evaluated the effect of short- term HG exercise on vasodilatory responsiveness after 10 minutes of arterial occlusion

(14). Lastly, the supine rest period prior to obtaining the resting blood flow varied between studies.

55

Table 4.3. Literature overview Study Population Type of training Results

Yasuda, 83 10 healthy, HG, 33 & 50% MVC, 6X/wk (6 34 & 26% ⇑ BF, respectively. young wks)

Sinoway, 87 6 healthy, HG, 70% MVC, 4X/wk, 30min ⇔ BFrest. young (4 wks) 30% ⇑ RHBF in trained arm

Green, 94 11 healthy, HG, 75% MVC, 4X/wk, 30min ⇔ BFrest young (4 wks) 23% ⇑ RHBF in trained arm

Bond, 94 7 healthy, Plantar Flexion, 75% MVC, ⇔ BFrest. young 4X/wk, 6sets of 15reps (4 wks) 22% ⇓ in RHBF, ⇑ VR.

Franke, 98 6 healthy, HG, 70% MVC, (4 wks) ⇔FVCrest. young 35% ⇑ FVCpeak. Bank, 98 11 healthy, HG, 70% MVC, 4X/wk, 30min, ⇔ in CHF. young & 7 CHF (4-6 wks) 24% ⇑ RHBF in healthy, young. BF = Blood flow; RHBF = Reactive hyperemic blood flow; VR = Vascular resistance; FVC = Forearm vascular conductance.

Our findings stand in contrast to other studies of different training duration or muscle groups (1, 12, 15). For example, Yasuda et al. reported an increase in FBF in the ipsi- and contralateral arm following 6 weeks of HG exercise training (15) and attributed the improvements to changes in neurogenic balance (15). In contrast, other investigators report a reduction in limb blood flow following short-term training programs. Bond et al.

(1) attributed reductions in leg blood flow to an increase in muscle cross-sectional area without proportional changes in microvascular density, whereas Saito et al. attributed the lower leg blood flow to a more effective metabolic process (12). Clearly, more studies are needed to further understand the vascular adaptive responses to exercise training.

Similar to the study by Yasuda et al. (15) the current study found no statistical differences in the degree of change in vasoresponsiveness between the LO and HI groups following the training period. Subsequent to training, FBF increased 16.5%, and 17.7% in the LO group, and 20.6%, and 29.2% in the HI group after OCC, and ROCC, respectively. The lack of a significant difference between the groups suggests that the

56 signal for vascular adaptations may have been similar for both groups. Moreover, it indicates that significant local vascular adaptations can be induced with small muscle group conditioning programs at intensities as low as 25% of MVC. Such adaptations may be particularly valuable in populations for whom low-intensity exercise training is indicated. In fact, recent studies have reported short-term HG training results in enhanced endothelium-dependent vasodilation in the forearm skeletal muscle circulation of patients with chronic heart failure (9). Likewise, patients with other disease conditions (e.g. diabetes mellitus, peripheral vascular disease, hypertension and coronary artery disease) who suffer from compromised vascular function might benefit from such small muscle group conditioning regimes. Clearly, further investigations are needed to verify this theory and whether such training programs can ultimately contribute to an individual’s improved exercise performance.

Another finding of this study is the observed difference in vasodilatory capacity between the dominant and non-dominant arm. This data is consistent with Sinoway et al. who reported a 25% and 29% difference following OCC and ROCC respectively in the dominant arm versus non-dominant arm of tennis players (13). Likewise, in the current study, vascular capacity in the dominant arm was greater than the non-dominant arm

(10% and 12% difference following OCC and ROCC respectively) perhaps as a consequence of preferential usage. The difference between Sinoway’s findings and ours is most likely the result of regular participation in the sport of tennis, as all the subjects in

Sinoway’s study were considered advanced players (13). Significantly, the observed differences in FBF between the dominant and the non-dominant arm prior to training in this study disappeared after training. This indicates that forearm vascular adaptations in

57 the dominant arm were matched by short-term forearm HG exercise training in the non-

dominant arm.

Although this study is not adequately or purposely designed to explain the exact

mechanism of the training adaptations, several previous investigations have indicated the

role of the endothelium. Shear stress associated with exercise has been proposed to

contribute to the increased production of endothelial derived relaxing factors, such as NO

(8). It is postulated that the contraction and relaxation phases of exercise results in brief

circulatory arrest followed by a hyperemic response (3, 5). This intermittent blood flow

might create shear stress on the vascular wall that is responsible for the vascular

adaptations demonstrated in this study. Clearly, other factors (e.g. increased muscle

capillarizations, or changes in local metabolites, neuroendocrine balance, other myogenic

factors and/or cardiac function) associated with exercise might also contribute to improved vascular function (3). The fact vasoreactivity is a complex interplay of central, vascular and local control mechanisms limits this study’s ability to precisely identify the mechanism(s) for change.

In conclusion, FBF and FVR at rest and following forearm occlusion (with and without HG exercise) showed a graded response with the greatest hyperemic flow recorded after five minutes of forearm occlusion with HG contraction. Furthermore, the non-dominant FBF was found to improve in response to low-intensity as well as high- intensity short-term HG exercise training. Furthermore, blood flow in the non-dominant forearm following training was increased to match the flow in the dominant forearm. The exact mechanism(s) behind these vascular adaptations are not known and warrants further investigations.

58 4.5. REFERENCES

1.Bond V, Jr., Wang P, Adams RG, Johnson AT, Vaccaro P, Tearney RJ, Millis RM, Franks BD and Bassett DR, Jr. Lower leg high-intensity resistance training and peripheral hemodynamic adaptations. Can J Appl Physiol 21: 209-217, 1996.

2.Copeland SR, Mills MC, Lerner JL, Crizer MF, Thompson CW and Sullivan JM. Hemodynamic effects of aerobic vs resistance exercise. J Hum Hypertens 10: 747-753., 1996.

3.Delp MD and Laughlin MH. Time course of enhanced endothelium-mediated dilation in aorta of trained rats. Med Sci Sports Exerc 29: 1454-1461, 1997.

4.Green DJ, Cable NT, Fox C, Rankin JM and Taylor RR. Modification of forearm resistance vessels by exercise training in young men. J Appl Physiol 77: 1829-1833, 1994.

5.Green DJ, O'Driscoll JG, Blanksby BA and Taylor RR. Effect of casting on forearm resistance vessels in young men. Med Sci Sports Exerc 29: 1325-1331, 1997.

6.Hepple RT, Babits TL, Plyley MJ and Goodman JM. Dissociation of peak vascular conductance and V(O2) max among highly trained athletes. J Appl Physiol 87: 1368- 1372, 1999.

7.Holling HE, Boland HC and Russ E. Investigation of arterial obstruction using a mercury-in-rubber strain gauge. Am Heart J 63: 194-205, 1961.

8.Hornig B, Maier V and Drexler H. Physical training improves endothelial function in patients with chronic heart failure. Circulation 93: 210-214, 1996.

9.Katz SD, Yuen J, Bijou R and LeJemtel TH. Training improves endothelium-dependent vasodilation in resistance vessels of patients with heart failure. J Appl Physiol 82: 1488- 1492, 1997.

10.Kerslake DM. The effect of the application of an arterial occlusion cuff to the wrist on the blood flow in the human forearm. J Physiol 108: 451-461, 1949.

11.Patterson GC and Whelan RF. Reactive hyperemia in the human forearm. Clin Sci Lond 14: 197-211, 1955.

12.Saito M, Matsui H and Miyamura M. Effects of physical training on the calf and thigh blood flows. Jpn J Physiol 30: 955-959, 1980.

13.Sinoway LI, Musch TI, Minotti JR and Zelis R. Enhanced maximal metabolic vasodilatation in the dominant forearms of tennis players. Journal of Applied Physiology 61: 673-678, 1986.

59

14.Sinoway LI, Shenberger J, Wilson J, McLaughlin D, Musch T and Zelis R. A 30-day forearm work protocol increases maximal forearm blood flow. J Appl Physiol 62: 1063- 1067, 1987.

15.Yasuda Y and Miyamura M. Cross transfer effects of muscular training on blood flow in the ipsilateral and contralateral forearms. Eur J Appl Physiol Occup Physiol 51: 321- 329, 1983.

60 CHAPTER 5. ARTERIAL AND VENOUS ADAPTATIONS TO SHORT-TERM HANDGRIP EXERCISE TRAINING

61 5.1. INTRODUCTION

Structural and functional changes in the arterial system following exercise training are

well documented. These adaptations are associated with reduced vascular resistance,

improved blood delivery and diffusion in the contracting muscle and contribute to

improved vascular modulation and exercise performance during subsequent exercise sessions. The mechanism(s) by which these adaptations are induced are not fully understood, but are thought to include alterations in the arterial endothelium and/or

smooth muscle (4, 8, 12, 18, 32, 65). Recent studies report training-induced adaptations

may be observed following one week of training in animal models (35, 61), and in human

single large conduit arteries (i.e. brachial) (2).

Interestingly, the effect of exercise training on venous function has received

considerably less attention despite compelling evidence demonstrating its dynamic role

beyond a typical role as a “passive volume reservoir” (22, 72). Most of the available

studies have linked training-induced improvements in venous function to greater blood

volume associated with increased VO2 (10, 29, 41), as well as increased venular cross-

sectional area (49) and density (70). On the other hand, Wecht et al. attributed enhanced

venous function in “aerobically” trained individuals to improved venous hemodynamics,

including venous emptying, outflow and compliance (76). However no studies have

examined the effect of localized exercise training (i.e. handgrip) on regional venous

function.

Therefore, the purpose of this study was to examine the influence of 4-weeks of

handgrip exercise training on arterial and venous function. A secondary aim of this study

62 was to determine the time course of change for both arterial and venous indices associated with the handgrip exercise training program.

5.2. HYPOTHESES

This study tested two specific hypotheses regarding the influence of handgrip exercise on arterial (Hypothesis I) and veno-dynamics (Hypothesis II).

5.2.1. Hypothesis I Four-weeks of handgrip exercise training will result in an increase in arterial reactivity following 5-minutes of arterial occlusion;

5.2.2. Hypothesis II Four-weeks of handgrip exercise training will result in an increase in venous capacitance, compliance and outflow at rest and following 5-minutes of arterial occlusion.

5.3. METHODS

5.3.1. Participants and Design Apparently healthy sedentary young adult males were recruited to participate in the study. Individuals with manifestations of cardiovascular, metabolic, orthopedic or neurological disease or on any medication (e.g., digitalis) which could affect the results of this study were excluded.

The study was executed in 4 weeks, and involved once a week evaluation session and 5-time-a-week exercise training sessions. Prior to and at the end of each week, participants underwent handgrip strength and vascular function evaluations in the dominant and non-dominant forearms. The primary outcomes of this study include: (1)

Peak blood inflow, defined as vascular response to 5 minutes of upper arm arterial occlusion (64); (2) Venous capacitance; defined as the changes in limb blood volume following 6 minutes of venous occlusion; (3) Venous outflow; defined as the changes in

63 limb blood volume following the release of venous occlusion and; (4) Venous

compliance; defined as the ratio of change in volume to concomitant change in

transmural distending pressure (i.e. occlusion pressure) (11, 23, 54, 76). Additionally,

handgrip strength, arm exercise tolerance, and hemodynamic measures (i.e. blood

pressure, heart rate and heart rate variability) were also assessed. Following the

completion of the initial assessment, each subject was asked to report to the exercise

physiology laboratory 5 days/week to perform 20 minutes of forearm flexion exercise at a cadence of 1 contraction every 4 seconds starting at an intensity of 60% of maximum

handgrip strength. The non-dominant arm was trained while the dominant arm served as control.

5.3.2. Study Measurements:

Vascular function assessments Vascular function indices were obtained in the dominant and non-dominant forearms

using mercury strain gauge plethysmography at rest and following occlusion (64) at the

beginning of the study, and at the end of each week throughout the study period. This

technique is noninvasive, and based on the assumption that alterations of pressures in

strategically placed cuffs allows examination of the rate of change of limb volume

thought to reflect venous function indices including blood inflow, vascular resistance,

venous capacitance, venous outflow and venous compliance (11, 23). Prior to the

experiment, blood pressure cuffs were positioned around the participant’s upper arm and

wrist, and a mercury-in-Silastic strain gauge placed around the forearm approximately 10

cm distal to the olecranon process (4). Resting forearm vascular function measures were

obtained following 15 minutes of supine rest. Immediately before the measurements, hand

circulation was occluded for 1 minute by inflating the cuff at the wrist to 240 mmHg.

64 Forearm blood inflow was obtained using an upper arm venous occluding pressure of 7 mmHg below diastolic blood pressure. Forearm venous capacitance was measured after

an additional 6 minutes of venous occlusion and venous outflow following the release of

pressure (3, 66). Subsequently, peak forearm blood was examined following arterial

occlusion achieved by inflating the cuff on the upper arm to 240 mmHg for 5 minutes.

Forearm vascular indices were then determined as described above. These were then

performed on the other arm. Prior to, during and following each procedure blood pressure and heart rate were obtained.

Autonomic nervous function assessments Indices of autonomic function were assessed using measures of heart rate variability.

These measures were obtained at the same time as the vascular assessments described

above. The ECG electrodes were placed on the chest and interfaced with a Biopac

MP100 and its companion software Acqknowledge (model MP100A, Biopac Inc., Santa

Barbara, CA) to allow for continuous data acquisition. Five minutes (at rest and during

forearm occlusion) of ECG data were analyzed for mean heart period (mean R-R

interval), standard deviation (SDNN), low frequency power (LF) and high frequency

power (HF) using the guidelines set forth by the Task Force of the European Society of

Cardiology and the North American Society of Pacing and Electrophysiology (7, 71).

These data were obtained at the beginning and at the end of the study period.

Arm ergometer test Arm exercise tolerance was evaluated using a graded arm ergometer exercise test at the

beginning and at the end of the study period (5, 48). The test was performed by initially

turning the arm ergometer at a speed of 50 rpm and a resistance of 0.5 kp for 2 minutes

65 with the participant in a seated position. The resistance was then increased by 0.25 kp

every 2 minutes until the subject reached exhaustion, requested to stop, or exhibited signs

or symptoms, which require test termination. Prior to, during, and following the exercise

test, heart rate, blood pressure measurements, and ratings of perceived exertion were also

obtained at the end of each minute throughout the test using a standard sphygmomanometer and the Borg's perceived exertion scale, respectively.

Forearm circumference and handgrip strength assessments The participant’s forearm circumferences and handgrip maximal voluntary contraction

were evaluated at baseline and at the end of each week throughout the study period.

Forearm circumference was examined using a weighted measuring tape 10 cm distal to

the midpoint between the lateral epicondyle and olecranon process. Handgrip strength

was evaluated using a handgrip dynamometer (Lafayette instruments, OH, 02) with the

participant upright, but slightly bent forward at the waist. The test involved an all-out

gripping effort for 3 seconds, without movement of the arm. The average of three

consecutive trials was used as the measure of strength.

Body composition Three-site skinfold method was used to estimate body percents of fat and fat-free masses.

The recommended body sites for this method were chest, abdominal and thigh (80).

Additionally dominant and non-dominant forearm skinfolds were obtained at the upper

1/3 of the forearm.

5.3.3. Data Reduction and Analysis Resting blood inflow was recorded at a paper speed of 5 mm/seconds and values were

derived from the slope drawn at a best-fit tangent using the first 2-3 pulses. Calculations

66 were made as a function of 60 seconds divided by the horizontal distance (mm) needed for the slope to rise vertically from baseline to the top of the recording paper and multiplied by the full chart range (11, 23). Peak forearm blood inflow was recorded at a paper speed of 25 mm/second. Analyses were performed using a slope drawn at a best-fit tangent to the curves of the first 3-pulse flows post cuff release. Blood flows were then calculated from 60 seconds multiplied by the paper speed, and divided by the horizontal distance (mm) needed for the volume slope to increase by 20 mm vertically (4). Forearm vascular resistance was then determined by dividing mean arterial pressure by the blood inflow. Forearm venous capacitance was calculated as the vertical distance (mm) representing the change in forearm-volume graph following 6 minutes of venous filling

(11). Forearm venous compliance was calculated as the ratio of venous capacitance and occlusion pressure multiplied by 100 (76). Forearm venous outflow was derived from a tangent line representing the vertical drop in the volume-graph from the excursion line and drawn at 0.5 second and 2 second following the release of the venous occlusion pressure (11). All vascular indices were expressed in ml•100ml tissue-1•min-1.

5.3.4. Statistical Analysis All statistical analyses were completed using the SPSS statistical program (version 11.00,

Chicago, Ill). To examine the influence of 4-weeks of handgrip exercise training on arterial reactivity (Hypothesis I) a 2 (Arms) * 5 (visit 1-vist 5) split-plot ANOVA was used. To examine the influence of 4-weeks of handgrip exercise training on venous capacitance, compliance and outflow at rest and following 5-minutes of arm occlusion

(Hypothesis II) a second 2 (Arms) * 5 (visit 1-vist 5) split-plot ANOVA was used.

Pearson product moment correlations were used to examine the relationships between

67 handgrip strength and vascular function indices, between arterial and venous function

indices as well as between arterial and venous adaptations. Finally a paired t-test was

used to compare between pre and post training measures of HRV indices, the skinfold

measures, and handgrip strength. Alpha was set a-priori at p<0.05.

5.4. RESULTS

5.4.1. Participants Twenty-two men participated in the study. Five participants did not complete the study

due to personal decision (3 individuals) or lack of compliance (2 individuals). Participant

descriptions are presented in Table 5.1.

Table 5.1. Subject characteristics Visit 1 Age (yrs) 22.6 ± 3.5 Height (cm) 175.4 ± 1.6 Weight (kg) 75.5 ± 4.0 Control arm length (cm) 27.5 ± 0.3 Trained arm length (cm) 27.4 ± 0.3 Sum of Skin fold (mm) 54.4 ± 7.7 Values are mean ± SE.

5.4.2. Hemodynamic, Anthropometric and Exercise Measures Table 5.2. Cardiovascular hemodynamics and HRV before and after 4-week handgrip exercise training Variable Visit 1 Visit 5 HR (bpm) 58.2 ±1.6 57.5 ± 1.8 SBP (mmHg) 112.6 ± 2.4 111.1 ± 2.1 DBP (mmHg) 60.7 ± 1.9 59.7 ± 1.8 MAP (mmHg) 78.0 ± 8.1 76.8 ±7.6 SDNN (ms) 73.4 ± 4.1 79.6 ± 4.0 LF power 36.9 ± 1.3 37.2 ± 1.4 HF power 63.1 ± 1.3 62.8 ± 1.4 Values are mean ± SE.

68 Hemodynamic, anthropometric and exercise measures at baseline (visit 1) and the end of the study (visit 5) are presented in Tables 5.2 and 5.3. No significant changes in heart rate, blood pressure, and heart rate variability indices were observed. Moreover, there were no changes in anthropometric measures in the non-trained arm. In contrast, forearm circumference increased by 1.56% (p=0.03) in the trained arm. Handgrip strength was not modified in the non-trained arm, whereas the trained arm showed a 14.5% (p=0.014) increase in maximal voluntary contraction. Finally, there were no changes in the time to fatigue for the symptom-limited graded arm ergometer exercise test.

Table 5.3. Arm ergometer exercise test, forearm skinfolds, and sum of three skinfolds before and after 4-week handgrip exercise training Visit 1 Visit 5 Arm ergometer test (mints.) 10.3 ± 0.6 9.75 ± 0.9 Control handgrip strength (kg) 40.6 ± 2.9 42.8 ± 3.2 Trained handgrip strength (kg) 36.1 ± 2.1 41.3 ± 1.7 * Control arm circumference (cm) 26.3 ± 0.5 26.5 ± 0.5 Trained arm circumference (cm) 25.6 ± 0.6 26.0 ± 0.5 * Control arm skin fold (mm) 1.1 ± 0.2 1.1 ± 0.13 Trained arm skin fold (mm) 1.2 ± 0.3 1.0 ± 0.13 Values are mean ± SE. *=p<0.05 vs. visit 1

5.4.3. Effect of Training on Vascular Function Values for the blood flow responses at rest and following occlusion for visits 1 and 5 are

presented in table 5.4 and 5.5. Furthermore, Figures 1 through 5 present data for the

vascular measures across the entire study.

69

Table 5.4. Resting arterial and venous indices Control arm Trained arm Visit 1 Visit 5 Visit 1 Visit 5 Arterial inflow 1 1 2.9 ± 0.2 2.6 ± 0.2 3.0 ± 0.2 3.0 ± 0.2 (ml·100ml tissue- ·min- ) Vascular resistance 29.4 ± 2.3 31.7 ± 2.3 28.1 ± 2.3 28.6 ± 2.3 (U) Venous capacitance 1 1 4.6 ± 0.13 4.5 ± 0.13 4.8 ± 0.13 4.5 ± 0.13 (ml·100ml tissue- ·min- ) Venous compliance 8.6 ± 0.32 9.1 ± 0.33 9.9 ± 0.32 9.2 ± 0.33 (%) Venous outflow 1 1 42.5 ± 1.9 43.3 ± 1.9 44.8 ± 1.8 45.4 ± 1.8 (ml·100ml tissue- ·min- ) Values are mean ± SE.

Table 5.5. Post occlusion arterial and venous indices Control arm Trained arm Visit 1 Visit 5 Visit 1 Visit 5 Arterial inflow 1 1 22.1 ± 0.7‡ 21.6 ± 0.7 19.0 ± 0.7 22.6 ± 0.7* (ml·100ml tissue- ·min- ) Vascular resistance 3.7 ± 0.2‡ 3.9 ± 0.2 4.3 ± 0.2 3.5 ± 2.3* (U) Venous capacitance 1 1 3.4 ± 0.13 3.4 ± 0.13 3.6 ± 0.13 3.5 ± 0.13 (ml·100ml tissue- ·min- ) Venous compliance 6.2 ± 0.25 6.2 ± 0.26 6.4 ± 0.24 6.3 ± 0.25 (%) Venous outflow 1 1 42.1 ± 1.7 44.9 ± 1.9 45.1 ± 1.7 45.5 ± 1.7 (ml·100ml tissue- ·min- ) Values are mean ± SE. *=p<0.05 vs. visit 1; ‡=p<0.05 vs. Trained arm.

Resting arterial function indices. The 2 X 5 split-plot ANOVA tests revealed no significant main effects for arms or visits, or for the arms by visits interactions for resting forearm arterial inflow and vascular resistance (see Table 5.4).

70 Control arm 25 Trained arm * ‡ * 1 * *

20 Arterial inflow (ml.100ml-1.min-

15 12345 Vis it

Figure 5.1. Changes in peak forearm blood inflow. Values are mean ± SE. *=p<0.05 vs. visit 1; ‡=p<0.05 vs. Trained arm.

Control arm 5 Trained arm

‡ 4 * * * * (U) 3 Vascular resistance

2 12345 Visit

Figure 5.2. Changes in peak forearm vascular resistance. Values are mean ± SE. *=p<0.05 vs. visit1; ‡=p<0.05 vs. Trained arm.

Peak arterial function indices. As depicted in Figures 5.1 and 5.2, the 2 X 5 split-plot ANOVA test revealed no

significant main effects for arms or visits for peak arterial indices, however there were

significant in the arms by visits interaction for trained forearm peak forearm arterial

inflow (p=0.02) and vascular resistance (p=0.009). Subsequent LSD post-hoc comparison

demonstrated increased peak arterial inflow, and decreased peak vascular resistance

71 following the 1st week in the trained forearm only. Finally, post-hoc comparison revealed greater peak forearm blood inflow (p=0.0013) and vascular resistance (p=0.05) in the

untrained verses the trained arm at visit 1, however, no differences were observed in the

subsequent visits.

Control arm 11 Trained arm † 10 † † ‡ * 9 (%) *

Venous compliance 8

7 12345 Vis it s

Figure 5.3. Changes in resting forearm venous compliance. Values are mean ± SE. *=p<0.05 vs. visit 1; ‡=p<0.05 vs. Trained arm; † =p<0.05 vs. visit 2.

Resting venous function indices. The 2 X 5 split-plot ANOVA tests revealed no significant main effects of arms or visits,

or arms by visits interactions for resting forearm venous capacitance and outflow (see

Table 4), however revealed significant (p=0.04) arms by visits interaction for venous

compliance. As depicted in Figure 5.3, subsequent LSD post-hoc comparison

demonstrated a decrease in trained arm venous compliance in visit 2 followed by a

gradual return to baseline by visit 5.

72 4

Control arm 3.8 ) Trained arm

3.6

3.4 (ml.100ml-1.min-1 Venous capacitance 3.2

3 12345 Vis it

Figure 5.4. Changes in post occlusion forearm venous capacitance. Values are mean ± SE.

Post-occlusion venous function indices The 2 X 5 split-plot ANOVA tests revealed no significant main effects for arms or visits,

or arms by visits interactions for post-occlusion forearm venous compliance and outflow.

However there was a significant main effect of visits (p=0.04) for venous outflow (see

Table 5.5 and Figure 5.4). As depicted in Figure 5.5, Subsequent LSD post-hoc

comparison demonstrated a decrease in trained forearm venous outflow in visit 2

followed by a gradual return to baseline by visit 5. The control forearm venous outflow

was greater at visit 2 than at visit 5.

Relationships between handgrip strength, and arterial and venous function indices The Pearson product moment correlations are presented in Table 5.6 and Figures 5.6 through 5.9. Significant associations were noted between handgrip strength and venous function indices (see Table 5.6 and Figures 5.6 and 5.7), and between arterial and venous function indices (see Table 5.6 and Figures 5.8 and 5.9).

73 50 Control arm Trained arm † † † † 1 45

40 Venous outflow (ml.100ml-1.min-

35 12345 Vis it

Figure 5.5. Changes in post occlusion forearm venous outflow. Values are mean ± SE. † =p<0.05 vs. visit 2.

7 r=0.58; p=0.001

6

5

4 Venous capacitance (mL·100mL tissue-1·min-1) 3 20 30 40 50 60 Strength (kg)

Figure 5.6. Relationship between handgrip strength & resting venous capacitance

74

Table 5.6. Pearson product moment correlation between arterial, venous and physical fitness measures Variables Resting vascular indices Post occlusion vascular indices Handgrip Blood Vascular Venous Venous Venous Blood Vascular Venous Venous Venous

strength inflow resistance capacitance compliance outflow inflow resistance capacitance compliance outflow

Handgrip strength 1.0 (kg) Blood inflow r=0.4* 1.0 (mL/100mL/min) Vascular resistance r=-0.26 r=-0.84* 1.0 (U) Venous capacitance r=0.58* r=0.35* r=-0.35* 1.0 (mL/100mL/min) Venous compliance r=0.27 r=-0.007 r=-0.3 r=0.67* 1.0 (%)

Resting vascular indices Resting vascular Venous outflow r=0.6* r=0.44* r=-0.5* r=0.81* r=0.52* 1.0 (mL/100mL/min) Blood inflow r=0.30 r=0.05 r=-0.2 r=0.4* r=0.45* r=0.4* 1.0 (mL/100mL/min) Vascular resistance r=-0.05 r=0.08 r=0.17 r=-0.26 r=-0.5* r=-0.27 r=-0.87* 1.0 (U) Venous capacitance r=0.57* r=0.23 r=-0.15 r=0.63* r=0.42* r=0.6* r=0.34# r=-0.15 1.0 (mL/100mL/min) Venous compliance r=0.25 r=-0.059 r=-0.07 r=0.43* r=0.55* r=0.5* r=0.5* r=-0.52* r=0.74* 1.0 (%) Venous outflow r=0.53* r=0.42* r=-0.51* r=0.76* r=0.56* r=0.95* r=0.43* r=-0.3‡ r=0.6* r=0.5* 1.0 Post occlusion vascular indices Post occlusion (mL/100mL/min)

*p<0.05; #p=0.06 † p= 0.07. ‡p=0.08

75 5.5. DISCUSSION

The main finding of this study was a unilateral improvement in regional peak forearm

arterial inflow and vascular resistance following handgrip exercise training. Furthermore,

the improvements were evident after the first week of training and appeared to plateau

during the remainder of the study. These findings confirm data from animal models (35,

61) as well as from human studies involving single conduit artery reactivity (2).

Uniquely, this study indicates a significant decrease in venous compliance after 1 week of training followed by a gradual return to baseline measures over the remainder of the study.

5.5.1. Effect of Handgrip Training on Arterial Indices Four weeks of handgrip training resulted in a 19% increase in regional peak arterial

inflow. This increase confirms previous data from our laboratory (21%) (4) as well as

others (25%) (65). Additionally, differences in peak arterial inflow and vascular

resistance between the dominant (non-trained) and non-dominant (trained) forearms at

visit 1 diminished by visit 2 suggesting vascular adaptations to habitual activities can be

matched following only 1 week of training.

The increased peak blood inflow in the trained arm without alterations in the

contra-lateral arm vasoreactivity, cardiovascular hemodynamics (i.e. blood pressure,

heart rate, and heart rate variability indices) and measures of cardiorespiratory fitness suggests that these adaptations are locally modulated. Although it is beyond the scope of this study to determine the mechanisms involved, these localized adaptations could involve either mechanical or chemical stimuli. The muscle contraction and relaxation phases associated with exercise training increase shear stress against the muscle bed arterial walls causing the release of endothelium-derived relaxing factor(s) (46). The

76 location of these changes is currently unclear. However, several studies have reported

increase in vasodilator mRNA and proteins at the site of large conduit vessels (13, 14)

while more recent studies suggest these changes occur in smaller distal vessels (36).

Alternatively, local accumulations of metabolites and vasoactive substances in distal

tissue beds could play a role in these adaptations. In fact, previous studies have shown an

increase in oxidative enzyme activity may precede the increases in vasoresponsiveness

(13).

Other studies have suggested regional neurohumoral changes following handgrip

exercise training (77). For example up-regulation of nitric oxide/prostacyclin is believed

to inhibit the release of norepinephrine from adrenergic nerves (9) and blunt norepinephrine-mediated vasoconstriction (74). Handgrip exercise training also appears to result in attenuated sympathetic vasoconstriction (56) attributed to changes in muscle metaboreceptors (67), mechanoreceptors (63) or a combination of both (53).

Additionally, Sinoway and colleagues demonstrated that this attenuation is accompanied by reduced norepinephrine arterial spillover and venous clearance (63).

Additional potential candidates that may contribute to the changes in vascular reactivity include up-regulation of prostaglandin (35, 61) and endothelium-derived hyperpolarizing factor (47), increased sensitivity of arterial smooth muscle to vasodilatory agents (34, 69), smooth muscle sarcoplasmic reticulum calcium unloading

(73) and reduced arterial wall sensitivity to endothelin-1 (28, 38). Clearly the speculations regarding the possible mechanism(s) for these adaptations need further refinement and experimental confirmation. Moreover, these speculations do not exclude the possibility of other plausible explanations for the regional vasoreactivity changes.

77 As depicted in Figure 5.1, post-occlusion blood inflow increased 13% after the first week, followed by a more gradual increase over the remainder of the training period.

These results confirm previous data from our laboratory (2) and appear to be consistent with the hypothesis advanced by McAllister and Laughlin. The authors speculate that vascular endothelial function is enhanced after just a few days of training and these adaptations could serve to buffer the increase in shear stress experienced during exercise

(47). Given that nitric oxide is an important signaling mechanism during exercise, and appears to be modifiable with exercise training, the nitric oxide system has emerged as a strong candidate responsible for the rapid changes in vascular reactivity. Both Wang et al.

(75), and Sessa et al. (59) report an enhanced nitric oxide-mediated dilation in the dog circumflex coronary artery in vivo, following 7 and 10 days of chronic exercise training.

More recently Johnson et al. (27) indicate increased acetylcholine induced relaxation and eNOS protein in porcine pulmonary arteries following 1 week of short-term training.

Kingwell and colleagues speculate that exercise-induced adaptations aimed at

meeting increased metabolic demands evolve from alterations in vasodilation to

adaptations in metabolic enzymes and vascular restructuring (31). In fact, recent data

suggest that nitric oxide contribution to exercise-training-induced arterial adaptations

might be a signal for subsequent increase in arterial wall dimensions (40). Increased shear stress (6) and flow-mediated nitric oxide production (46), associated with rhythmic exercise, inhibits intimal thickening and promotes arterial diameter enlargement (12, 40,

45). Wijnen and colleagues reported greater diameter of femoral but not brachial artery in

cyclists as compared to age-matched controls (78). In a more recent study, the diameter

of subclavian artery in the racket arm of tennis players was 19% greater than in the

78 opposite arm (26). Moreover, numerous studies have described exercise training-induced

increases in cross-sectional area by either recruiting already existing or developing new

vessels (24). These arterial structural alterations are associated with elevated skeletal

muscle blood inflow (57), and are believed to normalize increased blood shear stress and

to meet the increased metabolic demands associated with exercise training (12, 40, 52).

The plateau following the 1st week suggests a possible diminishing role of the

mechanism(s) involved in these early arterial adaptations. Since the individuals

participated in the study were sedentary, it is not clear whether the current training

regimen is sufficient to induce vascular adaptations in trained individuals. Therefore,

future studies should examine the effect of greater intensity and/or longer exercise

training program on arterial adaptations in sedentary as well as trained individuals.

The unilateral improvements in vasoreactivity indicate that exercise benefits are

specific to the trained arm confirming the importance of well-rounded exercise training

programs as advocated by ACSM (1). Therefore endurance and resistive exercise training

program targeting the heart and major muscle groups might be essential to induce overall

vascular adaptations. This is particularly relevant given the role improved arterial

function plays in blood delivery and perfusion, overall cardiovascular function, and exercise performance as well as in the prevention of cardiovascular diseases. In 1991,

Silber et al. reported increased peak forearm arterial inflow following cycle endurance

exercise training program (62). These findings were further confirmed following a

combined endurance and resistance exercise training program in healthy middle age

adults (44) as well as in patients with type 2 diabetes mellitus (42) and heart failure (43).

Exercise training-induced improvements in arterial function were also observed in the

79 coronary circulation of animals (37, 51) as well as in patients with coronary artery

disease (20), and recent myocardial infarction (25). Additionally, these arterial

adaptations were associated with improved endothelial function and exercise capacity

(42-44).

5.5.2. Effect of Handgrip Training on Venous Indices As presented in Tables 5.4 and 5.5, four weeks of handgrip training did not result in an

increase in forearm venous capacitance, compliance or outflow. In fact, venous measures

in the trained arm, at rest, and following 5-minutes of arterial occlusion showed a

consistent decline at the end of the first week followed by a gradual return to pre-training

values by visit 5 (see Figures 5.3-5.5). Figure 3 depicts a significant drop in resting

venous compliance in, the trained arm at visit 2 which remained below pre-training

values during visits 3 and 4.

The venous function responses to handgrip exercise training in this study are

intriguing. Since these responses are previously unreported, they are difficult to compare

to previous studies. However, the pattern of the data resembles the classic physiological

theory known as the Stress-Response-Adaptation. Hans Selye hypothesizes that

physiological stress causes an initial drop in physiological function followed by periods

of recovery and adaptation (58). The behavior of the data in the present study may reflect

the initial stress of training followed by a gradual recovery and beginning of an adaptive

period. Given that venous compliance is mainly affected by the sympathetic nervous

system (50, 76), regional elevation in sympathetic drive associated with the early phases of training (15, 56, 79) might result in venoconstriction. The increase in venoconstriction could lead to a reduction of venous blood volume in the forearm (15), which may account for the decline in venous function after the 1st week of training. Assuming that total

80 forearm blood volume remains the same following localized handgrip exercise training,

increased venoconstriction might decrease venous volume by driving blood towards the

arterial system. This theory may be supported by the concurrent increase in peak arterial

inflow during the 1st week of exercise training. Further analyses of the present data

revealed a weak relationship (p=0.023, r=0.58) between the magnitude of drop in venous

function and the increase in peak arterial blood inflow at visit 2; but it is premature to

suggest a direct link between those measures (see Figure 5.10).

Resting venous compliance gradually increases during visit 2-5. This response to

exercise training might be explained by two possible physiological changes. The persistence of the training stimulus may have resulted in an increased venous cross- sectional area by either recruiting already existing or developing new venules (49, 70).

Assuming the venoconstriction during the 1st week of training contributes to the increased

arterial inflow, newly accessible veins may allow for the delivery of more blood away

from the venous system into the arterial circulation. Additionally, venoconstriction may

also be attenuated subsequent to the readjustments in the sympathetic/parasympathetic

balance associated with exercise training (39). Although these mechanisms are simply

mere speculation, they reflect the recovery and adaptation phases of Selye’s theory (58).

The variations in training effect on the arterial and venous systems during the

training period might be due to the fundamental differences in the arterial and venous

structures and functions. For example, veins have thinner walls and are mainly composed

of smooth muscles, fibrous proteins, collagen, and elastin tissues (19). Additionally,

although an extensive neural network connects both sides of the vasculature, there are

differences in adrenoceptor distribution that might cause differential vascular

81 responsiveness to sympathetic stimulation (68). Endothelium-dependent dilatation also appears to operate differently in the venous side as a result of reduced relaxation, vasoagent production and/or smooth muscle responsiveness (16, 60). Given these differences and the effect of handgrip exercise training on the sympathetic nervous system (56) and endothelial tissue (17, 33, 46), it is reasonable to observe greater stiffness, less compliance and subsequent lower venous outflow under such conditions.

However, these speculations need further verifications in future studies.

Moreover, the unilateral changes in venous compliance suggest a stress response mediated by regional neural reflex and/or adrenoceptor sensitivity changes. Alterations in metaboreceptors (67), mechanoreceptors (63) stimulation and/or a combination of both

(53) have been proposed to induce attenuation in regional sympathetic activities. In the current study, cardiac autonomic modulation, supposedly a measure of autonomic nervous balance, did not change during the 4-week training program. However, heart rate variability might not be sensitive enough to detect localized neurohumoral changes.

Therefore, future studies should probably examine the effect of exercise training on regional neurohumoral modulations of the venous circulation, particularly the venoarteriolar sympathetic axon reflex.

The effect of an exercise training program with longer duration or higher intensity is unclear. Since the apparent increase in venous measures after visit 2 may lead to speculations regarding the adaptive changes using longer duration and/or higher intensity exercise training programs. Previous data from our laboratory showed a significant correlation between venous indices and handgrip strength as well as greater venous capacitance and outflow in individuals with higher handgrip strength. Individuals in the

82 higher strength group reported long-term involvements in resistance exercise training (3).

Moreover, Wecht et al reported greater venous compliance and total venous outflow in

“aerobically” trained individuals (76). As presented in Table 5.6 and Figures 5.6 and 5.7, results from the current study confirm the relationships between venous function indices and handgrip strength. Additionally, as in Figures 5.3-5.5, venous measures at visit 2 gradually increased reaching baseline levels at visit 5. However since training ended after

4 weeks, further venous adaptations to longer exercise training program are not known warranting future studies to examine the effect of different exercise training modes, intensities, and durations on venous indices.

70 r=0.6; p=0.002 60

50

40

30 Venous outflow 20 (mL·100mL tissue-1·min-1) 10 20 30 40 50 60 Strength (kg)

Figure 5.7. Relationship between handgrip strength & resting venous outflow 5.5.3. Relationship Between Arterial and Venous Functions Previous findings from our laboratory showed an association between forearm arterial and venous function indices (30). The results from the current study, as presented in

Table 5.6 and Figures 5.8 and 5.9, confirm these relationships, suggesting inter- communication between the arterial and venous circulations. Recent compelling evidence has emerged illustrating the importance of various aspects of the venous system beyond its typical role as a “passive volume reservoir”. For example, Tschakovsky & Hughson

83 reported greater arterial inflow with increased venous emptying following arm elevation

(72). The authors linked these findings to the local venoarteriolar sympathetic axon

reflex. The presence of nerve fiber collaterals from the sympathetic arteriolar plexus to

adjacent venules reflex was initially described in 1991 by Rygaard et al., and may serve

as the anatomical substrate for the local venoarteriolar sympathetic axon reflex (55).

35 r=0.5; p=0.006 30

25

20

Blood inflow 15

10 (mL·100mL tissue-1·min-1) 5 46810 Venous compliance (mL·100mL tissue-1·min-1)

Figure 5.8. Relationship between post occlusion blood inflow & venous compliance

Finally, evidence of cross-communication between the arterial and venous circulations indicates that venular endothelium releases a relaxing factor (s) that contributes to the vasodilatation of adjacent arterioles (21). Therefore, the relationship between arterial and venous function indices suggests that vasoreactivity and muscle perfusion may be influenced by venous hemodynamics, confirming the essential role of venous function in overall cardiovascular control, muscle perfusion and exercise performance (21, 72). Additionally, the correlation shown in Figure 10 between the decrease in venous compliance and the increase in arterial inflow following 1 week of handgrip training might further confirm the cross-communication between the arterial and venous circulations. The optimal location of the venous system in relation to the

84 arterial circulation may provide a mechanism for monitoring tissue metabolic status and potentially provide a valuable feedback mechanism for influencing the arterial reactivities. However more studies are certainly needed to further verify these speculations.

35 r=0.43; p=0.014 30

25

20

Blood inflow 15

10 (mL·100mL tissue-1·min-1) 5 30 40 50 60 70 Venous outflow (mL·100mL tissue-1·min-1)

Figure 5.9. Relationship between post occlusion blood inflow & venous outflow

5 r=0.58; p=0.023

0 Changes in blood inflow (mL·100mL tissue-1·min-1) -5 -10-8-6-4-20 2 Changes in venous compliance (mL·100mL tissue-1·min-1)

Figure 5.10. Relationship between changes in peak blood inflow & resting venous compliance

85

5.6. CONCLUSION

Data from the current study demonstrate a unilateral improvement in regional peak forearm arterial inflow and vascular resistance after the first week of handgrip exercise training and appeared to plateau during the remainder of the study. Additionally, the 4- week training did not result in increase in venous capacitance, compliance, and outflow.

However uniquely this study indicates a significant decrease in venous measures after 1 week of training followed by a gradual return to baseline values over the remainder of the study. The unilateral responses to handgrip exercise training further advocate the importance of whole-body exercise training programs targeting major muscle groups as recommended by recent ACSM exercise guidelines.

5.7. REFERENCES 1.American College of Sports Medicine Position Stand. The recommended quantity and quality of exercise for developing and maintaining cardiorespiratory and muscular fitness, and flexibility in healthy adults. Med Sci Sports Exerc 30: 975-991, 1998.

2.Allen JD, Geaghan JP, Greenway F and Welsch MA. Time Course of Improved Flow- Mediated Dilation after Short-Term Exercise Training. Med Sci Sports Exerc 35: 847- 853, 2003.

3.Alomari M, Reyes R, Wood R and Welsch M. Short-term reliability of forearm venous indices using strain gauge plethysmography: influence of handgrip strength on venous outflow and capacitance. SEACSM, Atlanta, Ga, 2003.

4.Alomari MA, Welsch MA, Prisby RD, Lee CM and Wood RH. Modification of forearm vascular function following short-term handgrip exercise training. Int J Sports Med 22: 361-365., 2001.

5.Balady GJ, Weiner DA, Rose L and Ryan TJ. Physiologic responses to arm ergometry exercise relative to age and gender. J Am Coll Cardiol 16: 130-135, 1990.

6.Barakat AI. Responsiveness of vascular endothelium to shear stress: potential role of ion channels and cellular cytoskeleton (review). Int J Mol Med 4: 323-332, 1999.

86 7.Berntson GG, Bigger JT, Jr., Eckberg DL, Grossman P, Kaufmann PG, Malik M, Nagaraja HN, Porges SW, Saul JP, Stone PH and van der Molen MW. Heart rate variability: origins, methods, and interpretive caveats. Psychophysiology 34: 623-648, 1997.

8.Clarkson P, Montgomery HE, Mullen MJ, Donald AE, Powe AJ, Bull T, Jubb M, World M and Deanfield JE. Exercise training enhances endothelial function in young men. J Am Coll Cardiol 33: 1379-1385, 1999.

9.Cohen RA and Weisbrod RM. Endothelium inhibits norepinephrine release from adrenergic nerves of rabbit carotid artery. Am J Physiol 254: H871-878, 1988.

10.Convertino VA, Mack GW and Nadel ER. Elevated central venous pressure: a consequence of exercise training-induced hypervolemia? Am J Physiol 260: R273-277, 1991.

11.Cramer M, Langlois Y, Beach K, Martin D and Strandness Jr DE. Standardization of venous flow measurements by strain gauge plethysmography in normal subjects. Bruit VII: 33-39, 1983.

12.Delp MD. Differential effects of training on the control of skeletal muscle perfusion. Med Sci Sports Exerc 30: 361-374, 1998.

13.Delp MD and Laughlin MH. Time course of enhanced endothelium-mediated dilation in aorta of trained rats. Med Sci Sports Exerc 29: 1454-1461, 1997.

14.Delp MD, McAllister RM and Laughlin MH. Exercise training alters endothelium- dependent vasoreactivity of rat abdominal aorta. J Appl Physiol 75: 1354-1363, 1993.

15.Duprez DA, De Buyzere M, De Sutter JM, Deman SA, De Pue NY and Clement DL. Venous responses to rhythmic exercise in contralateral forearm and calf. Eur J Appl Physiol Occup Physiol 65: 94-98, 1992.

16.el Khatib H, Lupinetti FM, Sanofsky SJ and Behrendt DM. Production of endothelium-dependent relaxation responses by saphenous vein grafts in the canine arterial circulation. Surgery 110: 523-528, 1991.

17.Fukaya Y and Ohhashi T. Acetylcholine- and flow-induced production and release of nitric oxide in arterial and venous endothelial cells. Am J Physiol 270: H99-106, 1996.

18.Green DJ, Cable NT, Fox C, Rankin JM and Taylor RR. Modification of forearm resistance vessels by exercise training in young men. J Appl Physiol 77: 1829-1833, 1994.

19.Guyton AC, Richardson TQ and Langston JB. Regulation of cardiac output and venous return. Clin Anesth 3: 1-34, 1964.

87

20.Hambrecht R, Wolf A, Gielen S, Linke A, Hofer J, Erbs S, Schoene N and Schuler G. Effect of exercise on coronary endothelial function in patients with coronary artery disease. N Engl J Med 342: 454-460, 2000.

21.Hester RL and Choi J. Blood flow control during exercise: role for the venular endothelium? Exerc Sport Sci Rev 30: 147-151, 2002.

22.Hester RL and Hammer LW. Venular-arteriolar communication in the regulation of blood flow. Am J Physiol Regul Integr Comp Physiol 282: R1280-1285, 2002.

23.Hokanson DE, Sumner DS and Strandness Jr. DE. An electrically calibrated plethysmograph for direct measurement of limb blood flow. IEEE transactions on bio- medical engineering 22: 21-25, 1975.

24.Holloszy JO and Coyle EF. Adaptations of skeletal muscle to endurance exercise and their metabolic consequences. J Appl Physiol 56: 831-838, 1984.

25.Hosokawa S, Hiasa Y, Takahashi T and Itoh S. Effect of regular exercise on coronary endothelial function in patients with recent myocardial infarction. Circ J 67: 221-224, 2003.

26.Huonker M, Schmid A, Schmidt-Trucksab A, Grathwohl D and Keul J. Size and blood flow of central and peripheral arteries in highly trained able-bodied and disabled athletes. J Appl Physiol, 2002.

27.Johnson LR, Rush JW, Turk JR, Price EM and Laughlin MH. Short-term exercise training increases ACh-induced relaxation and eNOS protein in porcine pulmonary arteries. J Appl Physiol 90: 1102-1110, 2001.

28.Jones AW, Rubin LJ and Magliola L. Endothelin-1 sensitivity of porcine coronary arteries is reduced by exercise training and is gender dependent. J Appl Physiol 87: 1172- 1177, 1999.

29.Kenney WL and Armstrong CG. The effect of aerobic conditioning on venous pooling in the foot. Med Sci Sports Exerc 19: 474-479, 1987.

30.Khalil MS, Reyes R, Alomari M, Wood R, Brady S and Welsch MA. Relationship between cardiovascular endurance and forearm vascular function in women. SEACSM, Atlanta, GA, 2003.

31.Kingwell BA. Nitric oxide-mediated metabolic regulation during exercise: effects of training in health and cardiovascular disease. Faseb J 14: 1685-1696, 2000.

88 32.Kingwell BA, Sherrard B, Jennings GL and Dart AM. Four weeks of cycle training increases basal production of nitric oxide from the forearm. Am J Physiol 272: H1070- 1077, 1997.

33.Koller A, Dornyei G and Kaley G. Flow-induced responses in skeletal muscle venules: modulation by nitric oxide and prostaglandins. Am J Physiol 275: H831-836, 1998.

34.Koller A, Huang A, Sun D and Kaley G. Exercise training augments flow-dependent dilation in rat skeletal muscle arterioles. Role of endothelial nitric oxide and prostaglandins. Circ Res 76: 544-550, 1995.

35.Laughlin MH. Endothelium-mediated control of coronary vascular tone after chronic exercise training. Med Sci Sports Exerc 27: 1135-1144, 1995.

36.Laughlin MH, Pollock JS, Amann JF, Hollis ML, Woodman CR and Price EM. Training induces nonuniform increases in eNOS content along the coronary . J Appl Physiol 90: 501-510, 2001.

37.Laughlin MH, Rubin LJ, Rush JW, Price EM, Schrage WG and Woodman CR. Short- term training enhances endothelium-dependent dilation of coronary arteries, not arterioles. J Appl Physiol 94: 234-244, 2003.

38.Laughlin MH, Schrage WG, McAllister RM, Garverick HA and Jones AW. Interaction of gender and exercise training: vasomotor reactivity of porcine skeletal muscle arteries. J Appl Physiol 90: 216-227, 2001.

39.Lee CM, Wood RH and Welsch MA. Influence of short-term endurance exercise training on heart rate variability. Med Sci Sports Exerc 35: 961-969, 2003.

40.Lehoux S, Tronc F and Tedgui A. Mechanisms of blood flow-induced vascular enlargement. Biorheology 39: 319-324, 2002.

41.Louisy F, Jouanin JC and Guezennec CY. Filling and emptying characteristics of lower limb venous network in athletes. Study by postural plethysmography. Int J Sports Med 18: 26-29, 1997.

42.Maiorana A, O'Driscoll G, Cheetham C, Dembo L, Stanton K, Goodman C, Taylor R and Green D. The effect of combined aerobic and resistance exercise training on vascular function in type 2 diabetes. J Am Coll Cardiol 38: 860-866, 2001.

43.Maiorana A, O'Driscoll G, Dembo L, Cheetham C, Goodman C, Taylor R and Green D. Effect of aerobic and resistance exercise training on vascular function in heart failure. Am J Physiol Heart Circ Physiol 279: H1999-2005, 2000.

89 44.Maiorana A, O'Driscoll G, Dembo L, Goodman C, Taylor R and Green D. Exercise training, vascular function, and functional capacity in middle-aged subjects. Med Sci Sports Exerc 33: 2022-2028, 2001.

45.Marano G, Palazzesi S, Vergari A and Ferrari AU. Protection by shear stress from collar-induced intimal thickening: role of nitric oxide. Arterioscler Thromb Vasc Biol 19: 2609-2614, 1999.

46.McAllister RM, Hirai T and Musch TI. Contribution of endothelium-derived nitric oxide (EDNO) to the skeletal muscle blood flow response to exercise. Med Sci Sports Exerc 27: 1145-1151, 1995.

47.McAllister RM and Laughlin MH. Short-term exercise training alters responses of porcine femoral and brachial arteries. J Appl Physiol 82: 1438-1444, 1997.

48.Milani J, Fernhall B and Manfredi T. Estimating oxygen consumption during treadmill and arm ergometry activity in males with coronary artery disease. J Cardiopulm Rehabil 16: 394-401, 1996.

49.Miyachi M, Tanaka H, Yamamoto K, Yoshioka A, Takahashi K and Onodera S. Effects of one-legged endurance training on femoral arterial and venous size in healthy humans. J Appl Physiol 90: 2439-2444, 2001.

50.Pang CC. Autonomic control of the venous system in health and disease: effects of drugs. Pharmacol Ther 90: 179-230, 2001.

51.Parker JL, Oltman CL, Muller JM, Myers PR, Adams HR and Laughlin MH. Effects of exercise training on regulation of tone in coronary arteries and arterioles. Med Sci Sports Exerc 26: 1252-1261, 1994.

52.Prior BM, Lloyd PG, Yang HT and Terjung RL. Exercise-induced vascular remodeling. Exerc Sport Sci Rev 31: 26-33, 2003.

53.Ray CA and Hume KM. Sympathetic neural adaptations to exercise training in humans: insights from microneurography. Med Sci Sports Exerc 30: 387-391, 1998.

54.Rothe CF. Reflex control of veins and vascular capacitance. Physiol Rev 63: 1281- 1342, 1983.

55.Rygaard K, Moller M and Henriksen O. Presence of nerve fiber collaterals from the sympathetic arteriolar plexus the concomitant venules in dog skeletal muscle. Acta Physiol Scand Suppl 603: 115-118, 1991.

56.Saito M, Kato M and Mano T. Comparison of sympathetic nerve activity during handgrip exercise performed with the dominant and non-dominant arm. Environ Med 44: 60-62, 2000.

90

57.Schmidt-Trucksass A, Schmid A, Brunner C, Scherer N, Zach G, Keul J and Huonker M. Arterial properties of the carotid and femoral artery in endurance-trained and paraplegic subjects. J Appl Physiol 89: 1956-1963, 2000.

58.Selye H. The stress of life. New York: McGraw-Hill, 1976.

59.Sessa WC, Pritchard K, Seyedi N, Wang J and Hintze TH. Chronic exercise in dogs increases coronary vascular nitric oxide production and endothelial cell nitric oxide synthase gene expression. Circ Res 74: 349-353, 1994.

60.Shapira OM, Xu A, Aldea GS, Vita JA, Shemin RJ and Keaney JF, Jr. Enhanced nitric oxide-mediated vascular relaxation in radial artery compared with internal mammary artery or saphenous vein. Circulation 100: II322-327, 1999.

61.Shen W, Zhang X, Zhao G, Wolin MS, Sessa W and Hintze TH. Nitric oxide production and NO synthase gene expression contribute to vascular regulation during exercise. Med Sci Sports Exerc 27: 1125-1134, 1995.

62.Silber D, McLaughlin D and Sinoway L. Leg exercise conditioning increases peak forearm blood flow. J Appl Physiol 71: 1568-1573, 1991.

63.Sinoway L, Shenberger J, Leaman G, Zelis R, Gray K, Baily R and Leuenberger U. Forearm training attenuates sympathetic responses to prolonged rhythmic forearm exercise. J Appl Physiol 81: 1778-1784, 1996.

64.Sinoway LI, Musch TI, Minotti JR and Zelis R. Enhanced maximal metabolic vasodilatation in the dominant forearms of tennis players. J Appl Physiol 6: 673-678, 1986.

65.Sinoway LI, Shenberger J, Wilson J, McLaughlin D, Musch T and Zelis R. A 30-day forearm work protocol increases maximal forearm blood flow. J Appl Physiol 62: 1063- 1067, 1987.

66.Solomito AN, Marsh L, Alomari M and Welsch MA. Technical considerations for the assessment of venous function: influence of occlusion time and pressure. SEACSM, Atlanta, GA, 2003.

67.Somers VK, Leo KC, Shields R, Clary M and Mark AL. Forearm endurance training attenuates sympathetic nerve response to isometric handgrip in normal humans. J Appl Physiol 72: 1039-1043, 1992.

68.Stein CM, Deegan R and Wood AJ. Lack of correlation between arterial and venous beta-adrenergic receptor sensitivity. Hypertension 29: 1273-1277, 1997.

91 69.Sun D, Huang A, Koller A and Kaley G. Short-term daily exercise activity enhances endothelial NO synthesis in skeletal muscle arterioles of rats. J Appl Physiol 76: 2241- 2247, 1994.

70.Suzuki J, Kobayashi T, Uruma T and Koyama T. Time-course changes in arteriolar and venular portions of in young treadmill-trained rats. Acta Physiol Scand 171: 77-86, 2001.

71.Task Force of the European Society of and the North American Society of Pacing and Electrophysiology. Heart rate variability: standards of measurement, physiological interpretation and clinical use. Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology. Circulation 93: 1043-1065., 1996.

72.Tschakovsky ME, Shoemaker JK and Hughson RL. Vasodilation and muscle pump contribution to immediate exercise hyperemia. Am J Physiol 271: H1697-1701, 1996.

73.Underwood FB, Laughlin MH and Sturek M. Altered control of calcium in coronary smooth muscle cells by exercise training. Med Sci Sports Exerc 26: 1230-1238, 1994.

74.Vanhoutte PM and Miller VM. Alpha 2-adrenoceptors and endothelium-derived relaxing factor. Am J Med 87: 1S-5S, 1989.

75.Wang J, Wolin MS and Hintze TH. Chronic exercise enhances endothelium-mediated dilation of epicardial coronary artery in conscious dogs. Circ Res 73: 829-838, 1993.

76.Wecht JM, de Meersman RE, Weir JP, Bauman WA and Grimm DR. Effects of autonomic disruption and inactivity on venous vascular function. Am J Physiol Heart Circ Physiol 278: H515-520, 2000.

77.Wiegman DL, Harris PD, Joshua IG and Miller FN. Decreased vascular sensitivity to norepinephrine following exercise training. J Appl Physiol 51: 282-287, 1981.

78.Wijnen JA, Kuipers H, Kool MJ, Hoeks AP, van Baak MA, Struyker Boudier HA, Verstappen FT and Van Bortel LM. Vessel wall properties of large arteries in trained and sedentary subjects. Basic Res Cardiol 86 Suppl 1: 25-29, 1991.

79.Williams CA, Mudd JG and Lind AR. Sympathetic control of the forearm blood flow in man during brief isometric contractions. Eur J Appl Physiol Occup Physiol 54: 156- 162, 1985.

80.Williams DP, Going SB, Milliken LA, Hall MC and Lohman TG. Practical techniques for assessing body composition in middle-aged and older adults. Med Sci Sports Exerc 27: 776-783, 1995.

92 CHAPTER 6. SUMMARY

Four projects were conducted in partial fulfillment for the requirement of this degree. The main findings of these studies were arterial and venous measures are reproducible, can differentiate between populations, and are modifiable with exercise training.

The results of the first study (chapter 2) indicate that a pressure of 7 mmHg <

DBP and a period of 10 minutes venous occlusion produced the greatest venous capacitance and outflow, without altering arterial indices. Additionally, the examination of forearm vascular function indices using mercury strain-gauge plethysmography are reproducible and associated with fitness measures (estimated VO2peak and handgrip strength). The association between handgrip strength and venous system hemodynamics implies that muscular conditioning might be important for venous health, which warrants future studies examining the effect of muscular strength modifications on venous function.

The results of the second study (chapter 3) indicate that arterial function indices are reduced in heart failure patients confirming previous studies (8, 15, 16). Additionally, venous capacitance and outflow are diminished and associated with maximum walking distance on a 6-minute walk test in heart failure suggesting the importance of venous function to exercise intolerance.

The results of the third study (chapter 4) indicate unilateral improvements in arterial inflow following 4-week handgrip exercise training similarly for the LO and HI intensity groups. The unilateral improvements suggest that vascular adaptations are locally mediated confirming previous findings (11, 14).

93 The results of the final study (chapter 5) indicate unilateral improvement in arterial inflow and vascular resistance following 1 week of handgrip exercise training.

Additionally, the exercise training resulted in a decrease in venous function indices followed by gradual increase to baseline values. The one week improvement in arterial function indices confirm previous findings from animals models (3) and human single artery (2) and suggest vascular adaptations to habitual activities can be matched following only 1 week of training. The venous function responses to handgrip exercise training in this study are previously unreported thus they are difficult to compare to other studies. However, the pattern of the data resembles the classic physiological theory known as the Stress-Response-Adaptation. Hans Selye hypothesizes that physiological stress causes an initial drop in physiological function followed by periods of recovery and adaptation (10). The unilateral responses to handgrip exercise training further advocate the importance of whole-body exercise training programs targeting major muscle groups as recommended by recent ACSM exercise guidelines (1).

The findings regarding the venous system are particularly uniqe and relevant considering venous abnormalities in aging (6, 7), and patients with heart failure (5, 9, 13) and venous diseases (4, 12). Therefore, recognizing the contribution of skeletal muscle in venous return and overall cardiovascular performance during exercise, factors associated with muscle fitness modification might contribute to increased overall cardiovascular health subsequent to improved venous function, namely venous outflow and capacitance.

However, it is recognized that the aforementioned studies are limited and more investigations are cairtainly needed to futher expand our knowledge regarding the effect of exercise training on the venous system.

94 6.1. REFERENCES

1.American College of Sports Medicine Position Stand. The recommended quantity and quality of exercise for developing and maintaining cardiorespiratory and muscular fitness, and flexibility in healthy adults. Med Sci Sports Exerc 30: 975-991, 1998.

2.Allen JD, Geaghan JP, Greenway F and Welsch MA. Time Course of Improved Flow-Mediated Dilation after Short-Term Exercise Training. Med Sci Sports Exerc 35: 847-853, 2003.

3.Delp MD and Laughlin MH. Time course of enhanced endothelium-mediated dilation in aorta of trained rats. Med Sci Sports Exerc 29: 1454-1461, 1997.

4.Hartmann BR, Drews B and Bassenge E. Venous Function in Patients with Venous Disease and Healthy Controls Before and After a Bathing Procedure and Subsequent Cold Stimulus. International Journal of Angiology 7: 252-254, 1998.

5.Ikenouchi H, Iizuka M, Sato H, Momomura S, Serizawa T and Sugimoto T. Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Jpn Heart J 32: 17-34, 1991.

6.Lanne T and Olsen H. Decreased capacitance response with age in lower limbs of humans--a potential error in the study of cardiovascular reflexes in ageing. Acta Physiol Scand 161: 503-507, 1997.

7.Olsen H and Lanne T. Reduced venous compliance in lower limbs of aging humans and its importance for capacitance function. Am J Physiol 275: H878-886, 1998.

8.Poole-Wilson PA, Buller NP and Lipkin DP. Regional blood flow, muscle strength and skeletal muscle histology in severe congestive heart failure. Am J Cardiol 62: 49E- 52E, 1988.

9.Sato H, Ikenouchi H, Aoyagi T, Matsui H, Mochizuki T, Momomura S, Serizawa T, Iizuka M and Sugimoto T. Pathophysiology and evaluation of severity of congestive heart failure on the basis of venous characteristics. Jpn Circ J 53: 141-145, 1989.

10.Selye H. The stress of life. New York: McGraw-Hill, 1976.

11.Sinoway LI, Shenberger J, Wilson J, McLaughlin D, Musch T and Zelis R. A 30- day forearm work protocol increases maximal forearm blood flow. J Appl Physiol 62: 1063-1067, 1987.

12.Tretbar LL. Chronic venous insufficiency of the legs: pathogenesis of venous ulcers. J Enterostomal Ther 14: 105-108, 1987.

95 13.Welsch MA, Allen JD and Geaghan JP. Stability and reproducibility of brachial artery flow-mediated dilation. Med Sci Sports Exerc 34: 960-965, 2002.

14.Yasuda Y and Miyamura M. Cross transfer effects of muscular training on blood flow in the ipsilateral and contralateral forearms. Eur J Appl Physiol Occup Physiol 51: 321-329, 1983.

15.Zelis R and Flaim SF. Alterations in vasomotor tone in congestive heart failure. Prog Cardiovasc Dis 24: 437-459, 1982.

16.Zelis R, Mason DT and Braunwald E. A comparison of the effects of vasodilator stimuli on peripheral resistance vessels in normal subjects and in patients with congestive heart failure. J Clin Invest 47: 960-970, 1968.

96 APPENDIX A: LITERATURE REVIEW

97 SECTION 1. INTRODUCTION

With the latest developments in technology various vascular functions can be measured

in human, noninvasively, with minimum expense and relative ease. These functions

include arterial inflow, vascular resistance, venous filling capacity, venous emptying rate,

vessel diameter, vessel density, and volume changes. These recent advancements have

created opportunities for unprecedented prospects in vascular physiology research.

Presently, research is approaching integrative understanding of vascular function. This

kind of understanding requires gathering information obtained from different vascular

investigations and applying it to different conditions including exercise, diseases,

pharmacological therapy and alternative therapy. Additionally relate this information to

other physiologically-related disciplines including cardiology, neurology, endocrinology,

microbiology as well as psychology.

The main focus of this review is to discuss peripheral vascular function and

control at rest and during activities as well as adaptations to exercise training.

Additionally, vascular adaptations to diseases, the basic structure of the vasculature,

physical characteristics of the circulation, and the available tools to assess vascular

function are also discussed. This introduction is followed by seven sections. Section two

of the review focuses on the basic structure of different segments of the vascular system

including arteries, arterioles, , venules and veins. The wall component of each

segment is described as well as the distribution and density of each segment throughout

the body. The basic physical principles govern the circulation are discussed in section

three. This section sheds the light on various vascular physical characteristics that contribute to the control of blood flow in the arterial and venous side of the circulation

98 including vessel diameter, vessel length, transmural pressure, and blood velocity. The relationship between pressure, vascular resistance, and blood flow as well as hemodynamics essential for the venous system are also explained. Section four is divided into two parts. The first part deals with central neurohormonal factors contribute to vascular function control whereas the second part addresses the peripheral aspect of vascular control. Acute responses and control of the vascular system to exercise are discussed in section five. Similar to section four the details of the central and peripheral aspects are addressed in this section. However, since the cardiovascular system is functionally integrated, cardiac as well as blood acute adjustments to exercise training are briefly discussed. Section six deals with vascular adaptations to exercise training. This section provides a review of some of the studies examined arterial as well as venous adaptations to exercise training. Exercises promote systematic vascular adaptations are first discussed followed by regional vascular adaptations to small muscle group exercises.

In section seven, peripheral vascular adaptations to various diseases are briefly discussed.

Finally section eight overviews some of the available techniques to measure peripheral vascular function.

Since the subjects covered in this review spans over broad aspects of vascular physiology the author obtained information from various original articles, review articles, as well as textbooks. Additionally, text in this review is supported with illustrative figures and pictures selected from textbooks, review articles and original articles to explain some physiological phenomena covered in this review.

This review is certainly less than complete, however it is an attempt to reveal the available knowledge pertaining to vascular physiology. Additionally, this review is

99 expected to serve as a self-learning tool and platform for future self-advancements in integrative understanding of cardiovascular physiology.

100 SECTION 2. VASCULAR SYSTEM STRUCTURE

Blood vessels of the body form a closed delivery system that allows blood to be transported from the heart to the capillaries and back to the heart. Oxygenated blood leaving the heart passes through vessels of progressively smaller diameters referred to as arteries, arterioles and capillaries [1, 2]. On the other hand, deoxygenated blood, returning to the heart from the capillaries, passes through vessels of progressively larger diameters called veins (figure 2-1). The capillaries, which link the arterial side to the venous side, permit the exchange of nutrients and waste products between blood and living cells [1, 2]. This section will focus on the general structure of the peripheral vascular system. The functional anatomy details of all the vasculature segments, including arteries, arterioles, capillaries venules and veins, will be discussed.

2.1. ARTERIAL SYSTEM

There are two great arteries that emerge from the heart, the aorta from the left ventricle and the from the right ventricle. The arteries are efferent vessels that carry blood away from the heart to the capillary beds throughout the body. All arteries in the body transport oxygenated blood from the heart to body tissues except the pulmonary vascular tree that carries deoxygenated blood from the heart to the [2, 3].

Therefore, the most reliable way to classify blood vessels is by the direction in which they carry blood, either toward or away from the heart. Every artery enters an organ branches into 6-8 smaller vessels to become arterioles while arterioles branch 2-5 times to eventually become capillaries [4].

101 2.1.1. Arteries The lumen, the central canal of all blood vessels, is formed from a thick wall composed of three layers or tunicae. These tunicae are , and tunica adventitia [5]. The tunica intima, the innermost layer, has a lining of endothelial cells, a thin subendothelial layer of fine areolar tissue, and an internal elastic layer called the internal elastic lamina (figure 2-1 & 2-2). The tunica media (the middle layer) is the thickest layer of the arterial wall in large arteries [2, 3].

Figure LR2.1. Vessel wall characteristics Internal diameter, wall thickness, and relative amounts of the principle components of the vessel walls of the various blood vessels that compose the . Taken from [6].

It is composed mainly of connective tissues, smooth muscle cells and elastic fibers. The tunica media of the largest artery walls have more elastic tissue than smooth muscle, whereas smaller arteries contain more smooth muscle cells than elastic fibers [5]. The greater prevalence of elastic tissue helps larger arteries to stretch in order to adjust for the large change in pressure as the ventricles contract forcefully and pump blood into these type of vessels. For example, the aorta and pulmonary wall elasticity allows these vessels

102 to adjust for the pressure created by the ejected blood from the ventricles [3]. The

outermost layer of the vessel, the tunica adventitious, is mainly made of collagen and

elastic fibers with some smooth muscle fibers extending longitudinally next to the outer

border of the tunica media. This layer comprises nerves and lymphatic vessels in addition

to a capillary network formed by small vessels called “” to nourish the walls of large arteries (> 20 mm) [4, 7].

Figure LR2.2. Typical . Shown different segments of a vessel. Taken from [8].

2.1.2. Arterioles Arteries located further away from the heart branch into smaller and smaller vessels, then

into arterioles shortly before merging with the capillaries. The arterioles contain more

smooth muscle cells (figure 2-1 & 2-2). Arterioles and capillaries are connected by

structurally intermediate vessels called terminal arterioles or [7]. The

metarterioles are wrapped around by highly innervated nerves and smooth muscles to

form precapillary sphincters however they do not have an internal elastic layer or their

103 own supply. When these sphincters contract, blood flow is routed bypassing the capillary bed, however when the sphincters are relaxed, blood flow to the capillary bed and surrounding tissues increases [7]. The opening of the sphincters depends on the

metabolic needs of the surrounding tissues. As the metabolic activities of the surrounding

tissues increase the sphincters are opened to increase blood perfusion. Whereas when the

blood supply is no longer needed in larger quantities, the sphincters close [3, 7]. [4].

Therefore arterioles and metarterioles, due to their high content of smooth muscles,

sympathetic innervations, and location are the segments of the arterial tree most

responsible for mediating flow of blood to the capillaries [7, 9].

2.2. CAPILLARIES

As terminal arteries extend further away from the heart they branch out to form

capillaries. The capillaries run throughout all body tissues to form capillary beds in each

tissue. This serves as a point of connection between the arterial system and venous system. Capillaries are the most abundant blood vessels, which give them an enormous surface area available for exchange of gases, fluids, nutrients, and waste products between blood and nearby cells (figure 2-3) [5, 10]. They are the smallest among all vessels with a diameter of 8-10 µm, which is just large enough for the passage of one erythrocyte (figure 2-1). For example, the diameter of a medium-sized capillary is about

500 and 600 times smaller than a medium-sized artery and vein respectively. However, capillary diameters vary depending on the function of the tissue that is being supplied [5].

104

Figure LR2.3. Blood volume distribution in different compartments of the circulation. Taken from [2].

Capillary walls are mainly composed of only a tunica intima layer that contains endothelial cells on a thin basement membrane of glycoprotein (figure 2-1) [3, 11]. The endothelial cells are held together by tight junctions to prevent small molecule passage.

However, these junctions have gaps of unjoined membranes called intercellular clefts to allow for molecular movement across the capillary wall [6]. However in areas where high rate of molecular exchange is required, there are fenestrations or pores that penetrate the endothelial tissue to increase the capillary permeability for small molecules. These kinds of capillaries are called fenestrated capillaries. They are found in areas like the small intestine and the synovial membranes of joints, as large quantities of fluids and nutrients are exchanged with blood. In low molecular exchange areas (i.e. skeletal muscle, skin, and the central nervous system) are fed with continuous capillaries, in which the endothelial tissues contains no pores or fenestrations [2, 3]. Most of the small molecules are transported across capillary walls through the intercellular cleft. Moreover,

105 cytoplasmic vesicles that invaginate from the plasma membrane and migrate across the cell are used to move larger molecules across. Whereas oxygen and carbon dioxide seem

to be the only important molecules transported through diffusion [2].

2.3. VENOUS SYSTEM

As blood passes through capillary beds, it gets depleted of oxygen and nutritional

molecules and loaded with by-products (i.e. carbon dioxide and hydrogen ions).

Additionally blood pressure is reduced due to the high-resistance in arterioles and

capillary beds. Moreover, blood returns back to the right atrium through the smaller

venules then to the larger veins until it reaches the largest veins in the body, which are the

superior and inferior vena cava [11, 12]. Therefore, veins receive blood from smaller

vessels instead of giving blood off to branches as arteries do. Veins are more abundant

than their counterpart arteries and are located superficially beneath the skin as well as deeply in body tissues (figure 2-3) [1, 3, 13].

2.3.1. Venules Thoroughfare channels serve as an extension and are structurally intermediate between

capillaries and venules. These channels are connected to postcapillary venules, which are

the smallest of all veins. Postcapillary venules comprise of a thin layer of endothelium

and sometimes function like capillaries as inflammatory fluids and leukocytes leave

circulation through them. Whereas the larger venules consist of one or two layers of

smooth muscle cells, a tunica media as well as a thin tunica adventitia (figure 2-1 & 2-4)

[3, 5, 12].

106

Figure LR2.4. Typical vein. Shown longitudinal cut for a medium- sized vain. Emphasized the valves and different layers compose the venous vessel. Taken from [8].

2.3.2. Veins As blood approaches the heart through the circulatory system venules get larger to form veins. Opposite to their arterial counterpart, the tunica adventitia in veins is thicker than the tunica media. Moreover, the vein walls are thinner with larger lumens and contain less elastic tissue, collagen tissue, and smooth muscle thus veins are very distensible and comprisable (Figure 2-1) [5, 13].

Blood pressure is considerably lower in the veins than in the arteries, which reduces the propelling power of the blood as it travels back to the heart [13, 14]. Venous valves prevent blood from flowing backwards and assist in moving blood toward the right atrium. These valves are formed by the in-foldings of the tunica intimae, where each fold forms a bicuspid valve [3, 13]. Venous valves are more prevalent in the lower extremities where blood is moved against gravity. Further, fewer valves are found in the head and neck veins and there are none in the veins of the thoracic and abdominal

107 cavities. Additionally the compression of the venous walls by contracting skeletal muscle aids in propelling venous blood back toward the heart [3].

108 SECTION 3. CIRCULATORY HEMODYNAMICS

The primary function of the peripheral circulation is to deliver oxygenated blood to

tissues around the body and transport deoxygenated blood back to the heart. Meanwhile it

also transports other essential chemical substances (e.g. hormones, enzymes, and

minerals) to maintain the survival of the living cells [3]. In order to accomplish these physiological essentials blood flow is shunted according to tissue metabolic requirements

for substrate delivery and waste removal, which differs during the course of various daily

life situations. For example, skeletal muscle’s demand for nutrients and oxygen during

exercise is higher than its needs during resting condition. Likewise, the nutritional

requirements for stomach smooth muscle increase following a big meal [2, 6]. Several

aspects within the cardiovascular system interact together to achieve proper substrate

delivery and waste removal including cardiac muscle pump, vascular function, and blood

volume. These functions are further controlled by neurohumoral modulation, electrolyte balance, and tissue metabolic activities [7].

The vascular system plays a profound role in controlling the flow of blood to

tissues throughout the body. The parallel arrangement throughout the vascular system

allows for proper blood flow to organs around the body [2]. Small arteries and arterioles

are the principal site of blood flow control where terminal arterioles act as “stopcocks” to

accurately alter blood distribution to these organs according to their needs (figure 3-1).

However due to its structure and arrangement the vasculature is governed by physical principles that control blood flow [6]. This section will review the functional physical

principles of the peripheral vascular system.

109

Figure LR3.1. Schematic diagram of the parallel and series arrangement of the vessels composing the circulatory system. The drawing shows small valves, “stopcocks” likes, to control blood inflow representing vascular resistance on the arterial side (on the right). Larger venous diameter is also shown indicating the capacitance specialty of the venous system. Taken from [15].

3.1. GENERAL PHYSICAL PRINCIPLES OF PERIPHERAL CIRCULATION

The vasculature is divided into arterial, capillary and venous systems and proceeds within the body in the order of arteries, arterioles, capillaries, venules and veins. Vessels of the arterial system progressively get smaller in diameters, reaching the smallest as they approach the capillary beds. This is in contrast to the diameter of the venous system, where vessels get larger as they reach the heart [16]. The capillaries operate as an exchange structure for metabolic substances between blood and tissues whereas the

110 arteries and veins serve as distribution and collecting structures, respectively, for blood

[5, 17].

The systemic circulation contains the largest blood volume (83%) of total blood volume with 64% stored in the veins, 16% in the arteries and 4% in the capillaries.

Additionally, 8% is in the and 4% is usually within the heart chambers (figure 3-2) [2].

Figure LR3.2. Blood volume distribution in different compartments of the circulation. Taken from [2].

The capillaries have the largest total cross-sectional area (2500 cm2) allowing for sufficient blood/tissue exchange. In addition, the total cross-sectional area of all veins is

338 cm2, which is about four times that of arteries (62.5 cm2) (figure 3-3)[1]. Blood pressure also fluctuates at different segments of the peripheral circulatory system averaging approximately 100 mmHg at the aortic level and progressively declining to pressures approaching 0 mmHg in the (figure 3-3). This drop in blood

111 pressure is a function of blood flow rate and peripheral resistance within the circulation, as will be explained later [6].

Figure LR3.3. Changes in blood pressure and velocity at different cross-sectional areas. Taken from [18].

3.2. VASCULAR DISTENSIBILITY

The vascular system has an astonishing capacity to accommodate large changes in blood volume across all vessel segments due to vascular distensibility [19]. This physical

112 feature is characterized as the ability of the vasculature to expand and described as the volume of fluid needed to increase pressure by one unit. In other words, it is the ratio increase in volume for each millimeter mercury increase in pressure and could be expressed in the following formula [3, 20]:

Vascular distensibility = Increase in volume*Original volume/increase in pressure

Vascular distensibility is vital for proper blood pressure control and to accommodates the pulsatile output of the heart. Additionally, it provides smooth and continuous flow of blood through the tissues by averaging out the pressure pulsations so that tissue blood flow will not be disrupted [20]. Arteries are far stronger and less elastic than the veins due to thicker walls and a greater smooth muscle content. Therefore, veins are about 6-10 times more distensible than their arterial counterparts, making them the ultimate blood reservoir within the circulatory system [21, 22].

Capacitance is similar to and influenced by distensibility however these two physical characteristics of the vasculature are not identical. Vessel capacitance or compliance is the ability to accommodate changes in volume for changes in pressure when fluid is added. Thus capacitance describes volume-pressure relationship with respect to blood vessel size and volume added as stated in the following formula [3]:

Vascular capacitance = Increase in volume/increase in pressure

All of the cardiovascular system compartments, including the heart, are considered capacitance vessels since they contain volume. However as the veins hold

~70% of the circulating blood and large changes in veins’ volume cause little changes in pressure they provide the major capacitance function in the circulation [15]. Distensibility

113 decreases with aging as the lining of the vasculature becomes infiltrated with less elastic

fibers. Additionally diseases (i.e. diabetes & atherosclerosis), changes in autonomic

nerves balance, and various drugs might further increase the rigidity of the vascular

system [2, 23, 24].

3.3. BLOOD VELOCITY AND BLOOD FLOW

It is important to distinguish between blood velocity and blood flow in order to describe

the hemodynamics of the peripheral circulatory system. Velocity is referred to the linear

velocity or rate of blood displacement with respect to time and has properties of distance

per unit time, for example, cm•sec-1. Meanwhile, flow is volume or quantity of blood passing through a certain point in the circulation in a given time and has properties of

volume per unit time, such as, cm3•sec-1 [3, 16].

For a given constant flow, the velocity varies inversely as the cross-sectional area. For

example when 10 cm3•sec-1 of blood passes through a vessel with a cross-sectional area

of 2 cm2, its velocity is 5 cm•sec-1. However velocity drops down to 1 cm•sec-1 when the

cross-sectional area is 10 cm2 and it elevates to 10 cm•sec-1 when the cross-sectional area

is 1 cm2 (figure 3-4) [3]. This relationship is expressed in the equation [16]:

V=Q/A, Where V=velocity, Q=flow and A=area.

Velocity throughout the circulation depends on blood flow. Flow, which

decreases as blood proceeds in the circulation, depends on blood pressure, fluid

properties of blood and dimensions of the circulatory system [25]. Therefore, due to

reduction in flow, blood velocity decreases as it proceeds in the peripheral circulation until blood reaches the capillaries. However, as blood drains out of the capillaries, blood

114 progressively gains velocity again as the venules continue to propel blood toward the venae cavae and the heart (figure 3-3) [3].

Figure LR3.4. Relationship between cross-sectional area, A, and velocity, v, in the systemic circulation. Taken from [6].

3.4. FLOW, PRESSURE AND RESISTANCE

Blood flow (cardiac output) and peripheral resistance interact to create appropriate pressure (force) throughout the vasculature to deliver blood to needed tissues. This relationship is expressed in Ohm’s law [26]:

Q=P/R

Where Q is cardiac output, P is pressure and R is resistance

This law states that cardiac output is directly proportional to pressure and inversely proportional to resistance. The heart pump eventually generates cardiac output whereas resistance depends mainly on vessel shape (length and diameter), blood viscosity, and type of tissue involved [3, 16].

3.4.1. Blood flow Total body blood flow or cardiac output (Q) at rest is about 5,000 ml•min-1 and increases to 35,000 ml•min-1 with a rise in tissue activities (i.e. exercise). Its called cardiac output because it is the quantity of blood pumped by the heart in volume unit per time [15].

115 Laminar and turbulent flow Under normal conditions, the flow of fluids in a cylindrical tube is laminar. The thin fluid

layer adjacent to cylinder wall is extremely slow or even maybe motionless (figure 3-5).

Figure LR3.5. Laminar and turbulent blood flows within a vessel. Emphasized vortices develop from turbulent flow. Taken from [18].

However the fluid layer interior to this external layer must shear against this “motionless” layer and therefore move a bit faster. Similarly layers that are more interior, follow suite and travel more rapidly. Therefore, the velocity of fluid at the center of the laminar is maximal and is twice the mean velocity of flow across the entire cross-section of the tube

(figure 3-6). In laminar flow, substances remain in one lamina as the fluid proceeds longitudinally along the tube [3, 27].

Figure LR3.6. Laminar flow depicts the motion of a series of thin, concentric cylindrical shells sliding over each other. Velocity is higher as the shell position is closer to the central axis. Taken with permission from [27].

Turbulent flow of fluid in a tube however might develop forming eddies and

vortices (figure 3-5). Under these conditions, rapid mixing occurs causing fluid elements

not to remain in one definite laminae. Therefore, turbulent flow requires considerably

116 greater pressure than laminar flow to drive fluid through the same tube. This in turn requires a pump, such as the heart, to work harder to generate a given flow if turbulence develops. Reynolds’ number is a dimensionless number that represents the ratio of inertial to viscous forces for a fluid flowing through a cylindrical tube:

NR=ρDv/η Where D is the tube diameter, v is the mean velocity, ρ is the density, and η is viscosity

When NR <2000, the flow usually is laminar; and it would be turbulent when NR >3000.

Reynold’s number between 2000-3000 is in transition phase where both types of flows

might exist. Thus Reynold’s number principle indicates that large diameter, elevated

velocity, and low viscosity are sources for turbulence flow. Additionally, abrupt

variations in tube dimensions or irregularities in the tube walls may produce turbulence

[3, 27].

3.4.2. Blood pressure Blood pressure is the force exerted against any unit area of the vessel wall by the blood.

The rhythmic contraction of the heart generates this pressure during the left ventricular

contraction and relaxation phases of the cardiac cycle [6]. Blood ejected from the left

ventricle surges into to the aorta creates a pressure gradient throughout the entire vascular

system. A portion of the ejected blood is stored by the aorta, which allows it to stretch

and then recoil during ventricle relaxation, and drives the blood in pulsatile manner. In

preparation for the next contracting phase the left ventricle expands and the left atrium

contracts to drive blood to the left ventricle [4]. A pressure of 50 mmHg means that the

force exerted is sufficient to push a column of mercury up to a level 50 mm high.

However pressure is also measured in centimeters of water. Thus, a pressure of 10

117 centimeters of water means that the force exerted is sufficient to push a column of water

for 10 centimeters. 1 mm Hg = 1.36 centimeters of water [3, 26].

Systolic blood pressure (SBP) During the contraction (systole) phase of the cardiac cycle pressure generated by the

heart at rest is about 120 mm Hg. This measurement is usually referred to as pressure in

the brachial artery at the level of the heart. Additionally, it is an estimate of the work

generated by the heart and of the strain against the arterial walls during ventricular

contraction [26]. Following contraction, the ventricle relaxes, however the elasticity of

the arterial system maintains a driving pressure for steady flow of blood to the periphery, until the next surge of blood arrives [3].

Diastolic blood pressure (DBP) Blood pressure in the brachial artery decreases to about 80 mm Hg during the relaxation

(diastole) phase of the cardiac cycle [26]. The recoiling aorta and the contraction of the

atrium maintain this pressure. Diastolic blood pressure is indicative of peripheral vascular

resistance especially at the level of arterioles and capillaries [3].

Mean arterial pressure (MAP) Due to a longer diastolic phase, the mean pressure across the circulation is not simply an

average of systolic and diastolic pressures. However it is estimated from the following

equation [3]:

MAP = DBP+[(SBP-DBP)/3]

This formula provides an optimal estimation of MAP during rest. However when the

intensity of physical stress changes (e.g. exercise or sleeping) and the length of the

diastolic period is altered (e.g. an increase or decrease in heart rate), accuracy of the

estimation is changed [3].

118 Shear stress Shear stress is the external force applied against the arterial wall as blood is propelled down the arterial tree. This stress is the product of blood viscosity and shear rate as it would increase when viscosity and/or velocity of blood increases [28]. The stress of shear produces a differential motion in the layers of the flowing blood through the circulatory system thereby blood exerts shearing stress on the surface of the arterial wall in contact with blood. Shear stress is essential for the homeostasis of the vascular system as it stimulate the release of vasoactive agents, influence vessel wall structural remodeling and plays a role in the pathogenesis of the vascular system [29].

Volume-pressure relationship The relationship between blood volume and pressure in the peripheral circulatory system is described with volume-pressure curves. Pressure in the arterial and venous systems is influenced by volume of blood within the circulation, however due to differences in wall structure each responds differently to changes in volume [19]. The walls of the arterial system have stronger smooth muscles and are thereby far less expandable than veins.

Therefore slight modifications in blood volume result in huge changes in arterial pressure, whereas slight changes in venous pressure require tremendous changes in volume [22]. For example, mean arterial blood pressure would decrease from 100 to zero mmHg when blood volume drops from 750 ml to 500 ml. On the other hand blood pressure in the venous side decreases from 15 to 5 mmHg as a result of volume reduction of 500 ml (figure 3-7). Volume-pressure relationship is largely influenced by changes in vascular smooth muscle tone (i.e. due to changes in sympathetic drive) (figure 3-7) [3,

20].

119 Blood volume can continuously change within the circulation segments, however

blood pressure integrity must be maintained to assure proper flow of blood to needed

tissues [6]. Delayed compliance is a valuable mechanism by which changes in volume are

accommodated in the circulation. An increase in blood volume will first cause a rise in blood pressure followed by a stretch in the vasculature, allowing pressure to be adjusted

back to normal values. Vise versa, reduction in blood volume will cause a rapid drop in

blood pressure, however the smooth muscle fibers readjust their tension to allow blood

pressure to increase back to proper values. This mechanism is vital when a large loss (e.g.

hemorrhage) or extra increase (e.g. infusion) in blood volume is exhibited within the

circulation [15, 27].

Figure LR3.7. Volume-pressure curves within the circulatory system. Shown the effects of sympathetic stimulation and inhibition. Taken from [3].

Vasculature wall tension Tension in the vascular wall is a function of vessel radius, vessel thickness and

transmural pressure across the vessel. Transmural pressure in turn is the difference

between inside and outside pressure. The exerted force on the inner side of the vessel

wall causes the internal pressure, whereas external pressure is the combined pressure of

120 the vessel-surrounding environment. The relationship between these factors is expressed in LaPlace’s theory, which is [6, 27]:

Pt = Pi - Pe, Where Pt=transmural pressure, Pi=internal pressure, Pe=external pressure

PtR T = ------(LaPlace’s theory) µ Where T=Tension, R=radius, and µ=thickness

LaPlace’s theory indicates; wall tension increases as a result of transmural pressure increase, vessel radius increase or thickness decreases. In a normal vessel, radius and thickness are in approximate proportion, however wall tension differs as transmural pressure changes and the two will tend to stay in reasonable equilibrium (figure 3-8).

Reaching this equilibrium is particularly important to avoid vessel walls from bulging.

For example aortic aneurysm is due to weakening of particular point in the wall of the vessel [27]. Due to increased wall tension, this weak site will soon bulge, causing a decrease in wall thickness and an increase in the radius of the aorta, which will further weaken the aortic wall. This cycle of deterioration might require surgical restoration to avoid further damage or rupture of the wall [2].

Arterial pulse pressure The distensibility of the arterial system plays a role in tissue blood flow during the diastolic phase of the cardiac cycle. After blood surges from the heart with each vertical contraction the aorta expands then recoils to assist pumping blood through the vasculature during diastolic phase of the cardiac cycle, which causes pressure pulsation.

Pulse pressure is estimated as the difference between systolic and diastolic pressure,

121 which would be ~ 40 in young normal adults. However, vascular resistance causes

pressure pulsation to diminish as blood proceeds within the vasculature, thus minimizing

its influence on tissue blood flow. Pulse pressure is influenced mainly by stroke volume

and arterial compliance. For example, a greater stroke volume would require greater

accommodation by the arterial tree thereby greater differences between systolic and

diastolic blood pressures [2, 3].

Figure LR3.8. Pressure-volume-tension relations as described by Laplace's law Left (A); normal vessel. Right (B); aneurysmic vessel due to athrosclerotic injury. Taken from [2].

3.4.3. Resistance Resistance is the combined forces in the circulation that obstruct the flow of blood to the

periphery. Direct measurement of resistance is not available, however it could be

estimated using Ohm’s low (R=P/Q) and expressed in arbitrary unit (U). Blood viscosity,

structure of vessels (diameter, length, and arrangement), and blood flow are factors that

influence vascular resistance, which are included in Poiseuille’s law of fluid conductance

in tubes.

122 Poiseuille’s law In 1846 J.L.M. Poiseuille demonstrated the relationship between steady flow and pressure in a cylindrical tube. He discovered that the relationship between volume flow (Q) and the drop in pressure (P) along a tube of a given length (L) and inner diameter (D) was

[30]:

Q=KPD4/L.

The coefficient K was found to be independent of Q, P, L, and D and changed with alterations in temperature. However, in subsequent developments, Wiedemann in 1856 and Hagenbach in 1860 found that K was related to blood viscosity (K=π/128η); where η is viscosity and expressed in 1 dyn sec.cm2 or poise. Therefore, in the light of these developments, Poiseuille’s law has become [31]:

Q= π∆Pr4/8ηL Where ∆P is the difference in pressure between two ends of vessel, and r is radius

This edition of Poiseuille’s law states that flow is directly proportional to difference in pressure between two ends of a vessel and the radius of that vessel.

Additionally, blood flow is inversely proportional to the viscosity of blood and the length of the vessel (figure 3-9) [27].

123

Figure LR3.9. The relationship between vascular resistance and blood inflow as described by Poiseuille's law in a straight vessel. Taken from [18].

The theoretical derivation of this equation however describes steady laminar flow

in rigid cylindrical tubes and not the pulsatile flow through elastic vessels as in the body.

Additionally, Poiseuille’s law is based on the assumption that each liquid particle moves

at a constant velocity (v) in parallel to the cylinder wall, and that the force opposing this

motion is proportional to viscosity and to the velocity gradient perpendicular to the direction of flow [27].

In the vascular system, length of the vessel plays an important role in determining the magnitude of resistance to blood flow. The longer the vessel the larger the pressure

gradient needed to propel the blood further. As the vascular system progresses through

the body, vessels divide and increase in number forming an in-parallel arrangement

within each vascular bed. The increasing number of vessels causes each segmental cross- sectional area for blood/tissue exchange of materials to increase also. However if this increase in vessel number and cross-sectional area didn’t occur and each microvascular beds were connected end-to-end, an in-series vascular beds or resistance would be formed. The total resistance of these in-series resistances equals the sum of the individual resistances, which adds up to an enormous resistance against flow. Therefore, such

124 arrangement makes it extremely difficult for the cardiovascular system to function given the large pressure needed to drive blood flow through, even for resting condition, let alone under stressful condition (e.g. exercise) [15, 27]. However, typically multiple vascular beds are created from one vessel forming in-parallel resistance (figure 3-10). In this situation, the total resistance equals the sum of the reciprocals of the individual resistances, which is obviously far less than in-series resistance [6].

Figure LR3.10. Vascular resistance in-series and in-parallel circuits. Taken from [18].

According to Poiseuille’s law, a slight change in vessel diameter causes a tremendous change in flow due to the fourth power and is attributed to the physical characteristics of blood flow [27]. In laminar flow, blood velocity is minimum in the layer adjacent to the vessel wall and progressively increases until it reaches maximum at the center of the lamina. In a narrow vessel, where a lot of the blood is very close to the vessel wall and minimum volume of blood is flowing centrally blood would be flowing slowly. In contrast vessels with larger diameter where more blood is flowing centrally, average blood flow would be faster (figure 3-12).

125

Figure LR3.11. A; Relationship between blood flow and vessel diameter. B; Velocities of different concentric layers of blood flow, layers in the center are the fastest. Taken from [3].

For example when a pressure gradient of 100 mmHg is applied to drive blood through three separate blood vessels with diameters of 1, 2 and 4, blood flow will be in these three perspective vessels 1 ml/min, 16 ml/min and 256 ml/min, respectively (figure

3-12) [3]. One third of peripheral resistance is in the small arterioles, which have internal diameter that ranges between 8 to 30 µm with a tremendous capability of changing their diameter by about 4 folds. A 4-fold increase in vessel diameter could increase the flow by as much as 256 folds. This phenomenon allows arterioles to have an important role in vascular resistance and to be very effective in changing blood flow to tissues as a result of central nervous system or local neurohumoral signals [3].

Viscosity of blood also contributes to total vascular resistance. Percent of blood cells and solid matters in blood vessels are called hematocrit and determines viscosity of blood [6]. Hematocrit level usually averages about 40% of blood plasma, which is about

3 times as great as viscosity of water. However an increase in hematocrit to 60 or 70% form a viscosity as great as 10 times of water [3]. Concentration of hematocrit in plasma

126 depends on gender, disease status, degree of bodily activity and altitude at which a person reside. Higher blood viscosity generates greater friction created by dragging blood cells and matters against adjacent cells and vessel wall. This friction is the source of resistance created by viscosity of blood to flow across the circulation [27].

3.5. VENOUS HEMODYNAMICS

According to the Frank-Starling theory of contraction, the heart can pump only what passively fills it during diastole. The venous system operates as a collecting reservoir for blood and controls the quantity of blood returning back to the right atrium. Consequently cardiac output is primary regulated by venous return and the resultant end-diastolic volume by a simple mechanism of filling and stretching the left ventricular muscle fibers

[32]. The Frank-Starling law involving the end-diastolic volume however is not the only factor influences cardiac output. Cardiac contractility, heart rate, and arterial pressure are also important factors [33].

Low intra-vascular pressure, large volume present, intermittent nature of the flow, gravitational effect, venous valves, and the collapsible nature of the venous wall add complicity to flow in the venous cavity. Therefore, there are supplementary factors assisting the flow of blood in within the venous tree [20].

Flow of blood runs with no resistance in large veins when they are distended however as veins enter the thorax and abdominal cavities flow is impeded. Most veins are compressed at many points by the surrounding tissues and the intra-thorax and abdominal pressures, which obstructs flow of blood to the heart chambers [27]. Pressure in the right atrium also hinders blood returning back to the heart from the systemic circulation.

Pressure in the right atrium might sometimes increase above 0 mm Hg (e.g. heart failure),

127 causing blood to backup [32]. However pressure in the venous system doesn’t increase until all collapsed segments between peripheral veins and large veins have opened, which doesn’t occur until right arterial pressure reaches 4-6 mm Hg. A further increase in right atrial pressure would reflect a progressive pressure increase in large veins [33, 34].

However, pressure elevation in the right atrium doesn’t usually occur except under acute circumstances, such as exercise and heart failure [35].

During quiet standing, gravitational effects cause the drainage of thorax blood to the lower parts. Instantly skeletal muscles surrounding the veins start contracting and pushing against the veins to propel blood toward the heart [2, 15]. Meanwhile, the veins refill during muscle relaxation. In this case skeletal muscle is acting as a pump, which is clearly demonstrated during exercise as it plays a more effective role in driving blood toward the heart during exercise. While exercising, larger cardiac output is indeed required for tissue metabolic needs. This is facilitated in part by increasing skeletal muscle contraction force to enhance venous blood flow to the heart thereby increasing end-diastolic volume and cardiac output. However increased venous return and filling capacity is offset due to increased heart rate during exercise [15].

The contraction of the venous smooth muscle also plays a role in driving blood toward the heart, however its contraction depends on the intra-vascular pressure. The venous wall movement is mainly passive at low pressure, however when venous pressure increases above 5 mmHg wall tension increases providing an active role of the venous tone. Venous tone is controlled mainly by central and local neurohumoral signals [2, 27].

While the skeletal muscle pump and venous tone is important to propel blood forward, the venous valves prevent retrograded flow. Following skeletal muscle

128 contraction and propelling blood forward, the valve cusps close and the veins refill

rapidly, which allow for unidirectional flow [13, 20]. However in some venous disorders

(e. g. inflammation of the vein or congenital defects), venous valves become

incompetent. As a result of this venous deficiency valves don’t close properly allowing

for blood to back flow, which tends to increase venous congestion and venous pressure.

Consequently venous return is reduced causing fatigue associated with painful sensation

in the congested areas of the venous system [36].

The intra-thorax pressure varies during quiet breathing from –2 to –4 mm Hg

during expiration to about –5 to –7 during inspiration. The intra-abdominal pressure

decreases when intra-thorax pressure increases, which facilitates venous return form the

intra-thorax veins thereby increase thoracic vena cavae flow of blood and venous return.

The resultant changes in flow inside the veins cause ±25% changes in cardiac output

during respiration. This change in cardiac output is even more pronounced during exercise due to the increase in the rate and force of perspiration [27].

129 SECTION 4. CONTROL OF VASCULAR FUNCTION

Changes in blood vessel diameters are the underlying factor that control blood flows in the arterial and venous systems throughout the circulation. These changes in vessel diameter are the result of contraction and relaxation of vascular smooth muscle. The strategic location and the unique arrangement of the vascular smooth muscle throughout the peripheral circulation allow for phenomenal modulation of the vascular system and subsequently blood flow modifications to and from body tissues (figure 4-1).

The smooth muscle cells wrapped around the vessels of the circulatory system are small, mononucleate, and spindle shaped. They are generally arranged in helical or circular layers around the larger blood vessels and in a single circular layer around arterioles [7, 37]. The vascular smooth muscle cells contain small numbers of thick myosin filaments and large number of thin actin filaments forming sarcomeres with non- visible striations. These filaments are aligned in the long axis of the cell and the sliding filament mechanism operates by means of phosphorylation of cross-bridges to regulate their rate of cycling [3, 7]. Parts of the endothelial cells project into the vascular smooth muscle layer forming myoendothelial junctions at various points along the vessels. These junctions are used for functional communication between endothelium and adjacent vascular smooth muscle (figure 4-1) [7].

In contrast to skeletal muscles, vascular smooth muscles contract very slowly, develops high forces, maintains force for long periods with low adenosine triphosphate

(ATP) utilization, and operates over a considerable range of lengths under physiologic conditions. On the other hand similar to cardiac muscles, cell-to-cell signal conduction is responsible for synergistic contraction smooth muscle [3]. As in skeletal and cardiac

130 muscles, contraction caused by cross-bridge formation is controlled by myoplasmic Ca+2

concentration [38], however the Ca+2 mechanism of action for regulating contraction is

slightly different in vascular smooth. Contraction is elicited by the increase in

myoplasmic Ca+2 that come via voltage-gated calcium channels, through receptor- mediated calcium channels in the sarcolemma, and by the release from the sarcoplasmic

reticulum. Several neurohormonal and metabolic pharmacologic stimuli have been

proposed to be involved in modulating intracellular Ca+2 level including catecholamines

[39, 40], angiotensin II [41], vasopressin [41, 42], histamine [40], serotonin [43],

+ + adenosine [44], CO2 [45], K [46], H [47], nitric oxide [48], and prostaglandins [49].

Figure LR4.1. Typical elastic artery. Shown different segments of a vessel. Taken from [8].

131

Figure LR3.2. Schematic diagram of signal transudation involved in vascular smooth muscle modulation. PKC; protein kinase C, PLC; phospholipase C, G; guanine nucletide binding protein, PIP2; phosphatidyl inositol bisphophate, DAG; diacylglycerol, IP3; inositol triphosphate, Ca+2; Calcium ions, CaM; calmodulin, MLCK; myosin light chain kinase. Taken from [50].

The binding of agonists to vascular smooth muscle membrane receptors activates

phospholipase C (PLC) in a reaction coupled with guanine nucletide binding proteins (G

protein). Phospholipase C hydrolyzes phosphatidyl inositol bisphophate (PIP2) in the

membrane to yield diacylglycerol (DAG) and inositol triphosphate (IP3). Diacylglycerol

+2 +2 and IP3 increase Ca by activating voltage-gated calcium channels and releasing Ca

sarcoplasmic reticulum, respectively. Calcium ions bind to calmodulin, which in turn

binds to myosin light chain kinase. This activated Ca+2-calmodulin-myosin kinase

complex phosphorylates the light chains (20,000 daltons) of myosin. The phophorylated

myosin ATPase is then activated by actin, and the resulting cross-bridge cycling initiates

contraction (figure 4-2). Relaxation on the other hand occurs when the myosin light chain

132 phosphatases dephosphorylate the light chains of myosin and the cytosolic Ca+2 is lowered by sarcoplasmic reticulum uptake and by Ca+2 extrusion by the Ca pump and Na-

Ca exchanger [51, 52].

Vascular smooth muscle is normally tonically contracted and physiological

responses of increased and decreased contraction are a function of the integrated signals

from autonomic nerves, circulating and local hormones, endothelium tissue and local

metabolites [7, 53]. The endothelium, for example, regulates vascular smooth muscle

function by releasing a variety of vasoactive factors [54]. Endothelial cell calcium

regulates the activities of nitric oxide synthase, which metabolizes L-arginine to nitric

oxide, the most important endothelium-derived relaxing factor [55, 56]. After diffusing to

smooth muscle cell nitric oxide exerts its effects by negatively influencing calcium

release from the sarcoplasmic reticulum and its entry through calcium channels, as well as by inhibiting calcium activated contractile process itself [57, 58]. Endothelin-1 is

another endothelial-borne molecule however promotes vasoconstriction by sustaining

elevated levels of Ca+2 and activating muscle cell contractile process [56, 59].

Most of the body arteries and veins are supplied to different degrees by autonomic

nervous system fibers that exert a tonic effect on the blood vessels. Activation of the

sympathetic nerves either directly or reflexly causes smooth muscle contraction or

enhances vascular resistance by increasing cytoplasmic Ca+2 levels. In contrast, the

parasympathetic nerves tend to decrease vascular resistance, but they innervate smaller

number of the body blood vessels. Moreover, changes in the local environment alter the

contractile state of vascular smooth muscle, and alterations such as increased temperature

or increased carbon dioxide levels induce relaxation of this tissue [60, 61]. This section

133 will provide an overview of the mechanisms involved in vascular function modulation under normal conditions. Since vasoregulation is a complex process and include systemic

and local factors, the regulatory effect of each factor will be discussed.

4.1. SYSTEMIC CONTROL OF THE VASCULAR SYSTEM

Systemic modulation is maintained by homestatic reflex mechanisms that receive

mechanical and chemical signals from peripheral sensors within the circulation causing

either a local reaction or activation of the central nervous system [3, 15, 62]. The signals

sent to the central nervous system are processed then forwarded to the command centers

to take appropriate action in the periphery. These command centers send signals to either

the autonomic nervous system or glands around the body. Signals from both arms of the

systemic vascular control are reflected on total peripheral resistance, blood pressure and

blood flow [3, 62].

4.1.1. Nervous Regulation of the Circulation Neural control of the cardiovascular system affects mainly global functions such as blood

redistribution (blood shunting), cardiac output, metabolic adjustments and particularly

important for maintaining constant pressure within the circulation [15]. This control is

inflicted on the vascular system entirely by the autonomic nervous system. Both branches

of the autonomic nervous system innervate the heart however the vasculature has

extensive sympathetic network whereas the parasympathetic innervation is minimum.

Thus the sympathetic nervous system is very important in the control of the heart and

vasculature whereas the parasympathetic system is particularly essential for the

regulation of the heart and cardiac output with some control to the vasculature (figures 4-

3 and 4-4) [3, 32, 62].

134

Figure LR4.3. Functional anatomy of the sympathetic nervous control of the circulation. Taken from [3].

The most remarkable characteristic of the autonomic nervous system is probably the ability to modify heart rate, vascular resistance and blood pressure within few seconds. For example the sympathetic nervous system double heart rate within 3-5 seconds whereas blood pressure is increased to double normal within 15 seconds [15, 63].

This exceptional capability is due to the sophisticated and extensive neural network that connects the central nervous system to the heart and almost all segments of the vasculature (figures 3-3 and 3-4) [61, 64]. The innervation of the small arteries and arterioles allows the sympathetic stimulation to increase vascular resistance, blood pressure and blood flow to areas where needed. Whereas the innervation of the large vessels, particularly the veins, assists in propelling blood from these vessels to the heart

135 thus generating larger cardiac output [20]. Additionally sympathetic activities are especially important to increase heart rate and contractility force [61, 64]. Although, due to minimum vascular intervention, the parasympathetic stimulation is not particularly important for vasoregulation it plays a crucial role in the reduction of heart rate and contractility thereby blood pressure [65].

.

Figure LR4.4. Sympathetic innervation of the systemic circulation. Taken from [3].

The sympathetic vasomotor nerves project their fibers from almost all the thoracic and two lumbar spinal nerves. These fibers connect to the heart and its vessels and to the peripheral vessels (figures 4-3 and 4-4). Additionally a sympathetic nerve fiber branches of the intermediolateral horn cells of the spinal cord passing through the sympathetic chains and the splanchnic nerves directly into the adrenal medulla, which is responsible for the secretion of epinephrine and norepinephrine into the blood stream [8, 66]. The signals transmitted via the sympathetic fibers are initiated in the vasomotor center located in the pons and medulla (figure 4-5). The vasomotor center is controlled by locations within the central nerves system including the hypothalamus and cerebral cortex [8].

Within the vasomotor center there are areas responsible for initiating vasoconstriction

136 whereas others initiate vasodilation. Generally however increased sympathetic nervous system activities induce strong vasoconstriction throughout the circulation except in the heart and skeletal muscle vasoconstriction is less pronounced [15, 67, 68]. In addition to vasoconstriction and vasodilation signaling the vasomotor center is also responsible for receiving and coordinating reflexes from the sensory receptors around the circulation

(e.g. barareceptors and chemoreceptors). This particular function orchestrates a rather sophisticated feedback loop between the periphery and the central nerves system [65, 69].

The vasoconstrictor area, within the vasomotor center, fires vasoconstriction impulses to vessels around the circulation to form a sympathetic vasoconstrictor tone. These signals are to maintain a vasomotor tone that is important to preserve adequate blood pressure and flow. At rest this tone allows for partial state of contraction of blood vessels however during stress the firing of vasoconstriction impulses increases causing a corresponding increase in blood pressure (figure 4-5) [15].

The sympathetic terminal nerve endings are called adrenergic nerve fibers because they secrete the neurotransmitter norepinephrine. The manufacturing of norepinephrine is initiated in the axoplasm of the terminal nerve endings of adrenergic nerve fibers but is completed inside the vesicles [66, 69]. Once released norepinephrine binds to cellular membrane receptors called alpha (α) and beta (β) adrenergic. These receptors are highly specific, located in the vascular smooth muscle cell and penetrate through the cell membrane [8]. Upon norepinephrine binding to the receptor it cause conformational changes to the penetrating receptor thereby triggering cellular activities within the vascular smooth muscle cell by either increasing membrane permeability to ions or activate/deactivating enzymatic reaction (s) [3, 66]. In general α-adrenergic

137 receptors cause vasoconstriction β2 receptors cause vasodilation whereas β1 cause

increase in heart rate and contractility. The skin and internal organs vasculature has

mostly α receptors, skeletal muscle and splanchnic vasculature contains both α and β2

receptors however α are more abundant and dominant than β2 whereas the heart muscle

contains β1 and the heart vessels contain α and β2 receptors [3, 63]. The action of β2

receptors has probably been most confusing since it induces vasodilation. When

norepinephrine infused into tissues contain α and β2 receptors it activates only α

receptors, whereas β2 receptors are activated only when α receptors are blocked.

However norepinephrine released from nerve endings does not activate β2 receptors even

if α receptors are blocked. [15]. Therefore increase sympathetic nervous system activities

increase heart rate and contractility, heart and skeletal muscle vasoconstriction and strong

skin and internal organ vasoconstriction [70]. Although some reports indicated β2

receptors’ action might contribute to initial vasodilation observed at onset of exercise, skeletal muscle vasodilation is sustained by local mechanisms [15]. Since the autonomic nervous system is responsible for the flight-or-fight response, these cardiovascular adjustments are particularly important for the initial increase in blood delivery to the skeletal muscles during times of physical stress such as exercise [15, 70]. Dilation of the eye pupils, increased activities of the sweating glands, and increased metabolism are also implicated in the flight-or-fight response and stimulated by the sympathetic nervous system [1, 71].

138

Figure LR4.5. Systemic circulation regulatory areas within the brain. Taken from [3].

The removal of norepinephrine is accomplished by either reuptake by the nerve endings, diffuse to the circulation or destroyed by the enzyme monoamine oxidase or one of its derivatives. The reuptake mechanism accounts for 50-80% of the unused norepinephrine whereas norepinephrine diffuses to the blood accounts for most of the reminder. Finally norepinephrine destroyed by monoamine oxidase accounts for only a small fraction of the reminder [3, 69]. Norepinephrine released directly into the tissue stays active for only few seconds before its removal by one of the mechanisms. However norepinephrine released into the blood stream by the adrenal medulla or escaped from the nerve terminal stays active for 10-30 seconds until it diffuses into some tissue (e.g. liver) where its destroyed by monoamine oxidase [3, 15].

The parasympathetic nervous system projects its nerve fibers to the heart through the vagus nerve (cranial nerve X). The nerve endings at the heart release the neurotransmitter acetylcholine, which binds to cholinergic receptors, located mostly at the atria [66]. This system serves a double purpose, as it is particularly important when

139 the heart is excited and functioning under great workload. Activation of cholinergic

receptors of the heart decreases heart rate and contractility within a few seconds [1]. In fact a strong parasympathetic stimulus can stop heart pumping for a few seconds and reduce the strength of the heart contraction by 20-30% [3]. The parasympathetic contribution to the control of the vasculature is very limited and confined to a few tissue beds including the brain, heart, erectile tissue of the genitalia, some glands and, to a lesser extend, the kidney [72].

The activities of both branches of the autonomic nerves system are in constant

fluctuation mainly to defend blood pressure within a narrow range however this maintenance of blood pressure is intended only for short term whereas other mechanisms

have to be activated for long-term blood pressure maintenance [27]. The slightest change

in blood pressure is normally sensed by the baroreceptors and chemoreceptors within the

circulation, subsequently they send informative singles to the vasomotor center to unleash

a reflex to restore blood pressure back to normal [6]. This reflex usually includes a

withdrawal and/or activation of either of the autonomic nervous system branches. For

example in the case of elevated blood pressure, the autonomic nervous system deceases

sympathetic activities and increase parasympathetic activities to elicit a decrease in blood

pressure to normal [3].

4.1.2. Systemic Hormonal Control of the Circulation Systemic hormones are released into the bloodstream by the major endocrine glands and

bind to receptors located on the vascular smooth muscle. These hormones include

catecholamines, rennin-angiotensin-system hormones and arginine vasopressin [1, 73].

Hormonal agents released into the blood stream act directly on vascular smooth muscle

causing either dilation or constriction [73]. However hormones might also inflect their

140 effects on vascular tone indirectly vie modulating neural control or other hormonal

mechanistic [62]. For example angiotensin II binds to the neuronal membrane and

modulate the release of norepinephrine [74]. As well, angiotensin II controls blood pressure by modulating salt and water levels in the body thereby plasma volume [75].

Therefore considering the constant communication between the periphery and central

nervous system in addition to the interaction between hormonal and neural control, the

action of hormones in the body is often multifaceted and highly complex [62].

Catecholamines Due to a lot of similarities the adrenal medulla serves as an extension to the sympathetic

nerves system. In addition the sympathetic nerves system activities at the adrenal

medullae stimulate the release of large quantities of epinephrine and norepinephrine into

the blood stream. However catecholamines effects last 5 to 10 times longer as a result of

slower removal of the hormones [3]. Similar to the sympathetic nerves activities,

circulating catecholamines increase when the body is stressed mentally or by hunger,

exercise, or hypothermia. In addition they contribute to the fight-or-flight response as

they have profound and rapid effect on the heart, vascular system, and metabolism [15].

Both catecholamines hormones are derivatives of the amino acid tyrosine, have the same production procedures, and similar structure [18]. In addition at the same sites both epinephrine and norepinephrine have the same effect however mole for mole epinephrine is several times more powerful than norepinephrine. On average 80-90% of the adrenal medulla secretion is epinephrine whereas 10-20% is norepinephrine. Both catecholamines are extremely important in blood shunting and maintenance of adequate blood pressure thereby assuring sufficient blood delivery and removal during times of rest and stress

(e.g. exercise) [3, 76].

141 Epinephrine acts on the peripheral arteries by activation of α and β2 adrenal receptors throughout the body. Alpha-receptors cause vasoconstriction whereas β2- receptors induce vasodilatation. The concentration of these receptors is more abundant and dominant in areas than others. For example α is more abundant and predominant than

β2 receptors in the skeletal muscle and splanchnic arteries whereas β2 don’t exist in the skin and internal organs arteries [3]. Alpha-receptors on the other hand are highly concentrated in the skin and internal organs. Therefore increase in circulating epinephrine, such as during exercise, results strong vasoconstriction throughout the body except the skeletal muscle vasoconstriction is moderate [3, 76]. Increase in epinephrine to concentrations of 100 pg ml-1 induces elevation in heart rate and systolic blood pressure

-1 increases at concentrations of 125 pm ml . Activation of β2 causes vasodilatation followed by a drop in diastolic blood pressure in skeletal muscle however this drop doesn’t occur until epinephrine reaches 200 pg ml-1 which is ~ 135 pm ml-1 higher than basal levels [76].

In the venous side of the circulation existence and action of β2 and α adrenal receptors are not clear yet in human. However splanchnic and cutaneous veins contain α adrenal receptors and were found to induce venoconstriction. On the other hand the action of β adrenal receptors in veins is not known yet in human [77].

Norepinephrine is considered a neurotransmitter since most of it is derived from the sympathetic nerve terminals in the body [76]. However as a result of high sympathetic nervous activities its blood concentration might reaches four or five times above normal values (~1.8 ng ml-1) [15]. At these high levels norepinephrine acts as α

142 adrenal receptor hormone and induces vasoconstriction. In areas where α and β2 adrenal

receptors present vasoconstriction occur since β2 has low affinity to norepinephrine [3].

Angiotensin II The action of angiotensin has been revealed in the 1950s as the agent responsible for the

pressor responses observed following intravenous injection of renal extracts [62].

However aside from its capability to control arterial pressure through changes in vascular

tone, angiotensin inflect its effect on blood pressure by several mechanisms involves

controlling fluid volumes and interaction with the central and peripheral nerves systems,

thus its essential for long-term pressure control [78].

The production of angiotensin II is triggered by peripheral signals initiated by

reduction in blood pressure [67]. This reduction in blood pressure provoke the conversion

of prorenin to renin within the juxtaglomerular cells in the walls of the afferent arterioles

immediately proximal to the glomeruli in the kidney [79]. Following its release into the

blood stream the enzyme renin converts angiotensinogen to angiotensin I, a mild

vasoconstrictor with minimum effect in circulatory function. Renin persists in the blood

stream for 30 to 60 minutes as it continues forming angiotensin I [79, 80]. After few

seconds of producing angiotensin I, angiotensin-converting enzyme (ACE) converts it to

angiotensin II in the small vessels of the lungs. Angiotensin II persists in the blood for

only one or two minutes, then it gets converted to the inactive form by the enzyme

angiotensinase [80].

Angiotensin II powerful vasoconstriction capability is probably the main

mechanism by which it effect vascular resistance and blood pressure [81].

Vasoconstriction occurs mainly in the arterial side and to a lesser extends in the venous

side. Contraction of the arterial smooth muscle cells generates increase in vascular

143 resistance and blood pressure, whereas the mild contraction of the venous smooth muscle cells causes increase in venous return and cardiac output to help the heart pump against the angiotensin II-induced increase in blood pressure [15, 81].

Angiotensin II is not held to its vasoconstriction capability but also affects blood pressure by changing plasma volume. Angiotensin II acts directly on the kidney causing retention of salt and water and indirectly by inducing aldosterone secretion from the adrenal glands [62]. The effect of angiotensin II on fluid volume is particularly important for long-term control of arterial blood pressure [64]. High concentration of angiotensin II constricts the renal blood vessels thereby diminishing blood flow through the kidneys as a result, less fluid filters thorough the glomeruli into the tubules. Additionally the reduced blood flow into the peritubular capillaries reduces pressure causing rapid osmotic reabsorption of fluid from the tubules [3]. Angiotensin II also directly affects the tubular by increasing reabsorption of salt and water. The combined effect of these three mechanisms result in decreasing urinary output as much as four- to six-fold.

The reduction in kidney blood flow causes the release of aldosterone. Aldosterone increases sodium reabsorption by the kidney tubules, which increases the extracellular fluid sodium thereby increase in fluid retention. The total effect of direct and indirect mechanisms of fluid control is crucial for long-term control of arterial blood pressure.

However to the opposite of the popular belief, angiotensin II-induced retention of fluid by the direct-effect mechanism is probably three or more times as potent as the indirect effect [3].

Angiotensin II presser response vie vascular and fluid control has been confirmed for decades. However research have also revealed that angiotensin II facilitate the release

144 of sympathetic nerves system neurotransmitters in the periphery and adrenal medullary

catecholamines in addition to its possible influence on the central nervous system [62].

The first observation of the central effect of angiotensin II was observed when a dog head

was isolated from its circulation and connected to another animal while the brain

remained connected to the body only by the spinal cord. Following angiotensin II

infusion in the donor animal arterial pressure was increased in both animals. This presser

effect was inhibited by blocking the sympathetic activities, which indicate the effect was mediated by sympathetic nerves system [82]. Arterial pressure increased also in rabbits following vertebral arterial infusion of angiotensin II and failed to rise following systemic infusion confirming the central role of angiotensin II [83]. However the angiotensin II- induced augmentation of blood pressure in greyhounds was greatly hampered by atropine

(parasympathetic antagonist) or vagotomy indicating parasympathetically mediated response [84]. This difference might be due to species differences given the autonomic nerves system in greyhounds is predominantly vagal. The central site for angiotensin II effect is not clearly determined. However since angiotensin II doesn’t cross the blood brain barrier the site of action must be reachable from the circulation. Several investigations have found traces of angiotensin at varies location within the circumventricular organs including postrema area of the hind brain, the subfornical organ in the forebrain and the organum vasculosum of the lamina terminalis [85, 86].

The high level of interaction of these mechanisms demonstrates the multiple and complex action of angiotensin II in the control of blood pressure. Discovering the interaction of angiotensin II with several body systems have propelled the research related to hypertension and therapy pertain to (e.g. ACE inhibitors).

145 Arginine vasopressin (antidiuretic hormone) Arginine vasopressin is a nonapeptide hormone released by the posterior lobe of the

pituitary gland. Increased osmolality (increased plasma concentration) stimulates

osmoreceptors near the hypothalamus, which send singles to posterior pituitary to release arginine vasopressin into the circulation [87]. However reduction in central blood volume

as a result of fall in blood volume (hemorrhage) [88], orthostatic and/or positive pressure

breathing, are also potent stimuli of arginine vasopressin [89, 90]. Whereas mechanisms

by which either osmolality is decreased or central blood volume is increased inhibit the

production of the nonapeptide [91].

Arginine vasopressin receptors are located on the peritubular surface of the cells

of the distal convoluted tubules and medullary collecting ducts in the kidney. The binding

of the hormone to these receptors stimulates adenylyl cyclase activity and form cAMP

[92]. Thereafter cAMP initiates protein kinase activity and a phosphorylation cascade

resulting in the insertion of the protein, aquaporin, in the luminal membrane causing

antidiuretic action, the main effect of the hormone. Vasopressin inflicts its effect on the

kidney by increasing fluid permeability of the collecting ducts and tubules allowing water

to be reabsorbed back to the body [64].

In rats arginine vasopressin is several folds more potent vasoconstrictor than

angiotensin II, on a mole-for-mole basis [64, 93]. However the vasoconstrictor effects of

arginine vasopressin in human appear to be pathological rather than physiological as

vasopressin plasma levels are unlikely to reach those levels required to achieve

vasoconstriction [94]. At lower concentration of arginine vasopressin the contractile and

pressor responses to adrenergic and neural stimuli appear to be augmented [95].

146 Vasopressin once released remains in the plasma for as little as 2-3 minutes before

removed by the kidney and liver.

4.2. THE REFLEX CONTROL

The circulation is a mechanical feedback system thus any alteration in pumping or

peripheral vascular tone will influencethe intravascular pressures in various sections of

the circulation [96]. Continuous information about the state of the circulation is

transmitted from different sensory receptors strategically located around the circulation.

Any disturbances has direct effects on the heart and blood vessels which will tend to alter

cardiac output and tissue blood flow through local regulatory mechanisms [97]. These

peripheral sensory receptors are the baroreceptors and chemoreceptors.

4.2.1. Baroreceptors Baroreceptors are found in almost every large artery in the thoracic and neck areas

however they are found in abundant in the carotid artery and aortic arc. These sensors are

stimulated when these arteries are stretched [98]. Signals from the carotid receptors are

transmitted vie the Hering’s nerve that connects to the glossopharyngeal nerve, then to

the tractus solitarius in the medullary area, whereas the aortic baroreceptors connect with

the tractus solitarius by the vagus nerves (figure 4-6) [3, 9, 65].

The baroreceptors appear to have pressure-dose-response [3, 9]. The aortic

baroreceptors are stimulated at pressures around 30 mmHg whereas the carotid ones start

sending signals to the brain when pressure decrease below 60 mmHg and their activities

progressively intensify as they reach maximum at 180 mmHg. However the normal operating pressure is around 100 mmHg and a slight change in pressure calls for the autonomic nervous system for adjustments. The remarkable characteristic of the baroreceptors is the extremely rapid respond to changes in blood pressures. Additionally

147 constantly and rapidly increasing blood pressure triggers baroreceptors’ response far more rapidly than stationary pressure [3, 99]. For example a progressively increasing blood pressure of 150 mmHg cause twice as much increase in the baroreceptors’ firing rate as a stationary blood pressure of 150 mmHg. Whereas a falling pressure cause one quarter firing rate of that for the stationary pressure [3, 100]. Once the signals are received in the vasomotor center the sympathetic vasoconstrictor tone is inhibited and the vagus tone is activated (figure 4-5). This would decrease cardiac output and vascular resistance followed by blood pressure decrease. Additionally the baroreceptors firing rate increases during systolic and decreases during diastolic phases of the cardiac cycle [3,

15]. The magnitude of sympathetic tone inhibition depends on the firing rate from the arterial baroreceptors where higher firing rate cause more inhibition. Therefore a decrease in the firing rate, such as during a fall in blood pressure, cause an increase in sympathetic tone thereby elevation in blood pressure. This is apparent when blood pressure in the head and upper portion of the body falls upon standing up after laying down the baroreceptors elicit a reflex that triggers a strong and immediate sympathetic discharge result in rapid restoration of normal blood pressure [15, 68]. The baroreceptors serve as a short-term mechanism to defend blood pressure within a narrow range. However for long-term regulation of blood pressure other control systems, such as antidiuretic and associated hormonal systems, have to get involved [9, 98].

4.2.2. Chemoreceptors The chemoreceptors are highly vascularized, extremely sensitive to changes in blood gas proportion, and easily accessible by the blood [68]. The bifurcation of the common carotid artery and the aorta are the host of these really small (1-2 millimeters) organs.

Each common carotid artery harbor one sensor called carotid bodies whereas several

148 similar receptors, called the aortic bodies, cling to the wall of the aorta (figure 4-6) [9,

67]. The accessible location of these nerve endings allows for constant sensing of changes in blood gas continent, such as lack of oxygen, excess carbon dioxide and excess hydrogen ions. Additionally the small nutrient artery runs through these bodies supply them with abundant blood flow and keep them with close contact with arterial blood [3,

101]. Below a critical value fall in blood pressure changes blood flow in the nutrient artery, which results in a diminished oxygen and accumulation of carbon dioxide contents. Consequently these sensors are stimulated and start firing signals to the vasomotor center in the central nervous system [15, 101]. Within the vasomotor center these signals are processed and a stimulus is transmitted to the sympathetic nervous system to counteract this drop in arterial blood pressure. However these sensors play an important role in preventing a fall in blood pressure below 80 mm Hg whereas they are not acutely activated during normal blood pressures [3, 9, 101].

Figure LR3.6. Peripheral baroreceptors. Taken from [3].

149 4.3. LOCAL REGULATION OF VASCULAR TONE

The cardiovascular system delivers blood throughout the body according to tissue demands. This process involves constant adjustment of the heart and the vessel functions to assure sufficient blood pressure and volume. However an integral part of this adjustment occurs primarily in the capillaries, arterioles, venules, small arteries and small veins. This group of vessels is controlled by chemical substances mainly produced in the tissues surrounding theses vessels. Additionally these chemical substances are produced in accordance with the demands of the metabolic tissues. This mechanism is the local regulation of vascular function and involves mainly the contribution of the endothelium and metabolic tissues.

4.3.1. Endothelium-Dependent Regulation of Vascular Tone The endothelium is a monolayer of squamous cells located strategically at the interface

between the blood and vascular muscle cells implying an intimal and direct interaction

between these three compartments [56, 102, 103]. The endothelium is a very interactive

tissue capable of responding to numerous mechanical, chemical and cellar stimuli. The

surface of the vascular endothelium expresses a verity of receptors for potent vasotonic

agents that have rapid and profound effects on the vascular wall function.

Recent advances in cardiovascular science have demonstrated the integral role of

endothelium cell in the cardiovascular function beyond just a passive barrier of the

vasculature. Since the illustration by Furchgott and Zawadzki of the obligatory role of the

endothelium in the relaxation to acetylcholine in the isolated aorta of the rabbit [104], it

has become obvious that the endothelium modulates the degree of constriction of the

underlying vascular smooth muscle. It does so by liberating vasoactive substances

including endothelium-dependant relaxing factors [104, 105], endothelium-dependant

150 hyperpolarizing factors [106-108], and endothelium-dependent constriction factors [59,

109].

It is also clear now that the endothelium possesses a variety of modulatory

functions beyond the control of vascular tone including endothelium-dependent

modulation of platelet aggregation [110] and endothelium-dependent modulation of

vascular growth [111]. The endothelium tissue is also involved in vascular function

alterations associated with the pathogenesis of many cardiovascular diseases [112].

Endothelium-dependent relaxation Among a number of vasotonic factors, the endothelium lining of the vascular system

releases relaxing mediators including nitric oxide, prostacyclin, and hyperpolarization

factors. In addition to smooth muscle level of responsiveness, these vasodilatory agents

play an integral part in peripheral vasodilation at rest and under various stress conditions.

In a lot of cases these agents interact together to achieve proper peripheral blood flow.

However paradoxically Ca+2 ion, the very same agent that promote vasoconstriction, is

also involved in vascular smooth relaxation. All endothelium-dependent relaxation factors are released as a result of increasing Ca+2 concentrations in the endothelial cell.

This increase is triggered by either agonists (e.g. acetylcholine, histamine, bradykinin,

arginine vasopressin and ADP) or sheer stress (figure 4-7).

Nitric oxide Nitric oxide is a substance released continuously by the endothelial lining of the vascular

system and prominently directed to induce vasodilation [48]. Additionally it is currently

considered as one of the most endothelial function markers as its release believed to

induce a compensatory relaxing response to a stimulus of a constricting nature [112] and

prevent platelet aggregation [113]. Due to its gaseous nature, nitric oxide is rapidly

151 produced, metabolized and highly diffusible without even direct contact with vascular

smooth muscle, which makes it hard to detect it [104]. However using invasive and

difficult techniques, animal models have allowed researchers to reveal a lot of

information to help us understand the important role of nitric oxide [55, 114].

Furthermore, pharmaceutical infusion of substances to modify molecular markers of nitric oxide function have been used as simple and noninvasive method to assess nitric oxide production and synthesis in human [115, 116].

Interestingly the role of nitric oxide was discovered as a result of searching the

effect of acetylcholine. For a long time the role of acetylcholine in vascular tone was a

subject of controversy. Some investigators noted acetylcholine effect ranged from slight

to substantial vasodilatation, [117-119] whereas the majority of the literature reported

only additional contraction [120]. However the importance of an intact endothelium for

the acetylcholine-induced vasodilation in rabbit aorta rings was not demonstrated until

the accidental finding by Furchgott and Zawadzki in 1980 [104]. A failure to follow the

standard procedures for the effect of drugs on vascular smooth muscle tone resulted in

preserving the intimal surface (endothelium tissue) of the vascular smooth muscle and

subsequent muscle relaxation to acetylcholine. These unexpected results were followed

by another experiment in which extra care was taken during vascular smooth muscle

preparations to avoid rubbing the endothelium tissue. In this experiment a transverse strip

of rabbit aorta with endothelium removed was mounted between spring-clips with a

longitudinal strip of the same length and width but with endothelial cells present, intimal

surface against intimal surface. In this so-called “sandwich” preparation, the

endothelium-free transverse strip, which alone did not relax in response to acetylcholine,

152 exhibited good relaxation. This relaxation was attributed, at the time, to the release of

substance(s) from the endothelial cells of the longitudinal strip, and called endothelium-

dependent relaxation factor [104]. Independently Ignarro and coworkers in their

continued study of nitrovasodilator action discovered that nitric oxide could relax

vascular smooth muscle through stimulation of soluble guanylate cyclase [121].

Afterward the findings of these two groups were matched to conclude that the

endothelium-dependent relaxation factor was indeed nitric oxide [122, 123].

Like the rest of endothelium-dependent relaxation factors, production of nitric

oxide is initiated as a result of increasing Ca+2 ions concentration in the endothelium cell.

Epinephrine, arachodonic acid, acetylcholine, ADP, ATP, arginine vasopressin,

bradykinin (AVP), endothelin, histamine, serotonin, thrombin, in addition to sheer stress

are factors responsible for this increase in Ca+2 ions.

The binding of one of these agents or the physical pressure of sheer stress trigger

intracellular signals (e.g. cyclin adenosin monophosphate, phospholipase C, inositol

triphosphate and diacylglycerol) causing an increase Ca+2 ions within the cell. This

increase in Ca+2 activates nitric oxide synthase, which reduces L-arginine to nitric oxide

and L-citrulline [125-127].

Following the release of nitric oxide from the endothelium cell, it easily diffuses

to the adjacent vascular smooth muscle cell and produces vasodilation. This vascular

relaxation occur as a result of nitric oxide binding to the heme portion of guanylyl

cyclase in muscle cells and increase production of cyclic guanosine monophosphate

(cGMP) from guanosine triphosphate. The accumulation of cGMP activates cGMP-

dependent protein kinase, which phosphorylates myosin light chain kinase and regulates

153 activity of action-myosin ATPase (figure 4-8). Phosphorylation of myosin light chain

kinase lowers its affinity for the calcium-calmodulin complex, thus decreasing

phosphorylation of myosin light chain and promoting the formation of comparatively less

active, dephosphorylated form of myosin. Activation of cGMP-dependent protein kinases

can also result in the phosphorylation of calcium transporters, complexing intracellular

calcium and allowing relaxation of vascular smooth muscle [128]. However other

investigators have suggested that cGMP-mediated relaxation might involve other

mechanisms. These mechanisms include inhibition of a sarcolemmal calcium channel to

inhibit Ca+2 influx [57, 129], or the activation of sarcolemmal calcium extrusion pump

[130].

Figure LR4.7. Agents cause the release of endothelium-derived relaxing factors. NO; nitric oxide, PGI2; prostacyclin, AVP; arginine vasopressin, NA; norepinephrine. Taken from [124].

154

Figure LR4.8. Effect of endothelium-derived relaxing factors. NO; nitric oxide, PGI2; prostacyclin, R; receptor, cGMP; cyclic guanosine monophosphate, cAMP; cyclic adenosine monophosphate. Taken from [124].

Nitric oxide is also involved in deactivation of platelets aggregation. This process

is mediated by increasing intercellular levels of cGMP [58]. However the effect of nitric

oxide is strongly synergistic with that of prostacyclin. Thus the combined secretion of

nitric oxide and prostacyclin by the endothelin cells at the interface with the blood must

constitute a powerful antiaggregatory barrier (figure 4-8) [131].

Nitric oxide-independent relaxation maybe reduced by hemoglobin or methylene

blue, which antagonize the rise in cGMP either by inhibiting guanylate cyclase or by

scavenging nitric oxide and preventing its stimulation of the enzyme [54, 56].

Infusion of 0.8 mg ml-1 of nitrates was enough to produce venodilation but 3-4 mg

ml-1 was required to produce dilation in the arterial side in pre-constricted vessels [132],

whereas nitroglycerin produced venodilation even in unconstricted veins [133]. The effect of nitric oxide however varies in different segments of the venous system. In a study to compare the effect of nitric oxide in different veins found that acetylcholine

produced significant dilation in the external jugular, superior vena cava, brachiocephalic

veins and two portions of the inferior caval vein. Whereas the azygos, pulmonary,

155 splenic, renal, femoral and lateral saphenous veins exhibited 10-20% relaxation. On the

other hand the portal, mesenteric and the third segment of the inferior caval veins showed

constriction [134]. Similar observations were also found in human [135], as 0.4 of

nitroglycerine had no effect on superficial hand veins whereas produced dilation in the

calf and forearm veins.

Prostacyclin Physiologically, prostacyclin is locally produced and serves a double purpose as it causes

vascular smooth muscle relaxation and prevents platelets aggregation. In addition it is

responsible for metabolizing cholesterol esters in smooth muscle cell, prevent the

accumulation of cholesterol esters and suppress the release of vascular smooth muscle growth factors [105, 136]. Arachidonic acid is a membrane-bound fatty acid processed by cyclooxygenase into endoperoxides then to prostacyclin by prostacyclin synthase. This process is initiated by the break down of cell phospholipids by phospholipase A2 within

the endothelium cell then degraded to 6-keto-PGF1α, a biologically inactive form of

prostaglandin [137]. Amongst several substances proposed, bradykinin and choline esters

seem to be the strongest candidates to stimulate prostacyclin formation and release in

endothelium cells of human and animals [138, 139]. Angiotensin I and II also participate

in the production process, however they don’t seem to be as potent generators as

bradykinin and choline esters [139]. Nicotinic acid derivatives and the sheer stress of

blood on vessel walls have also been suggested to mediate the release of prostacyclin in

human and animal endothelium cells [140-142]. Moreover histamine, arginine

vasopressin, platelet activating factor, interleukin-1, leukotrienes and thrombin are

believed to be involved in the production of prostacyclin [31]. However regardless of the

external stimulus, prostacyclin biosynthesis is triggered by activating membrane G-

156 protein and increasing cytoplasmic Ca2+. This increase in [Ca2+] activates phospholipase

A2, the initial rate-limiting enzyme in prostacyclin production pathway (figure 4-8) [105].

Glucocorticosteroids and lipcortin inhibit the production of prostacyclin via

inhibiting phospholipase A2 [143, 144]. Whereas low-density lipoproteins [145] and

vitamin K [146] inhibit prostacyclin biosynthesis by interfering with prostacyclin

synthase function because they produce oxygen free radicals. On the other hand aspirin

and aspirin-like drugs affect prostacyclin production by blocking the activities of

cyclooxygenase thereby inhibiting the conversion of arachidonic acid to endoperoxides

[147].

Following binding to the receptors prostacyclin induces relaxation and prevents

platelet aggregation by increasing cAMP activities in vascular smooth muscle and in

platelets [105]. However some investigators suggest the contribution of prostacyclin in vascular function is not as significant as nitric oxide and might be considered negligible

[148].

The effect of prostacyclin on venous smooth muscle is a disperse, mainly due to minimum information available. Sinzinger and Fitscha questioned venodilation to prostacyclin as they found venoconstriction in human saphenous veins [149]. In general however prostacyclin was found to produce venous smooth muscle relaxation [150, 151].

For example in a study by Hiremath et al. prostacyclin produced dose-dependent relaxation [150] confirming previous observations by Robinson and colleagues [152].

Endothelium-dependent hyperpolarization factor In 1978 Kuriyama and Suzuki reported hyperpolarization of the cell membrane of smooth

muscle when exposed to acetylcholine [153]. Bolton et al. verified this finding in 1984

while hyperpolarization was converted to depolarization as a result endothelial tissue

157 removal suggesting the involvement of the endothelium tissue [154]. Bioassay experiments thereafter revealed that the endothelium indeed releases highly diffusible substance(s) causing hyperpolarization followed by smooth muscle relaxation [107].

Despite infusing endothelium-dependent relaxant inhibitors (e.g. methylene blue and hemoglobin) [155, 156] or exposing the endothelium tissue to substances don’t cause the release of nitric oxide or proctacyclin (e.g. substance P and oxotremorine) [106, 157] hyperpolarization-induced relaxation was maintained indicating additional endothelium- dependent relaxation mechanism other than nitric oxide and prostacyclin.

Several substances have been identified to cause the release of endothelium- dependent hyperpolarization factor. These substances include acetylcholine, the primary smooth vessel dilator, in addition to histamine [155] bradykinin [158, 159], substance P

[160] [107], calcitonin gene-related peptide [161], ADP [162], endothelin [163-165] and vasoactive intestinal polypeptide [160]. As nitric oxide, endothelium-dependent hyperpolarization factor is released as a result of an increase in free [Ca+2] within the endothelium cell. The initial response to a stimulus can be attributed to the release of intercellular endothelial calcium stores, whereas the entry of extracellular calcium into the cell prolongs the response to acetylcholine or bradykinin [166, 167]. Additionally endothelium-dependent hyperpolarization apparently involves calcium calmodulin, as inhibitors of this enzyme decrease hyperpolarization in porcine coronary artery to bradykinin (figure 4-8)[168, 169].

Endothelium-dependent hyperpolarization of smooth muscle is induced by an increase in K+ conductance of the cell membrane, which leads to relaxation by increasing the membrane potential of the muscle cell [154]. This has been evident as membrane

158 conductance of K+ in smooth muscle is increased during hyperpolarization [153, 154,

170], the magnitude of hyperpolarization is inversely related to [K] [170], radioactive

rubidium (marker of K+ activities) efflux is increased during hyperpolarization [171,

172], hyperpolarization has been inhibited by K+-channels blockers [173], acetylcholine

produces a transient increase in the K+ efflux, associated with the transient

hyperpolarization of the membrane [171] and K+-channels openers have caused

hyperpolarization-induced relaxation [170]. Potassium channels have been proposed to

account for most of the potassium trafficking across the smooth muscle membrane

because resting membrane potential relay mostly on the membrane potassium gradient.

This has been more evident in large conduits than small ones [174]. Moreover, several

investigations have demonstrated that endothelium-dependant hyperpolarization is

insensitive to ATP-sensitive potassium channel inhibitors, blocked by small conductance

calcium-activated potassium channel and voltage-dependent potassium channel inhibitors

but not affected by large conductance calcium-activated potassium channels inhibitors

[175]. However other mechanisms might be involved in inducing hyperpolarization. For

example Busse et al, demonstrated that the increase in intracellular free Ca+2 might be the

preliminary event leading to the secondary membrane hyperpolarization [176].

Several characteristics have been identified for the hyperpolarization mediator including high solubility and transferability, and ability to produce hyperpolarization- induced relaxation [107, 177]. Additionally the hyperpolarization of the smooth muscle usually is more transient than the accompanying relaxation induced by endothelium- dependent vasodilators [155, 170, 171, 178]. The exact nature of the endothelium- dependent hyperpolarization factor however has not been identified as several substances

159 have been proposed to cause hyperpolarization in smooth muscle cell. Prostacyclin, nitric

oxide, and products of cytochrome P450 have been reported to cause smooth muscle

hyperpolarization in different species at various locations in the vascular system. Fore

example prostacyclin and nitric oxide have been reported to increase potassium

trafficking across the smooth muscle membrane by activating ATP-sensitive [179] and

calcium-activated channels [180]. Moreover, experiments performed mainly in bovine

and porcine coronary arteries show that smooth muscle hyperpolarization is inhibited by

inhibitors of cytochrome P450 monooxygenases and associated with the release from

endothelial cells of epoxyeicosatrienoic acid (EET) (substance produces

hyperpolarization of vascular smooth muscle) [181, 182]. Alternatively,

hyperpolarization of endothelial cell causes accumulation of potassium ions in the

intercellular space between endothelial and smooth cells. This increase potassium

concentration can provoke the hyperpolarization of vascular smooth muscle cell by

activating the inwardly rectifying potassium conductance [183] and Na/K pump [184].

Considering all these dynamics, potassium could also be regarded as an endothelium-

dependent hyperpolarization factor. However these facts, about the nature of

endothelium-independent hyperpolarization factor, are not true in all blood vessels of

different species. For example in blood vessels of the pig as well as in various arteries

from human, chemically unrelated inhibitors of cytochrome P450 do not produce an

inhibition of hyperpolarization-responses [185-187]. Additionally these proposed factors might all be responsible for hyperpolarization-responses in the smooth muscle cell of vessels as they all contribute at different stages and to various degrees to the

160 hyperpolarization process. Therefore, more research is needed to further identify the

nature of this factor(s) in different metabolic tissues of different species.

Resent investigations have shown the importance of endothelium-dependent

hyperpolarization factor increases as the vessel size decreases in rat mesenteric as well as

in human gastroepiploic arteries [186, 188]. Additionally endothelium-independent

hyperpolarization factor is impaired in aging and hypercholesterolemic human and

animals [186, 189, 190] and in hypertensive and diabetic animals [189, 191].

Endothelium-dependent contraction It was as early as 1982 when the vasoconstrictory role of the endothelium was

demonstrated as arachidonic acid augmented smooth muscle contraction evoked by

norepinephrine in the canine systemic and pulmonary veins [192]. This augmentation disappeared when the endothelium tissue was mechanically removed suggesting that the endothelial cell can be a source of vasoconstrictor signals [192]. Endothelium-dependent

smooth muscle contraction has been described in a verity of blood vessels under various

conditions and plays a role in maintaining proper vascular tone without implying endothelium dysfunction [56, 193].

Endothelin The existence of a contractile factor released by the endothelium cell was demonstrated as early as 1985 with the presence of extracellular calcium and independent of the adrenergic, cholinergic, serotoninergic, or hitaminergic systems [195, 196]. This factor was then characterized and named endothelin, as it was isolated from the endothelial cells of pig aorta in 1988 by Yanagisawa et al. Endothelial was described as a peptide made up of 21 amino acids with a N-terminal domain provided a very strong affinity to the receptor [59]. Subsequently three isoforms of the peptide were identified and named

161 endothelin-1, endothelin-2 and endothelin-3. All subtypes are present in all mammalian

species, however endothelin-1 is the only one made in the endothelial cell [197].

Endothelin-1 is ten times more potent vasoconstrictor then angiotensin II and induces a

very long lasting hypertensive action [198]. However endothelin-2 is the most potent

among all of them and has longer duration of action on blood pressure whereas

endothelin-3 has 1/20th of the vasoconstrictor activity of endothelin-1 [59].

Figure LR4.9. Agents cause the release of endothelium-derived contracting factors. EDCF1; endothelium-derived contracting factor, PO2; partial pressure of oxygen, AI&II; angiotensin I & II, AA; arachidonic acid, SP; substance P, 5-HT; 5-hydroxytryptamine, PGH2; prostaglandin H2, TXA2; thromboxane A2, AVP; arginine vasopressin, TGFα; transforming growth factor beta, ACE; angiotensin converting enzyme. Taken from [194].

The secretion of endothelin-1 is regulated at the level of peptide synthesis due to the lack

of dense secretary granules where it can be stored then released as needed. The

expression and release of endothelin-1 is induced by several factors including adrenaline,

thrombin, vasopressin, angiotensin II, insulin, cytokines, growth factors and physical

stimuli (figure 4-9). The production of endothelin-1 involves the increase in intracellular calcium and protein kinase C activity, which is induced by many of these agents [199].

162 On the other hand nitric oxide [200], prostacyclin [201], natriuretic peptides [202],

heparin [203] and protein kinase C inhibitors reduce endothelin-1 expression and release

[199].

Two types of endothelin receptors were identified and termed ETA and ETB

receptors. The affinity of the three isopeptide to ETA is endothelin-1 > endothelin-2 >>

endothelin-3, where the affinity of endothelin-1 in man is 1000-fold more than

endothelin-3 [204]. However the ETB receptors show no selectivity for any of the

endothelin isoforms [205]. The vasoconstrictor response is produced by stimulation of

ETA receptors present in vascular smooth muscle and relaxing response by stimulation of

ETB receptors in the endothelium. However studies reported the existence of ETB in

vascular smooth muscle and produces vasoconstriction [206, 207].

Administration of endothelin-1 induces an initial reduction in arterial blood

pressure followed by prolonged increase. This reduction is transient and most noticed

during elevations in resting blood pressures however have been attributed to the release

of vasodilator substances including prostacyclin and nitric oxide from the vascular

endothelium [126, 208, 209]. Similar to other vasoconstrictors, endothelin induces

smooth muscle contraction by increasing intracellular C+2. The initial phase of this

+2 contraction is achieved by activating IP3 system thereby mobilizing intracellular C

stores. The prolonged phase of the contraction is maintained by stimulating the entrance

of extracellular C+2. However the mechanism by which the entrance of extracellular C+2

is achieved is not clear yet. Several investigations have suggested that the formation of

+2 inositol tetrakis phosphate (IP4) might trigger the entrance of C [210-212].

163 In addition to its vasodilator and vasoconstrictor effects endothelin-1 promotes

proliferation of vascular endothelium and smooth muscle cells and fibroblasts, which

results in structural changes to the vascular wall. The mitogenic effect of endothelin-1 is

attributed to stimulation of tyrosine residues phosphorylation and activation of two other

mitogen-activated protein kinases in the vascular smooth muscle [213, 214].

Numerous reports have been published regarding the possible role of endothelin

in a variety of phathophysiological conditions involving several body systems. These

systems include the renal system, pulmonary system, heart and vasculature. This

involvement in a wide range of diseases and conditions is due mainly to the endothelin capacity to induce constriction and cell proliferation. In the vascular system, endothelin

increases vascular tone, vascular resistance thereby blood pressure. It is also believed to

be involved in the atherosclerotic process and vascular structure alterations. Endothelin is

also involved in increasing the production of other vasoconstrictor including rennin,

angiotensin II, epinephrine, and aldosterone as well as increasing sympathetic activity.

Besides its vasoconstrictor and pressor effect, endothelin has positive inotropic properties

that increase the contractility of vascular smooth muscle [215, 216].

Veins appear to be super sensitive to endothelin as it was 5- to 10-fold more

potent constrictor in isolated large veins than isolated large arteries from dog coronary,

mesenteric, femoral, renal, and internal mammary vessels [217, 218]. Similar

observations were also noted in human as endothelin produced greater contraction in

human saphenuos vein then the gastroepiploic artery [219]. Endothelin effect in human

is described as slow onset venoconstriction. This constriction was maintained even

following the administration of Ca+2-blockers [220, 221]. On the other hand

164 venoconstriction was reversed when nitric oxide was administrated but to a lesser extent

than as observed in the arteries [219]. Additionally infusion of potassium channels opener

drugs and calcium channels antagonists prior to endothelin administration prevented

endothelin-induced constriction [222]. In a comparison of the constrictor effect in human superficial hand veins, endothelin-1 was 26 times and enodhtlin-3 was 3 times more potent than sympathomimetic drug (phenylephrine) [223, 224].

Activation of voltage-dependent channels [59, 225], closure of ATP-sensitive potassium channels [222], and most importantly mobilization of Ca+2 from internal

sources [226] are the mechanisms by which endothelin increase Ca+2 levels and

consequent venoconstriction.

4.4. METABOLIC VASOREGULATION

The metabolic theory of vascular control is characterized as changes in the concentration

of dilator substances(s) alter the excitation-contraction coupling of vascular smooth

muscle cells. Increased vascular resistance and decreased blood flow to metabolic tissues

causes increased vasoactive metabolic substance concentrations in the interstitial space

[97, 227]. This theory indicate that imbalanced blood demand/supply induce the

production and accumulation of vasoregulation mediators. These mediators interact with

the vascular smooth muscle of the resistance vessels inducing alterations in vascular tone

thereby changes in local tissue blood flow [228].

Hypoxia has been implicated as a vasomediator since oxygen might be a limiting

factor in aerobic ATP production. This hypothesis indicate that reduction in cellular

[ATP] during muscle exertion induces vasodilation [229]. However studies suggested

that respiratory chain function and ATP production wouldn’t be affected until

165 mitochondrial PO2 reaches ~0.05 mmHg [230, 231]. This drop in PO2 can’t be found in

normal situation even under hypoxic conditions experienced during exercise [232, 233].

Hypoxia-induced smooth muscle contractions in isolated peripheral, coronary and

cerebral arteries of the dog however have been documented. This response was attributed

to an unknown substance produced from the endothelium to the smooth muscle [234-

236]. Endothelium-dependent vasoconstriction to hypoxia appears to be dependent on activation of voltage-dependent calcium channels in smooth muscle because the contraction was blocked by calcium antagonists [237].

Upon the discovery of endothelin, as a vasoconstrictor, it has become an attractive candidate especially it produces a contractile effect sensitive to inhibition with calcium antagonists and does not require the activity of cyclooxygenase in animal models.

However further analysis of the characteristics of the substance(s) released by hypoxia indicated that endothelin is not responsible for the response [238]. In addition contractile respond to endothelin is slow and sustained whereas hypoxia-induced contraction was fast and transient [239]. On the contrary, hypoxia-induced contraction can be blocked when activity of nitric oxide is reduced by either methylene blue or inhibition of nitric oxide synthesis. Thus endothelium-dependent contractions to hypoxia might be due to the reduced activities of nitric oxide caused by hypoxia rather than the release of a diffusible contracting factor(s) [240, 241]. In human internal mammary artery hypoxia-induced contractions can be attenuated by indomethacin or inhibition of nitric oxide synthesis and are abolished by a combination of both. Therefore in human endothelium, two contracting components might be involved during hypoxia; a product of cyclooxygenase and the reduced activity of nitric oxide [242]. Interestingly in re-perfused canine coronary

166 arteries, the hypoxia-dependent contraction is augmented indicating the response may

contribute to myocardial ischemia in coronary artery disease [193].

On the other hand endothelium-dependent hypoxic contractions are not observed in peripheral veins [192]. This lack of response however might be due less sensitivity of the venous smooth muscle to the contracting effect as coronary arterial smooth muscle

“sandwiched” with venous endothelium exhibits hypoxic contraction [235].

Since carbon dioxide production is coupled with oxygen consumption and ATP production [243-245] it might be an attractive candidate as a vasoregulator. Assuming venous PCO2 is representative of tissue PCO2 investigators have found a direct

relationship between venous PCO2 and vasoactivity in both direction constriction and

dilation [45, 246, 247]. However since these investigations where not designed

+ particularly to distinguish between the effect of CO2 and H , changes in pH might also

play a role in vasodilation.

Potassium is constantly floating around during metabolism and muscle contraction as it plays an integral role in cell membrane depolarization and repolarization.

The progressive increase in interstitial [K+] noticed following increased metabolism is

associated with a vasodilator effect [248]. However this vasodilator effect continues as

the rise in [K+] disappears indicating transitory, complimentary or negligible role of potassium in the vasodilation attributed to metabolic out-products [249, 250]. A moderate increase in [K+] has minimum effect on smooth muscles of venous wall in animals [251,

252]. However, higher concentrations (above 12-20 meq/liter) seem to potentiate the

venoconstrictor effects of catecholamines and directly caused vasoconstriction of isolated veins [253-255]. These venoconstrictor effects of K+ ions have been attributed to

167 membrane depolarization and consequent influx of extracellular Ca+2 [256].

Concentrations above 30 meq/liter induce indirect venoconstrictor effect of K+ by stimulating norepinephrine from the nerve endings [257, 258]. Additionally concentrations of 10 or 15 meq/liter seem to inhibit the neuronal uptake of catecholamines. This inhibition was related to the Na+/K+ ATPase of the cell membrane

[257]. This venoconstrictor effect of potassium is appears not significant as compared to the effect to sympathetic or norepinephrine activities.

The relaxing role of adenosine appears to be significant during the early phase of vasodilation [259, 260]. This adenosine-induced vascular relaxation is activated by specific receptors located on the vascular smooth muscle and inhibited by methylxanthines [261, 262]. However the inhibition of adenosine activities didn’t prevent vasodilation to other factors including myogenic [228, 263] and acidosis [264].

Adenosine appears to promote dilation and to inhibit the effect of venoconstrictors, such as norepinephrine, in the venous side of the circulation in dogs [265, 266]. Similar observations were also found in human veins as adenosine promotes venodilation in a concentration-dependent manner [267]. In vivo, human hand veins precontracted with phenylephrine (a sympathomimetic) are relaxed by infusion of adenosine, showing no effect of aging on the venous respond to adenosine [268]. Venous relaxation in respond to adenosine or adenosine products was observed in several other tissue beds including erectile [269], portal [270] and jugular [271] veins of different species.

It appears that several metabolic out-product are involved concurrently in the control of the vascular function. However, it’s not known which of these substances is the sole mediator of metabolic regulation. Evidence from experiments that perturb a steady

168 state suggests that vascular function don’t depend on a single metabolite. Rather the

regulation of resistance vessels may reflect the combined influences of several

+ + vasodilators and the most likely candidates seem to be CO2/H , K , and adenosine [228].

Identifying the attribution of each mediator individually requires a series of

straightforward experiments designed to study the effect of each metabolic out-product in

isolation from the rest of out-products. In addition this relationship has to be established

under all physiological as well as pathological conditions [272]. However such a

designed is not easy to accomplish and requires complex preparations.

4.5. THE MYOGENIC VASOREGULATION

Blood vessels respond to elevation in transmural pressure with constriction and to

pressure reduction with dilation. These changes in the vessel diameter are induced independent from neural, hormonal, or metabolic influence and attributed to the

autoregulations of the myogenic response. This behavior is most observed in the

arterioles however it has been demonstrated in arteries, venules, veins, and lymphatics

[97, 273, 274].

Jones et al. reported the myogenic response as early as 1852 [275], and followed

by similar observations by Ostroumoff et al. in 1868 [276], and Gaskell in 1881 [277].

However in 1902 William Bayliss was the one who clearly attributed a considerable

contribution of vascular control to changes in transmural pressure. Bayliss concluded that

changes in transmural pressure induce changes in the diameter of the vessels too rapid to

be mediated by metabolic accumulation and is similar to isolated arteries constriction

observed following sudden distension [278]. These findings were a matter of debate for a

long time until 1949 Folkow demonstrated that denervated vessel preparations developed

169 pressure-dependant vascular tone [279] and autoregulation of blood flow was attributed

in part to pressure-dependent mechanism [280]. These findings were followed by

intensive investigations resulted in developing sophisticated techniques to study whole

organ circulation [34, 281] and quantify the myogenic responses in the

[282, 283]. These investigations confirmed previous reports and concluded that the

vascular responses to changes in transmural pressure are rapid [284] and induce

considerable changes in vascular resistance [34, 281]. Development of isolated vessel

techniques [285] has also allowed for more accurate quantification of myogenic mechanism in small arteries [286, 287] and arterioles [288, 289] where the effects of pressure could be clearly distinguished form flow, metabolic, neural and endothelial influences.

The magnitude of the myogenic respond is inversely related to vessel size among arterioles of a given vascular network [14, 290]. However the magnitude of the myogenic respond of similar size vessels varies in different vascular beds. For example the rat small mesenteric vessels have weak myogenic respond whereas it is strong in similar sized

cerebral [291] and skeletal muscle [292] vessels. These differences in the response were also observed between vessels with similar diameter in the same vascular bed [289].

Therefore the existence of the myogenic response is well documented however its strength varies between vessels even in vessels of similar diameter.

Several studies, performed in isolated and in vivo vessel preparations, have attributed part of the basal vascular tone to the myogenic respond [14, 293]. In these studies increase in vascular transmural pressure resulted in vasoconstriction and increased vascular resistance independent of other controlling factors [294]. The myogenic

170 response is also postulated to play an important role in maintaining constant blood flow

and capillary hydrostatic pressure during variations in systemic arterial pressure [97].

Whole organ data suggested that changes in arterial inflow and venous outflow pressures

produced changes in arterial resistance that would serve to minimize changes in capillary

hydrostatic pressure. Mellander et al. have reported constant cat hand-limp volume

during wide range changes in systemic pressure [294]. Given the tissue volume is a

reflection of capillary hydrostatic pressure, it was assumed that constant capillary

hydrostatic pressure was achieved by the autoregulation of the myogenic respond [295].

However these assumptions have disregarded the role of Frank-Starling law and other

local regulatory mechanisms. The myogenic response might be involved in maintaining

blood flow and capillary hydrostatic pressure, however other regulatory mechanisms

might play a more important role [14].

Vascular smooth muscle contraction and relaxation associated with the myogenic

response is thought to be the results of vascular smooth muscle membrane depolarization

and increased calcium ions permeability [296, 297]. However recently researchers have

found that myogenic contraction is initiated by depolarization of the vascular smooth muscle, which modulate calcium ions entry to the cell through voltage-gated calcium channels [298]. Data from cat cerebral artery showed graded (by –20m) vascular smooth muscle depolarization occurred as pressure was changed from 0 to 160 mmHg [299].

Similar observations were recorded in pig coronary [300], bladder myocytes [301] and mesenteric artery myocytes following longitudinal stretch [302]. This depolarization was

associated with activation of voltage-gated calcium channels and subsequent myogenic

behavior [14].

171 SECTION 5. ACUTE VASCULAR RESPONSES TO EXERCISE

Greater blood flow to skeletal muscles during exercise is a combination of increased cardiac output, increased inactive tissue vascular resistance and decreased skeletal muscle vascular resistance. Certainly this increase is in accordance with active muscle metabolic demands [303] and the result of the integration of central and peripheral mechanisms

[15]. Cardiac output increases as a result of greater venous return and stroke volume, additionally increased sympathetic activities and parasympathetic withdrawal strengthen contractility and accelerates heart rate, thereby larger cardiac output [304].

Vascular adjustments to exercise however are controlled centrally and peripherally. The large sympathetic discharge and increased hormonal activities cause total body vasoconstriction thereby increased vascular resistance. This “global” vasoconstriction is in favor of active muscle blood flow as local vasodilatory mechanisms are activated causing vascular smooth muscle relaxation during exercise [305]. This dynamic behavior of the vasculature, systemic vasoconstriction and local vasodilation, causes blood redistribution. Since blood is more needed in the contracting muscles the majority of cardiac output is redistributed or shunted toward the active muscles while blood flow to most of the inactive tissues is reduced [306].

Vascular adjustments are often regulated by different mechanisms during different phases of exercise as none of these mechanisms could alone explain the integrated control of blood flow to the active muscles [307]. While vascular control during exercise is complex this chapter will try to explain the mechanisms responsible for increasing blood flow to the active muscles. The systemic control of vascular function

172 will be briefly discussed whereas the peripheral vasoregulatory mechanisms will receive greater attention.

5.1. SYSTEMIC VASOREGULATION DURING EXERCISE

Muscle contractions stimulate the vasomotor center to increase sympathetic activities and to cause parasympathetic attenuation [308]. The mechanism (s) by which the vasomotor center is stimulated during physical activities is still a mater of debate.

Since heart rate and ventilation at onset of exercise increase almost immediately whereas electrically induced exercise (central command inoperative) developed a slower tachycardic response some researchers have suggested a centrally mediated sympathetic control of the cardiovascular function. Freyschuss et al. reported that static contraction increased heart rate and blood pressure with or without neuromuscular blockade, however this increase was one-half greater without blockade [309]. Therefore the supporters of this theory propose that motor outflow from the cerebral cortex interacts with the central neuron pools that regulate the cardiovascular responses to exercise. On the other hand, others reported electrically induced contraction caused an increase in heart rate and blood pressure indicating an activation of the sympathetic nervous system without central command involvement [310-312]. Additionally this increase in sympathetic nervous activities was intensity related [313, 314]. The supporters of this theory indicate that peripheral mechanical and chemical receptors are responsible for increased sympathetic activities associated with exercise. They further explain that reflexes from the contracting muscles seam to be the primary mechanism that activates the vasomotor center to initiate a sympathetic drive [315]. It seems muscle mechanoreceptors are the main source of peripheral reflexes during initial phase of exercise (0-15 sec.) whereas vascular

173 baroreceptors, carotid bodies, thermoreceptors and muscle chemoreceptors are activated

next (15 sec.-2 or 3 min.) by exercise. Peripheral reflexes during steady-state (3 min.

onwards) seems to be derived from the activation of all peripheral mechanisms [316].

However both sides of the argument have ignored the integration of central and

peripheral mechanisms to provoke sympathetic activities. In a recent study, Wichester et

al., [317] demonstrated the importance of both mechanisms to evoke sympathetic

discharge during exercise as demonstrated in patients with Brown-Sequard syndrome.

The spinal cord in these patients is hemisectioned leaving one side of the body with

sensory loss and normal motor function and the other side with normal sensation and

reduced motor function. Voluntary and electrically stimulated static knee extension of

either legs (with either deficit) elicited increases in heart rate and blood pressure, except

during electrical stimulation of the leg with sensory deficit heart rate and blood pressure

didn’t increase. In summary, neural mechanisms of cardiovascular control is redundant

[318] additionally modifications in autonomic nervous system activity could always find a strategy for stimulating the cardiovascular system stressing the plasticity of the neural

network [317].

Regardless of the mechanism (s) responsible for activating the vasomotor center, exercise appears to initiate sympathetic release that is rapidly propagated to the heart and most vascular networks. Probably the most remarkable characteristic of the autonomic nervous system is to elicit cardiovascular adjustments within few seconds. For example the sympathetic nervous system double heart rate within 3-5 seconds whereas blood pressure is increased to double normal within 15 seconds [15, 63]. This exceptional capability is due to the sophisticated and extensive neural network that connects the

174 central nervous system to the heart and almost all segments of the vasculature [61, 64].

Increased sympathetic activities during exercise to small arteries and arterioles increase

vascular resistance, blood pressure and blood flow to areas of need. Whereas the

sympathetic innervation to the large vessels, particularly the veins, assists in propelling

blood from these vessels to the heart thus generating larger venous return, stroke volume

and cardiac output [20]. For example the sympathetic stimulation increases arteriolar

resistance by ~eight-folds and can expel approximately 40% of the blood in the

capacitance vessels [319].

Although sympathetic outflow help increasing blood flow to the contracting muscles by triggering systematic vasoconstriction, substantial increase in sympathetic drive as a result of engaging in near maximal exercises involving large number of muscle groups might induce skeletal muscle vasoconstriction [320, 321]. In a study by Clausen et al., exercising upper and lower body at near maximal intensity resulted in decrease in lower body blood flow. This could be an active reflex redistribution of cardiac output and would indicate that all sympathetic efferent activity is not locally blocked or attenuated

[322]. This increase vascular resistance might be essential to maintain sufficient blood pressure with the circulation [314, 323]. However vasoactive agents resulting from the contracting muscles usually override skeletal muscle vasoconstriction by reducing sensitivity to [324] and maybe even inhibit the release of norepinephrine [102, 325, 326] thereby maintaining sufficient blood flow. Actually norepinephrine might stimulate the release of some of these agents thereby induce vascular smooth muscle relaxation and increase active muscle blood flow [327].

175 Cardiovascular responses to dynamic upper body exercise are different from lower body exercise. Due to the lack of lower body musculature involvement during upper body exercise a large vasculature remains undilated providing resistance to peripheral circulation [328]. This larger peripheral resistance induces substantially higher blood pressure at a given exercise intensity. Additionally at a given submaximal steady state oxygen consumption, heart rate during upper body exercise is higher. Due to less musculature involvement in upper body exercise venous return and stroke volume are lower. Therefore, since cardiac output is similar for upper and lower body exercises higher heart rate compensate for the reduced stroke volume [329, 330].

The complete parasympathetic withdrawal is achieved at heart rate of ~100 b/min.

This withdrawal causes an increase in cardiac output however it has minimum effect on the vasculature since they have very few vagal receptors [72]. Therefore as far as neural control, the sympathetic nervous system plays a greater role in controlling the vascular during exercise [65].

Increased sympathetic drive, during exercise, activates the medullary portion of the adrenal glands to release relatively large quantities of epinephrine and a smaller amount of norepinephrine. Essentially both adrenal hormones activate adrenergic receptors of the cardiovascular system causing increased cardiac output and generalized vasoconstriction. Hormonal control however seems to play a relatively minor role in regional blood flow during exercise compared with the more direct, rapid and potent sympathetic neural control. Furthermore it is more essential during stead-state exercise as it support the neural activities to maintain appropriate cardiovascular respond.

Additionally other physiological requirements, including thermoregulation, fluid

176 homeostasis and metabolic activities, compete for the adrenal medulla hormones [316,

331].

5.2. PERIPHERAL VASOREGULATION DURING EXERCISE

Increased blood flow and oxygen delivery to the contracting muscles are essential in order for exercise to continue. The first few seconds of exercise are associated with substantial increase in muscle blood flow, perfusion and oxygen extraction while these elements remain elevated as long as muscles are contracting [305]. In fact peak muscle blood flow during exercise is estimated to increase ~20 folds and to peak at 300-400 ml/min/100 g [306]. Certainly this increase is in accordance to muscle metabolic demands and related to muscle oxygen consumption [332], muscle contracting frequency

[333, 334], and running speed [335, 336]. Additionally within the active muscle group the recruited muscle fibers appear to receive most of muscle blood flow [334, 336].

Sympathetic vasoconstriction is induced during exercise throughout the body except in the contracting skeletal muscles. In a study by Johnson and Rowell blood flow to forearm decreased whereas it increased to the leg muscles during prolonged lower body exercise [337]. In fact blood flow was observed to increase to red portions and decrease to white portions of the same muscle during low-intensity exercise [338, 339].

The increased sympathetic vasoconstriction and decreased blood flow to inactive tissues while blood flow to the contracting remains elevated is achieved by functional sympatholysis. This process involves deactivating sympathetically induced vasoconstriction during exercise [325].

5.2.1. Local Mechanisms Contribute to Vasoregulation Various mechanisms have been proposed to contribute to locally induced increase in

muscle blood flow, as none of these mechanisms can be solely held accountable for this

177 substantial increase. Muscle pump, metabolic factors, myogenic response, and

endothelial-derived relaxing factors have been implicated as vasoregulatory mechanisms

responsible for the increased blood flow during exercise. These mechanisms are a

reflection of the contracting muscle activities and primarily to offset the vascular

responses to systematically induced vasoconstriction [254, 340-342].

Metabolic Mechanisms The oldest mechanism have been considered to contribute to hyperemic response

associated with exercise is metabolic vasodilation. As previously discussed, contracting

muscle metabolic byproducts are one of the mechanisms involved in peripheral reflexes

that initiate sympathetic vasoconstriction. Ironically however these metabolites also

appear to be involved in local muscle vasodilation during exercise. The byproducts of

metabolism accumulate in the interstitial fluid surrounding resistance vessels, thereby

establishing a local mechanism that is capable of matching oxygen delivery to the level of

contractile activity. The mechanism(s) by which the sympathetic vasoconstriction in the

active muscle is deactivated varies however the metabolic byproducts generally compete

with the sympathetic vascular regulation causing vascular smooth muscle to be less

sensitive to catecholamines [325]. For example Dodd and Johnson reported that the

metabolic byproducts reduced vasoconstriction mediated by post-junction α1-

adrenorecepoters but not by circulating catecholamines [343]. There is also a differential

sensitivity of α1- and α2-adrenoceptor constriction to metabolic inhibition during skeletal

muscle contraction as metabolic release of dilator substances attenuated α2-adrenoceptor

constriction 10-folds more than α1-adrenoceptor constriction [327].

Anyway the apparent relationship between muscle metabolism and muscle blood flow during exercise indicates a factor related to muscle metabolism might contribute to

178 vascular tone regulation. This relationship becomes evident when vasodilation was observed in resting recipient muscle upon the infusion of venous blood from a contracting muscle [344, 345]. It appears metabolites produced by the contracting muscles diffuse to the resistance vessels causing vasodilation and increased capillary recruitments. Adenosine, magnesium, potassium, calcium, oxygen and overall osmolarity are among the vasoactive factors associated with skeletal muscle metabolism and have been implicated in elevation of muscle blood flow during exercise [325]. Isolating the vasoactivity of each factor is a challenge however some studies have successfully observed the action of some of these factors.

Adenosine appears to accumulate in the vicinity of the contracting muscles [346].

The source of adenosine could be muscle tissue [347], endothelium tissue [348, 349], or a property of nervous activities [350]. However the muscle seem to be the major source of adenosine during contraction as electrically stimulated muscle produced adenosine without the present of nerve and endothelial cells [347].

Regardless of the source during contraction adenosine seems to induce vasodilation associated with increased muscle blood flow [15]. Arterial infusion of adenosine has been shown to produce vasodilation in most vascular beds including in human forearm and leg muscles [351] [352]. The magnitude of this vasodilation is comparable to the one observed during peak effort and greater than induced by acetylcholine [353].

Adenosine blockade seems to show the importance of adenosine vasodilatory role as exercise hyperemia is reduced in animals. During electrically induced muscle contraction exercise hyperemic response was reduced in cat [354], dog [355] and rabbit

179 [356] when adenosine receptors were blocked. Additionally peak and average arteriolar diameter and volume flow in hamster was reduced [357]. Data from human exercise investigations are however insufficient to form a conclusion. Systemic infusion of nonspecific adenosine receptor blockers in human reduced total reactive hyperemia after

5-10 min. of forearm occlusion [352]. However minimum information available regarding adenosine receptor blockade on exercise-induced hyperemia in human skeletal muscle. Additionally since the only drug used in human trials to block adenosine receptors is nonselective therefore other vasodilator mechanisms might be inhibited or activated and mask the blockade effect [358].

Muscle fiber type and intensity of exercise might play a role in the vasodilatory effect of adenosine. In a study by Schwartz and McKenzie the administration of adenosine deaminase (promotes the degradation of vasoactive adenosine) reduced muscle blood flow by ~40% during high intensity contractions of highly oxidative slow-twitch cat soleus muscle followed by reduction in muscle oxygen consumption and force production. Whereas the same variables in the same muscle were not effected by adenosine deaminase during low-intensity contraction. Moreover adenosine deaminase had no effect on these variables during neither low- or high-intensity contraction of fast- twitch cat gracilis muscle [359]. These data indicate that the vasodilatory effect of adenosine appears to be most evident during high-intensity contraction of highly oxidative slow-twitch muscles. This might be related to adenosine formation from AMP, which is proportional to oxidative capacity of the tissue, which greatest in slow oxidative muscle fibers during high intensity contraction. [359].

180 Since potassium is one of the byproducts of skeletal muscle it has been implicated as one of the contributors to the vasodilation associated with exercise [341, 345]. Several studies have illustrated the relation of accumulated potassium to reduced vascular resistance [333, 360] and increased muscle blood flow during exercise [345]. The vasodilatory effect of potassium however seems to be transient. The increase in venous potassium [361] as well as potassium infusion [345] were found to be effective for a short period of time indicating involvement in the initial phase but not in the maintenance of functional vasodilation. Additionally resting blood flow or the relationship between oxygen consumption and tissue blood during contraction were not changed as a result of potassium infusion [362]. Therefore potassium has been rolled out as a primary vasodilator during steady-state exercise [362, 363]. However the increase in systemic potassium observed during large mass muscle maximal contraction might be sufficient to induce vasodilation in other vascular beds [363].

While changes in hydrogen and subsequent reduction in pH during exercise are well established [345] no relationship was found between venous blood concentration of hydrogen ions and vascular resistance [360]. Additionally vasodilation observed during exercise was not associated with increased arteriolar hydorogen ions nor superfusion with highly acidic mixture resulted in significant vasodilation [364]. Therefore alterations in local pH don’t seem to play a role in increasing blood flow to contracting muscle.

Since ions and metabolic byproducts increase in the venous blood the subsequent increased in osmolarity [341, 345, 365] might be presumed as a vasodilator agent during muscle contraction. Although raise in osmolarity can result in relative increase in skeletal muscle vasodilation [366], elevated venous blood osmolarity during contraction is not

181 related to the magnitude of the decrease in vascular resistance [333, 360] and observed

dilation is transient [345, 365]. Additionally artificially increasing osmolarity in the

venous blood of resting muscle does not result in vasodilation similar to the one observed

in exercising muscles [365] nor alters the relationship between blood flow and tissue

oxygen consumption [362]. Moreover blood flow remained elevated while osmolality

returned to control values during long-lasting muscle contraction [367], whereas blood

flow returned to control values as osmolality in venous blood remained high during close

arterial infusion of hyperosmotic solution [368]. Therefore local osmolarity might play

minor role in the vasodilation observed during exercise other factors associated with

exercise seem to have greater influence.

As important as each metabolite in vasodilation associated with muscle

contraction simultaneous alterations of multiple chemical factors have a synergistic effect

on vascular function. In a study by Skinner and Costin, simultaneous manipulations of

oxygen content, potassium concentration, and osmolarity of arterial blood could produce

vasodilation comparable to that seen during contraction while manipulation of single

chemical factor resulted in weak vasodilation [341]. Similarly correcting venous blood

PO2 and pH completely abolished the vasodilator effect of venous blood from the

contracting muscle [361]. Additionally Mohrman and Regal found the combined effect

insufficient PO2 and PCO2 is profound and may account for most the metabolic

vasodilation observed during contraction [332]. These and other studies confirm the importance of each muscle metabolic byproduct during contraction however they also

emphasize the importance of the combined effect of these byproducts. They also suggest

that a variety of chemical combinations can result in profound increases in skeletal

182 muscle blood flow, which might differ according to muscle fiber, contraction speed,

stimulation parameters, species or other factors associated with muscle properties during

exercise.

As much as the metabolic theory is accepted, it doesn’t not explain all the hyperemia observed during exercise. Additionally the vasodilation generated by metabolic factors (e.g. potassium, PO2, adenosine, osmolarity) seem to be sufficient to

initiate exercise hyperemia however don’t persist as exercise hyperemia continues. This

indicate that they are either are not responsible for exercise vasodilation or are

superseded as exercise continue [250, 363]. Additionally vasodilation occurs at location

not within the metabolic vasodilatory factor reach. Therefore other mechanism(s) should

be involved in vasodilation observed during exercise especially after the initial phase of

exercise hyperemia and in locations away from interstitial fluid [369].

Mechanical Mechanisms Skeletal muscle during contraction undergoes a cycle of shortening and lengthening causing the muscle contracts and relaxes, respectively. The contracted muscle produces mechanical tension that reduces and, in some occasions, pauses blood flow to the active tissue even during maximal vasodilation [344, 370]. Any reduction in blood flow during contraction is rapidly compensated for by an overshoot during the relaxation phase.

Additionally muscle force development during rhythmic contraction fell sharply with small reduction in perfusion pressure and blood flow [371]. In the venous system, the muscle pump propels blood toward the heart as well as maintains venous pressure.

During quiet standing, gravitational effects cause the drainage of thorax blood to the lower parts. Instantly skeletal muscles surrounding the veins start contracting and pushing against the veins to propel blood toward the heart [2, 15]. Meanwhile with the help of the

183 venous valves the veins refill during muscle relaxation. In this case skeletal muscle is

acting as a pump, which is clearly demonstrated during exercise as it plays a more

effective role in driving blood toward the heart during exercise.

Transmural pressure is the pressure gradient across the vessel wall and calculated

as the difference between internal pressure and external pressure. Elevation in transmural

pressure causes vasodilation whereas reduction causes vasoconstriction. These changes in

the vessel diameter are induced independent from neural, hormonal, or metabolic

influence and attributed to the autoregulations of the myogenic response. This behavior is

most observed in the arterioles however it has been demonstrated in arteries, venules,

veins, and lymphatics [97, 273, 274]. Due to increase in byproducts within the muscle

tissue pressure increases during exercise causing a decrease in transmural pressure

thereby vasodilation. The reduction in transmural pressure and subsequent increase in

vasodilation and blood flow can even occur at intensities as low as 15-30% of MVC [372,

373]. In fact Mohrman and Sparks reported that vascular myogenic respond accounts for

~50% dilation observed following 1 sec tetanic contractions in the dog gastrocnemius

muscle [374]. Myogenic vasoconstriction, caused by rapid increased in intravascular

pressure, seems to be very potent and can’t be overridden by local vasodilators at rest and

during contraction. In a study by Meininger et al., the myogenic vasoconstriction is

maintained in the presence of extremely low tissue PO2 in resting skeletal muscle [375].

Moreover Lash and Shoukas reported reduction in functional vasodilation by 60-70% as a

result myogenic vasoconstriction during rest and exercise [376].

Increased vascular resistance and cardiac output associated with exercise result in systematic increase in blood pressure, which is also another mechanical factor contributes

184 to vascular control. Elevated mean arterial blood pressure increases perfusion pressure

across the vascular bed thereby increasing muscle blood flow [377, 378].

Endothelial Mechanisms While metabolites with vasodilator properties have been considered to amply contribute to vasodilation it doesn’t fully explain the magnitude, persistence and the sites of hyperemia observed during exercise. For example the vasodilatory effect of metabolites

seems to diminish after the initial stage of exercise (~ 1st 3 minutes) [250, 325, 363] as

well as hyperemia observed during exercise is more than the maximum possible

hyperemic respond can be produced by metabolic byproducts [305, 369]. Additionally

although it is generally assumed that small arterioles are the major resistance vessels

recent investigations revealed that 40-55% of the total vascular resistance reside in large

arterioles and small arteries. Metabolic byproducts can’t account for the vasodilation

detected in the larger upstream vessels, which are not bathed with interstitial fluid or

subjected to skeletal muscle metabolic factors [369, 379]. These observations indicate the

involvements of and have commenced the search for other mechanisms to explain all

aspects of the tremendous increase in skeletal muscle blood flow during exercise.

The involvement of the endothelium tissue in the regulation of blood flow has

been suggested first by Furchgott and Zawadzki [104] and has provided further

explanations to aspects associated with exercise hyperemia. Pharmacological agents such

norepinephrine, acetylcholine and bradykinin as well as physical factors such as shear

stress have been proposed to cause the release of nitric oxide. Both of these stimuli are

augmented during exercise [380, 381] and increase intracellular level of Ca2+, which

triggers nitric oxide synthase activities and subsequent nitric oxide production.

185 Indirect evidence, from pharmacological studies of large conduits, strongly

suggests that α2-adrenergic receptors are present on vascular endothelial cells [382, 383].

Since catecholamines increase during exercise, it would be justifiably predicted that they

might induce the release of nitric oxide. Additionally nitric oxide released has been

observed to diminish contraction induced by α1- and/or α2-adrenergic receptors located on

the vascular smooth muscle cell [102, 383]. These data imply stimulation of α2-

adrenergic receptors might be involved in the release of endothelial nitric oxide.

Flow-mediated dilation or shear stress-induced vasodilation is another function of

the endothelial cell. During exercise, increased blood flow velocity within the vascular

lumin causes an increase in the blood drag force against the vascular walls creating a

shear stress. The increased stress of blood shearing against the vessels is a signal that

causes vasodilation to reduce stress against the vascular wall. However, intact functional endothelium is obligatory for optimal regulation of shear stress [384, 385]. Stress of blood sheering against the endothelium layer produces vasodilatory factors, including

nitric oxide, from the endothelium tissue resulting in vasodilation of the proximal smooth

muscle cells [305]. Once released nitric oxide diffuse to the underlying vascular smooth

muscle cells, which activates guanylate cyclase to convert GTP to cGMP, thereby

vasodilation occurs [126].

Skeletal muscle blood flow at rest is estimated to be 2-4 ml/100g/minute whereas

it might exceeds 50-80 ml/100g/minute and values as high as 500 ml/100g/minute have been reported during strenuous exercise [306, 386]. This tremendous increase in muscle

blood flow is highly associated with the metabolic demands of skeletal muscle. Although

terminal arterioles and small arterioles regulate blood flow to the capillaries and

186 contribute largely to vascular resistance large arterioles and small arteries control the

amount of blood received by the terminal arterioles and account for 40-55% of vascular resistance [369]. Since they are within the influence of the muscle metabolic signals, the diameter of the smaller vessels change according to the muscle metabolic demands.

However upstream feed arteries are not within reach of the metabolic influence yet their

vasoreactivity are still in accordance to the contracting muscle metabolic demands [305].

Moreover smaller vessels are mainly controlled by changes in local metabolites thus they

regulate intra-organ blood flow whereas neural and endothelium regulation is

concentrated on the larger arteries which are better placed to regulate the organ perfusion

and systemic blood pressure [305]. Smaller vessel dilation is capable of increasing blood

flow by a factor of only 2-3 times more than resting levels, which is simply not adequate

to attain increase in flow of blood observed during exercise. Additionally without a

decrease in feed artery resistance, increased perfusion pressure through the microvessels

would be prohibited, thereby limiting the driving force for flow. Therefore concomitant

and coordinated dilation of the feed arteries located upstream seems detrimental to meet

the metabolic demands of the exercising muscle [305, 369].

Consequently it is clear, in order to match blood flow with muscle metabolic

requirements, the vasodilatory response to exercise has to be controlled by coupling

microvessel dilation with that of upstream arteries. This could be achieved by ascending

signals from the downstream vasculature to that of the upstream to induce feed artery

dilation [305]. These signals could be a functional vasoactive respond that propagates

vasodilation by direct cell-to-cell conduction along the vascular wall vie gap junctions of

either the endothelium tissue [387] or the smooth muscle cell [388]. Recent work

187 suggests that the signal might be a depolarization propagation triggered by acetylcholine along the endothelial cell layer ascending dilation along the arterial network [387, 389], however more research is needed to verify these findings. Therefore functional hyperemia seems to be originated in the smallest pre-capillary arterioles, which are generally presumed to be the most sensitive to vasodilator metabolites and spreads towards the lager arteries.

With flow-mediated dilation mechanism, sheer stress causes the release of endothelium-relaxing factor(s), which relaxes adjacent smooth muscle cells. Through this mechanism, the initial dilation of downstream vessels will increase flow through upstream vessels, inducing dilation of the upstream vessels [379, 390]. In fact cell-to-cell communication during exercise has also been reported to be influenced by venous factors. Venular endothelial cells can also release factors that relax adjacent arterioles and thereby contribute to functional hyperemia, without the present of arterial endothelium

[391].

Nitric oxide is clearly essential for the vasodilatory respond to exercise, recent evident however suggest the contribution of endothelin-1 in the increased blood flow to the contracting muscle. Maeda et al., demonstrated increased circulating endothelin-1 following 30-minute exercises at 90 and 130% of ventilatory threshold of young healthy volunteers. The concentration of endothelin-1 was the greatest after 30 minutes of exercise cessation and was greater following the higher intensity [392]. In a follow-up study, Maede et al. showed that circulating endothelin-1 increased in femoral artery of the noncontracting muscle and decreased in the contracting muscle during strenuous 30- minitues-leg exercise [393]. The authors of these studies concluded that endothelin-1

188 increase during exercise is intensity dependent and contribute to the distribution of

increased cardiac output during exercising.

5.3. SUMMARY

Muscle contraction triggers systemic and local mechanisms involved in vascular control.

Systemic factors, including neural and hormonal, appear to maintain the homeostasis of

all aspects of the cardiovascular system including the heart, vessels and blood

distribution. The raised sympathetic drive increases cardiac output, total vascular

resistance and direct blood toward the working muscle. The shift of the increased cardiac

output toward the active muscles is also facilitated by the vasodilatory factors produced

at the site of contraction by overriding the vasoconstrictory effect induced by the

sympathetic discharge. These factors include metabolic, endothelium and mechanical factors. Metabolic vasodilators tend to influence the arteries within the interstitial space, which includes capillaries, terminal and small arterioles. Whereas endothelium-induced vasodilation seems to be involved in flow-mediated dilation, which occurs mainly in the large arterioles and small arteries and triggered mainly by the dilation in the distal vessels and sheer stress. However these factors jointly increase the diameter of the vessels

thereby blood delivery to meet the metabolic demands of the muscle. The relative

importance of each of these factors to the vasodilatory response is not well understood,

however each factor seems to play a role and to contribute to increased blood flow to the

contracting muscle either throughout the exercise period or limited to only initiation of

the blood increase. Moreover some factors appear to be more important than others

depending on the muscle fibers involved in contraction. It seems, as might make sense,

muscle blood flow increase in oxidative fibers more than nonoxidative. Finally systemic

189 and local factors compete to takeover the control of the vessels however it appears as long as the muscle is contracting, factors produced locally usually claim “victory”.

5.4. VENOUS RESPONSES TO EXERCISE

According to the Frank-Starling theory of contraction, the heart can pump only what passively fills it during diastole. The venous system operates as a collecting reservoir for blood and controls the quantity of blood returning back to the right atrium. The increase in the amount of blood mobilized in the veins and propelled back to the heart during exercise is well established. This increase in venous return is essential to accommodate the required increase in cardiac output [32]. While the controlling mechanisms of blood flow in the arterial vasculature during muscle contraction are complex and interrelated more controlling mechanisms are even involved in the venous system. Low intra-vascular pressure, large volume present, intermittent nature of the flow, gravitational effect, venous valves, cardiac pressure, and the collapsible nature of the venous wall add complicity to flow in the venous cavity as all these factors interact with each other during exercise to achieve blood return to the heart [20].

Flow of blood runs with no resistance in large veins when they are distended however as veins enter the thorax and abdominal cavities flow is impeded. Most veins are compressed at many points by the surrounding tissues as a result of the intra-thorax and abdominal pressures, which obstructs flow of blood to the heart chambers [27]. Due to the increase in intercostals muscle contractions during exercise the physical pressor in the intra-thorax cavity increases causing further obstruction of blood in the veins.

Pressure in the right atrium might also hinder blood returning back to the heart from the systemic circulation. The increase in speed and force of cardiac contraction

190 during exercise causes elevation in cardiac pressure including pressure in the right atrium resulting in venous return blood to backup [32, 35]. However pressure against flow of blood in the venous system doesn’t increase until all collapsed segments between peripheral veins and large veins have opened, which doesn’t occur until right atrial pressure reaches 4-6 mmHg. A further increase in right atrial pressure would reflect a progressive pressure increase in large veins [33, 34].

Upon presuming standing position, gravitational effects cause the drainage of thorax blood to the lower parts. Instantly skeletal muscles surrounding the veins start contracting and pushing against the veins to propel blood toward the heart meanwhile, the veins refill during muscle relaxation [2, 15]. In this case skeletal muscle is acting as a pump, which is clearly demonstrated during exercise as it plays a more effective role in driving blood toward the heart. The efficacy of this pump is intensity dependent and certainly enhances venous blood flow [15].

The contraction of the venous smooth muscle also plays a role in driving blood toward the heart, however its contraction depends on the intra-vascular pressure. The venous wall movement is mainly passive at low pressure, however when venous pressure increases above 5 mmHg wall tension increases providing an active role of the venous tone. Venous tone is controlled mainly by central and local neurohumoral signals [2, 27].

While the skeletal muscle pump and venous tone is important to propel blood forward, the venous valves prevent flow regurgitation. Following skeletal muscle contraction and propelling blood forward, the valve cusps close and the veins refill rapidly allowing unidirectional flow only [13, 20].

191 The intra-thorax pressure varies during quiet breathing from –2 to –4 mm Hg during expiration to about –5 to –7 during inspiration. The intra-abdominal pressure decreases when intra-thorax pressure increases facilitate venous return form the intra- thorax veins thereby increase thoracic vena cavae flow of blood and venous return. The resultant changes in flow inside the veins cause ±25% changes in cardiac output during respiration. This change in cardiac output is even more pronounced during exercise due to the increase in the rate and force of perspiration [27].

In a series of studies by Bevegard and colleagues venous function was examined during and following exercise. Forearm and calf venous capacitance decreased and pressure increased during leg exercise and combined leg and handgrip exercises.

Additionally forearm and hand venous capacitance decreased in accordance to leg exercise intensity throughout the experiment. The authors attributed these changes to increased venous tone induced by sympathetic venoconstriction and with minimum contribution of the muscle pump especially since venous tone remained elevated even following cessation of exercise [394, 395]. In subsequent studies, Bevegard reported increase in forearm blood inflow without concomitant increase in venous capacitance following cessation of exercise indicating that local vasodilators contributed to arterial dilation but didn’t change venous tone confirming previous observations [396]. Similarly

Seaman and colleagues investigated forearm venous function during handgrip exercise at

20, 30 and 40 % of most voluntary contraction. Venous filling decreased and venous pressure increased indicating greater venous wall tension during exercise. The authors concluded that this increase was intensity dependent and is a reflection of venoconstriction essential to increase venous emptying and right atrial and ventricle

192 fillings [397]. In 1968 Robinson and Wilson examined the effect of exercise on venous function of nonworking muscle with and without propranolol (β-blocker) administration in young healthy males. Forearm venous compliance decreased during exercise than it had been at rest. Moreover despite reduction in heart rate, an indictor of cardiac performance, propranolol administration didn’t cause changes in forearm venous compliance however forearm blood flow and vascular resistance decreased and increased respectively. The authors concluded that venous compliance in the nonworking muscle is decreased during exercise however its not affected by reductions in cardiac output whereas the arterial system is affected by changes in cardiac performance [398]. Detry et al. investigated the effect of exercise on the venular blood volume of nonworking muscles with and without nitroglycerine, a potent vasodilator, in normal young men.

Forearm venous volume decreased by 25% during moderate-intensity leg exercise however this volume increased following nitroglycerine administration indicating volume changes might be due to venoreactivity. The authors concluded that the venous system in the nonworking muscle constricts during exercise however nitroglycerine administration causes venodilation. Additionally the effect of exercise on venous distensibility is essential to increase venous return however the effect of nitroglycerine caused venous relaxation and pooling thereby reducing preload which might partially explain the reason nitroglycerine is helpful for individuals with angina patients [399]. In order to determine variations in venous function hemodynamics during the course of the day, changes in valvular function, calf muscle pump and ambulatory venous pressures were examined during performing daily activities in the early morning and late after noon. These functions were examined during standing up from supine position and after tiptoe

193 exercises in males and females. Indexes of valvular venous function in males and females were lower in the afternoon than in the early morning whereas calf muscle pump and ambulatory venous pressure indexes remained the same. These data indicate that venous function is not affected by either daily circadian rhythm or gender, with the exception of the effect of daily circadian rhythm on venous valvular function [400].

In the contrary other investigations have reported venous dilation. The increase in sympathetic activity associated with exercise causes systemic arterial and venous constriction. However infusion of vasodilator substances as well as applying heat resulted in significant dilation in the arterial vessels but did not change venous tone [401-403].

These results indicate the involvement of other factors. Marshall and Tandon investigated venous function during exercise using direct measurements. The internal diameter of individual arterioles and venules of rat spinotrapezius were measured immediately after a period of electrical stimulation. The arteries as well as veins similarly dilated within the region of muscle contraction. Venodilation was faster and larger in magnitude in the smallest vessels as well as in individual venous and was contraction frequency dependant indicating metabolic vasodilation. The authors concluded that venodilation results from factors associated with muscle contraction and may play a role in outward capillary filtration and preventing edema formation following exercise [404]. Other investigators reported endothelium involvement in venodilation. Administration of nitric oxide inhibitor in rats resulted in a dose-dependent increase and decrease in venous tone and capacitance, respectively. The authors concluded that endothelium tissue contributes to venodilation observed in animals [405]. Inhibition of nitric oxide was also found to elevate venous tone and reduce return and cardiac filling, confirming the involvement of

194 endothelium-derived nitric oxide in the modulation of venous function [406]. Koller and

collogues examined the effect of shear stress on the walls of isolated venous diameter and reported increased nitric oxide and prostaglandin release and venodilation. The authors concluded that shear stress, as observed in the arteries, induces endothelium-derived nitric oxide and prostaglandin release, which contribute to the regulation of venous function. Additionally, since shear stress increases during exercise, the authors expected that endothelium is also involved in venous regulation during muscle contraction [407].

There is a general certainty that venous emptying increases during exercise to suffice the demanded increase in cardiac output however the mechanism by which venous emptying increases is still a debate. Contrary to that of the resistance vessels some investigators indicate sympathetically-induced increases in venous tone is not abolished by local vasodilators. Additionally this increase in venous tone propels blood toward the heart [304, 408]. Whereas other studies suggest the involvement of local vasodilators, including metabolic and endothelial, result in increased venous relaxation and venous diameter [404, 406, 407]. The mechanisms, suggested by both sides of the debate panel, seem to contribute to blood return to the heart during exercise. However there is a clear lack of well-designed studies and conflicting results, which limits our

understanding of the role of venous function during exercise. Additionally the control of

the venous system under resting conditions is complex and involves several mechanisms,

let alone during exercise. Therefore there is a tremendous need for more investigations

that examine each mechanism separately to have an integrative understanding of venous

function and overall cardiovascular physiology.

195 SECTION 6. CHRONIC CIRCULATORY ADAPTATIONS TO EXERCISE

TRAINING

Repeatedly challenging the body systems with exercise will result in adaptations of all body systems to assure maximum performance. While increased exercise tolerance is an obvious feature of the adaptations to exercise training a wide range of cellular, mechanical, chemical, structural, and regulatory changes associated with this increase.

These changes occur in a verity of systems and body parts including, cardiovascular, skeletal muscles, nervous and hormonal systems and depend largely on the type of exercise chosen for training. Changes in these systems are eventually integrated to achieve increase in exercise tolerance [409, 410].

Generally, exercise training involves rhythmic contractions of large muscle groups for a long period of time improves oxygen consumption. This improvement, first and foremost, corresponds to and is to meet the changes in the metabolic characteristics of the skeletal muscle involved and, for the most part, achieved by improvements in the cardiovascular system capacity to deliver oxygen and nutrients and remove carbon dioxide and byproducts. Essentially the delivery and removal aspects of the cardiovascular system become more efficient and result in reduced the workload placed on all components of the cardiovascular system at any given intensity, including resting.

This decrease in workload allows the cardiovascular system to function efficiently for longer period of time, which results in increased exercise tolerance [322, 411].

A variety of physiological changes occur in the cardiovascular system following long-term exercise training. These changes involve each aspect of the system including the heart, vessels and blood and all directly influence each other. For example decreased

196 heart rate following exercise training is attributed to increased stroke volume, which

attributed to changes in cardiac structure and increased blood volume. Moreover exercise

training results in increased vessel capabilities to deliver and remove blood to and from

skeletal muscles. This increase is due to the contribution of all aspects of the

cardiovascular system, however studies examining the effect of localized small muscle

group reported increase in blood flow without the involvement of the heart and blood

[322, 412-414]. Since the cardiovascular system is a continuum and its components directly interact with each other changes in the heart and blood will be briefly and

generally described. However this chapter will mainly focus on the vasoregulatory

mechanism adaptations associated with exercise training.

6.1. CENTRAL CIRCULATORY ADAPTATIONS TO EXERCISE TRAINING

6.1.1. Changes in the Heart and Blood Constituents Repeated bouts of exercise place the cardiovascular system under stress and induce

physiological adaptations. Maximum oxygen consumption has been the most useful to

quantify the results of exercise training. Values of oxygen consumption range from 30 ml-1●kg-1●min-1 in young sedentary, to 85 ml-1●kg-1●min-1 in elite athletes, and values of

45-50 ml-1●kg-1●min-1 have been observed in young active individuals [322]. Meanwhile

2 to 3 months of exercise training resulted in increases ranging between 15 and 30%,

depending on the initial level of fitness [413]. Oxygen consumption increases is

attributed to increase in maximal cardiac output and muscle tissue capabilities to extract

oxygen (A-Vd). Exercise training for 2-3 months resulted in 16% increase in oxygen

consumption as a consequence of 8% increase in cardiac output and 8% increase in A-

Vd. The increase in cardiac output has been attributed completely to increases in stroke

volume whereas maximum heart rate doesn’t change, if anything, it might slightly

197 decrease following exercise [322, 412, 413]. Whereas the increase in A-Vd might be due to several reasons including increased muscular vascular density, vasoreactivity and muscle oxidative capacity [415].

Reduced heart rate in athletes and in individuals following exercise training is

well documented. The decrease in resting and sub-maximal heart rate might results from structural, neural and hormonal changes [412, 416] including alterations in sympathetic/parasympathetic balance toward greater dominance of the parasympathetic component [417, 418]. However the reduction in heart rate has been reported to decrease during 7-weeks training program without concomitant reduction in plasma catecholamines indicating the involvement of other factors [419]. The reduction in resting heart rate might also results from changes in stroke volume, which might result from changes in cardiac structure and blood volume. Cross-sectional date indicate that cardiac dimensions and end-diastolic volume are larger in endurance athletes than athletes involved in activities of a short durations, such as resistance training (figure 6-1) [420].

These observations were also reported following training at 85-90% of maximum oxygen consumption for 7-weeks in sedentary individuals [421]. In fact increase in heart dimensions were observed following only 1 week of exercise training [422]. In addition to cardiac structure, Gillen et al. reported 10% increase in blood volume within 24 hours following only one bout of ergometry-cycling at 85% of maximum oxygen consumption

[423]. Additionally one-half of the difference in stroke volume normally observed between untrained and highly endurance-trained men during upright exercise is due to a suboptimal blood volume in the untrained men [424]. Coyle EF et al., concluded that the reduction in cardiovascular function following detraining is due, in large, to the decline in

198 blood volume [425]. Therefore increased cardiac dimensions and blood volume result in increased stroke volumes and subsequent reduction in heart rate.

Figure LR6.1. Cardiac mass index in elite athletes. Taken from [18].

Training-induced increase in blood volume and myocardial hypertrophy contributes also to reduced heart rate during submaximal exercise. However heart rate reductions might not exclusively due to central adaptations. In a study by Saltin et al., heart rate responses to one-leg submaximal exercise in the trained leg were lower than in the untrained leg [426]. Similarly, heart rate responses to leg exercise were higher than to arm exercise following arm exercise training program and vise versa following leg exercise [416, 427]. These studies suggest at least part of the fall in heart rate during submaximal exercise is induced by peripheral factors and specifically related to the trained body part. The relationship between heart rate and the trained muscle group might be mainly due to the decrease in sympathetic stimulation [417, 418]. Exercising with untrained leg or arm might result in greater feed back reflex induced by greater accumulation of metabolic byproducts. Specific nerve fibers within the skeletal muscle of most species can be activated by many different metabolic stimuli. Substances including

199 lactate, osmolality, and potassium that trigger these fibers appear in different concentrations in the extracellular space of trained and untrained muscle while exercise.

Actually the magnitude of heart rate increase during exercise in the trained and untrained muscles has been related to the concentrations of metabolites produced by the muscle

[318]. Thus cardiac adaptations to exercise training appear to result from central and peripheral changes. These changes include a shift in autonomic nerves system balance, structural changes in the cardiac muscle, increased in blood volume and reduced metabolic stimuli from the trained muscle.

6.2. PERIPHERAL VASCULAR ADAPTATIONS TO EXERCISE TRAINING

Resting muscle blood flow appears not to be affected by exercise training however endurance, high-intensity sprint and isometric training have been shown repeatedly to increase blood flow under varies conditions, including anticipatory state. These training effects might be the results of alterations in vascular smooth muscle and vascular endothelium structure and function, however the influence of cardiac and skeletal muscle adaptations cannot be totally excluded [428-432].

One of the early study examined the effect of exercise training on regional vascular function reported a decrease in blood flow in humans. In this study male subjects jogged 4-5km at 80% of maximum running capacity 5-6X/wk for five weeks.

Blood flows, measured immediately after maximal and submaximal exercise, were lowered following the training program. The authors attributed lowered leg blood flow to a more “effective” metabolic process [433]. Bond et al. also reported reduction in hyperemic blood flow following a training program consists 6 sets of 15 one-leg maximal isokinetic concentric contractions of planter flexion for 4 weeks while the nontrained leg

200 served as control. The authors attributed reductions in trained leg blood flow to an increase in muscle cross-sectional area without proportional changes in microvascular density [434]. More recently another study reported reduction in leg submaximal blood

flow following 9-12 wk high intensity lower body endurance training and this decreased

was coupled with decrease in submaximal oxygen consumption and unchanged cardiac

output. The authors attributed the decrease to unchanged cardiac output and blunted

splanchnic vasoconstriction [435].

Several other investigations however revealed conflicting results. For example

Klausen and colleagues reported that trained leg blood flow in six young healthy male

subjects increased during maximal and submaximal exercises following eight week one-

leg exercise training at ~85% of heart rate maximum [436]. The effect of endurance

training on the regional blood flow response to maximal exercise was also investigated in

foxhound by Musch et al. Training consisted of 8-12 wk of treadmill running at 80% of

maximal heart rate 1 h/day for 5 days/wk. Gastrocnemius muscle blood flow increased

and systemic vascular resistance decreased during maximal exercise following training

[437]. In another study by Musch et al., regional blood flow adaptations were determined

for rats after 6 days/wk for 6 wk of high-intensity sprint training or limited cage activity.

Training consisted of five 1-min bouts of treadmill running at work loads (15% grade, 97

m/min) in excess of the animals' maximal oxygen uptake interspersed with 90s of rest.

Blood flow to the different organs of the abdominal region was greatly reduced during

maximal exercise although. Furthermore blood flow to the entire hindlimb and to several

individual muscles (soleus, plantaris, gastrocnemius, flexor hallicus longus, vastus

lateralis, rectus femoris, biceps femoris, and adductor magnus and brevis muscles) was

201 greater in the exercise group during maximal exercise than in the sedentary control rats.

The authors concluded that increased cardiac output following high-intensity sprint training is primarily directed toward the working skeletal muscle and not toward the organs found in the abdominal region in rat [438]. Armstrong et al., reported increase also in preexercise anticipatory blood flow after running exercise training in rats. In fact this increase was intensity dependent, as the rats were assigned to a low-speed (daily treadmill walking up a 12 degree incline at 15 m/min), and high-speed (daily treadmill galloping up a 12 degree incline at 50 m/min) training program regimens for 2-4 wk. The authors also demonstrated that regional changes in blood flow to long-term endurance exercise training that elicits systemic adaptations were more pronounced in red muscles whereas short term elicited adaptations in only the white muscles indicating that blood flow adaptations are dependent on muscle type and the length the exercise training program [431, 439]. Blood flow was also reported to increase at onset of exercise following 10 days of cycle ergometer exercise training at 65% of maximal oxygen consumption [440]. The results of the above-cited studies are indeed very tempting to conclude that exercise training induce vascular adaptations. However the concomitant modifications in blood volume, stroke volume, heart rate, cardiac output, blood pressure, muscle metabolic capacity, and oxygen consumption demonstrate central cardiovascular involvement. Since the exercise training mode selected to examine vascular adaptations elicited systemic responses, the observed improvements in regional blood flow might be the result of increased blood volume, stroke volume, cardiac output, capillarization, A-Vd and oxygen consumption or might be basically due to improved cardiac output shunting to the active muscles rather than improved regional vascular function per say. Therefore

202 the question whether exercise training leads to regional vascular changes without central influence, remained unanswered.

TABLE LR6.1. Vascular function adaptations following small muscle group conditioning programs Study Population Type of training Results Yasuda, 83 10 HY HG, 33 & 50% MVC, 6X/wk (6 34 & 26% ⇑ BF, wks) respectively.

Sinoway, 87 6 HY HG, 70% MVC, 4X/wk, 30min (4 ⇔ BFrest. wks) 30% ⇑ RHBF in trained arm

Green, 94 11 HY HG, 75% MVC, 4X/wk, 30min (4 ⇔ BFrest wks) 23% ⇑ RHBF in trained arm

Franke, 98 6 HY HG, 70% MVC, (4 wks) ⇔FVCrest. 35% ⇑ FVCpeak. Bank, 98 11 HY & 7 HG, 70% MVC, 4X/wk, 30min, ⇔ in CHF. CHF (4-6 wks) 24% ⇑ RHBF in HY. Alomari, 01 28 HY HG, 25 & 75% MVC, 5X/wk (4 18, 29% ⇑ RHBF in trained wks) arm, respectively. BF = Blood flow; RHBF = Reactive hyperemic blood flow; VR = Vascular resistance; FVC = Forearm vascular conductance; healthy & young = HY.

In one of the early studies that examined the effect of localized exercise training on peripheral vasculature without central influence, Yasuda et al reported improvements in hyperemic forearm blood flow. The training program consisted of performing 6X/wk isometric handgrip exercise for 6 weeks at 33 and 50% of MVC to fatigue. Both training regimens resulted in similar increase in blood flow however adaptations after 6 months were greater than 3 months indicating training-period dependency [441]. Sinoway and colleagues also showed localized vascular adaptations to small muscle group training without systemic influence. In a cross-sectional study forearm blood flow, measured following 5-minutes of ischemia and 5-minutes of ischemia coupled with exercise, was higher in the dominant arm of tennis players whereas it was similar in both arms of a control group of comparable oxygen consumption. Additionally the difference in

203 hyperemic blood flow between 5-minutes ischemia and 5-minutes ischemia coupled with exercise in the dominant forearm of the tennis players was greater than the rest of the measured arms (nondominate forearm of the tennis players and both forearms of the control group) indicating greater vasodilatory reserve in the dominant forearm of the tennis players. The experiment populations and protocol allowed viewing the peripheral vasculature as an isolated local entity independent of the systemic cardiovascular involvements such as cardiac output and mean arterial blood pressure. Although the study was not proposed to reveal the mechanisms responsible for these adaptations, the authors attributed increase in dominant forearm blood flow of tennis players to peripheral vascular phenomena (increased capillarization), but also didn’t totally exclude systemic influence such as changes in sympathetic tone [428]. In a longitudinal study, Sinoway et al. further demonstrated modifications in regional vascular function following a small muscle group exercise-training program without eliciting systemic cardiovascular adaptations. The training regimen included four-week isometric handgrip exercise at 70% of maximum workload (workload was determined as the maximum amount of work the subjects could perform while maintaining 30 contractions/min for 3 minutes). The subjects exercised 4 times a week for 30 minutes and intensity was progressively increased as needed. The authors reported increase in forearm blood flow and reduction in forearm vascular resistance, measured immediately after ischemia and ischemia coupled with exercise, in the trained arm while oxygen consumption and vascular function in the untrained arm remained unchanged. The authors concluded that regional vascular modifications without systemic adaptations are attainable and attributed to enhanced intrinsic skeletal muscle vasodilatory capacity [442]. In an attempt to examine

204 the effect of training intensity on vascular function, subjects exercised at 25 and 75% of

MVC of isometric handgrip in our laboratory. The training program consisted of exercising 20 minutes 5X/wk for 4 weeks and both intensities resulted in similar increase in forearm blood flow, although the magnitude of increase was greater in the high intensity group [443].

Some investigations however have found that vascular adaptations to localized exercise training might be transferable to untrained body parts. Forearm blood flow in the ipsi- as well as the contra-lateral arms following 6 weeks were increased in the study of

Yasuda and colleagues. The authors attributed the improvements to changes in neurogenic balance however the metabolic contribution was not totally excluded [441].

Silber et al. also reported a 50% increase in forearm blood flow following leg cycle ergometer exercise however these adaptations might be the results of systemic adaptations rather than localized vascular adaptations as changes in oxygen consumption and heart rate were reported [444]. Therefore these studies have certainly demonstrated that peripheral muscle blood flow adaptation to exercise training program is a local entity independent of systemic cardiovascular modifications. However all the above studies were not designed to explain the mechanism(s) involved in peripheral vasculature adaptations although speculations attributed the observed blood flow adaptations to unclear local phenomena.

Hashimoto was the first to examine the effect of exercise training on endothelium- induced dilation [445]. The investigator reported no effect of exercise training on endothelium-dependent modulation (vasoconstriction or vasodilation) following 1- hour/day treadmill exercise training for 12 weeks in young (16wks) and old (98wks) rats.

205 However no documentations were provided to quantify the effect of exercise training

although it did induce lipoprotein profile changes (↑HDL- and ↓LDL-cholesterol). On the

other hand several other investigators reported different results. In well-designed

sequence of studies, Delp and colleagues repeatedly demonstrated training-induced

improvements in endothelium-dependent vasodilatory capacity. In one particular study

male rats ran on a treadmill at 30 m/min with 15O incline for 1 h/day 5 days/wk for 10-12

wk. Maximal vasodilator responses induced by acetylcholine were greater in vessel rings

from trained rats whereas dilatory responses by sodium nitroprusside were not different

following training, indicating that the augmented acetylcholine-induced dilatory response

resulted from endothelium adaptation. Additionally blockade of nitric oxide synthase activity, by L-NMMA, diminished acetylcholine-induced vasodilation by 100 and 79% in trained and nontrained rats, respectively [446]. Similar exercise training protocol was used to examine the effect of exercise training in hyperthyroid rats. Vasodilatory responses to acetylcholine and sodium nitroprusside were enhanced and not altered, respectively, in rat abdominal following the exercise program [447]. Vascular smooth muscle dilation is mediated by increasing cellular cyclic monophosphate, which is stimulated by activating guanylate cyclase. Acetylcholine indirectly increases the levels of cyclic monophosphate by stimulating endothelium-derived nitric oxide release, which activates smooth muscle guanylate cyclase whereas sodium nitroprusside increase the levels of cyclic monophosphate by directly activating smooth muscle guanylate cyclase. Thus the unaltered vessel response to sodium nitroprusside in the trained animals indicates that adaptations might be independent of cyclic monophosphate and related to the endothelium tissue. Additionally inhibiting nitric oxide activities, by L-NMMA,

206 completely obliterated the acetylcholine-induced vascular relaxation, which further

demonstrated the involvement of endothelium tissue and nitric oxide pathway.

Similar observations were reported by Chen and Li following only eight weeks of

treadmill exercise training. Acetylcholine- and sodium nitroprusside-induced relaxations were enhanced and unaltered, respectively, in thoracic aortae and pulmonary arteries, but not in carotid arteries from rabbits [448]. Vascular endothelium adaptations were reported following even shorter exercise program. Sun et al. found that vasodilatory responses to acetylcholine and L-NMMA were augmented in gracilis muscle of rats following treadmill exercise training as short as 2-4wks [449]. Therefore in an attempt to determine the time course of improvements in vasodilatory capacity in rats, Delp and et al. examined training-induced adaptation of vascular endothelium at different times of a 10- wk exercise-training program in rats. The exercise group was further subdivided into postexercise, 1-day, 1-wk, 2-wk, 4-wk, and 10-wk groups, whereas another group of animals didn’t exercise and served as control. The exercise program was similar to the previous studies [446, 447] and resulted in enhanced maximal vasodilation to acetylcholine and expression of endothelium nitric oxide synthase protein in aortas from

4-wk and 10-wk rats. These data indicate that the enhanced endothelium-dependent vasodilation of the rat aorta is present by 4 wk of endurance exercise training.

Additionally adaptations appear to be mediated primarily through the nitric oxide synthase pathway and are associated with an increased expression of endothelium nitric oxide synthase protein [450]. These studies clearly indicate that exercise training enhances vascular sensitivity and vasomotor function through adaptations of the endothelium as well as smooth muscle in animals. Additionally training-induced increase

207 in blood flow observed following exercise training could be partially allocated to greater vasodilatory capacity of nitric oxide. However thus far, the effect of exercise training on vascular endothelium function is not as clear in human as in animals therefore these findings still need to be confirmed in humans.

In a series of studies, Green et al. reported that enhanced peak blood flow following small muscle group exercise training is not endothelium-dependent. As previously reported [442, 443] Green and colleagues found increase in peak forearm blood flow following handgrip training program however with no changes in the vasodilatory capacity of nitric oxide [430]. Similarly they reported enhanced peak blood flow in the dominant forearm of tennis players however no evidence of difference were found in vascular endothelium function in the nondominant forearm of the tennis players and both forearms of the control group [451]. In fact the same group reported that forearm blood flow responses to L-NMMA and norepinephrine remained unchanged following 6 weeks of detraining, induced by forearm casting, indicating no detraining effect on nitric oxide pathway, despite evidence of physical deconditioning [452]. The authors attributed changes in blood flow to changes in vascular structure and vascular responsiveness without alterations in capillarization. Additionally they concluded that the lack of nitric oxide adaptations to exercise training (or detraining) as shown in animals, might be due to training protocol or species differences however didn’t exclude the contribution of other vasodilator pathways.

In contrast Kingwell and his group repeatedly observed improvements in nitric- oxide-mediated vasodilation following physical conditioning among humans. In a cross- sectional study they reported that endurance-trained athletes had greater vasodilatory

208 response to acetylcholine, greater vasoconstrictor response to norepinephrine, angiotensin

II, and L-NMMA, and similar response to sodium nitroprusside as compared to sedentary control group. Nevertheless these differences were linked to the level of serum cholesterol indicating that training-induced changes in endothelium function might be the result of reduction in cholesterol level [453]. The same research group substantiated these results in a longitudinal study as the subjects engaged in a 4-week cycle exercise-training program. Training increased maximal work-load and maximal oxygen consumption, and blood pressure was reduced. Additionally L-NMMA caused a greater vasoconstriction after training whereas there was no difference in vascular response to acetylcholine or sodium nitroprusside between pre and post training. The authors concluded that the improvements in vascular endothelium tissue are due to elevated shear stress resulting from increased blood flow and blood viscosity after cycle training [454]. More recently the same research group reported endothelium changes following home-based exercise training program for 4 weeks in hypercholesterolemic volunteers. The training program consisted of cycling 3X/wk for 3 minutes at 65% of maximum oxygen consumption and resulted in greater oxygen consumption, blood flow, vasoconstriction response to L-

NMMA along with improvement in lipid/lipoprotein profiles whereas vascular responses to acetylcholine and sodium nitroprusside didn’t change before and after training [455].

Endurance training was also found to improve nitric-oxide-mediated blood flow by several other investigations in young [456], older [457] heart failure [458], and hypertension [459, 460] participants. However the involvement of other vasodilatory pathways shouldn’t be denied since nitric oxide inhibition was found to have modest

209 effect on peak vasodilation during reactive hyperemia whereas prostaglandins appeared to play greater role [461].

Despite the overwhelming evidence of nitric-oxide-induced vasodilation improvements following exercise training the current discrepancy found between different studies need to be resolved. This discrepancy might be due to differences in training stimulus and/or experimental measurements and protocol. For example Bank et al. was the only one to report improvement in nitric-oxide-dependent vasodilation and peak hyperemic forearm blood flow in normal subjects but not in patients with heart failure following handgrip exercise training [462]. Otherwise all the studies reported changes in nitric oxide production used large muscle group exercises (e.g. running or cycling) whereas Green et al. examined vascular function following small muscle group exercise training (e.g. forearm handgrip or tennis playing). Additionally these improvements were reported in muscle beds that were not involved directly in exercise training (e.g. measuring forearm vascular function following leg exercise training), which indicate that improvements might be due to systemic adaptations. Therefore the search for other mechanism(s) that might be involved in vascular adaptations observed during exercise training is still warranted.

Increased blood flow observed following exercise training might also be due to reduction in vasoconstrictor control. Wiegman et al. reported reduction in norepinephrine-mediated vasoconstriction due to the reduction in the sensitivity to norepinephrine following 6wk-swim-training program in rats [463]. However the study was not design to determine whether the alterations in norepinephrine function are related to the endothelium or smooth muscle. Delp and colleagues found that endurance training

210 reduced vascular sensitivity to norepinephrine, in rat aortic rings, only when the

endothelium tissue was intact suggesting endothelium involvement [446]. Since α2-

adrengeric receptors are found on the endothelium tissue and mediate the release of nitric oxide or other endothelium-derived relaxing factors [382], binding of norepinephrine to the receptors on vascular endothelial cells from trained animals may induce greater released of endothelium-derive factor(s), which could blunt norepinephrine-mediated vasoconstriction. Additionally evidence indicates that endothelium tissue might inhibit the release of norepinephrine from adrenergic nerves [326] therefore exercise training

might alter the endothelium-induced inhibition of norepinephrine release.

Despite the minimum knowledge of endothelin-1 function during exercise

evidence shows that its production in the working muscles decreases while it increases

and causes vasoconstriction in the nonworking muscles. Additionally this increase is

exercise intensity- and time-dependent, contributes to norepinephrine-induced

vasoconstrictions and plays a role in blood redistribution during exercise [392, 393, 464].

However, very few studies have examined the effect of exercise training on the

endothelin-1 function and these studies are certainly far from conclusive. For example

Jones et al. found endothelin-1 adaptations to exercise training in miniature swine is gender dependent. The exercise program included treadmill running for 12 wks and

resulted in reduction of coronary artery sensitivity to endothline-1 of male swine but not

of the female [465]. From the same laboratory, Laughlin and coworkers also reported

gender-dependent adaptations however exercise training was found to reduce porcine,

femoral and brachial artery sensitivities to endothelin-1 in female swine but not in the

male [466]. In an effort to examine the effect of exercise training on endothelin-1

211 function in human, Maeda et al. examined plasma concentration of endothelin-1

following 8 wks of 1 hour cycling, 3-4 days/wk, at 70% of oxygen consumption.

Enduthline-1 plasma concentration were reduced and negatively correlated with

increased plasma concentration of nitric oxide. The observed alterations in endothelin-1 and nitric oxide concentrations were maintained for 4 weeks and didn’t return to basil

level until the 8th week after cassation of exercise training [467]. These studies suggest that exercise training might has some effect on endothelin-1 function however more research is definitely needed to expand our knowledge in the role of endothline-1- mediated modulation of the vascular function.

Therefore exercise training seems to modify the peripheral effect of vasoconstrictor factors however the mechanism by which this function is modified is not known. Additionally more studies are needed to examine the effect of exercise training modality, intensity, duration, frequency as well as training period length on the function of vasoconstrictor factors in different segments of the vasculature of various species, particularly human.

6.3. VENOUS SYSTEM ADAPTATIONS TO EXERCISE TRAINING

Improved venous return is probably the most recognized feature of the venous system that improves following exercise training. However since the effect of training on the heart, arterial system as well as blood contents have been extensively investigated, increased venous return has primarily been attributed to alterations in these aspects of the cardiovascular system, with minimum considerations given to the venous system. For example cardiac contractility and blood volume thereby stroke volume have been repeatedly found to increase following exercise training. Additionally vascular resistance,

212 vasoactivity, muscle vascular density, muscle perfusion, muscle blood flow and muscle blood volume have also been reported to be favorably altered. Since the cardiovascular system is a closed circuit, the combined exercise training-alterations within the system would certainly induce venous return changes. However recognized the venous system is a reservoir for ~65% of the total body blood and repeatedly challenged during exercise, it must be affected by exercise training. However very few studies examined the effect of exercise training on venous hemodynamics, therefore conclusions regarding the vasculature performance have for the most part ignored the contribution of the venous system adaptations to physical conditioning [32, 322, 413].

The available studies investigated sporadic aspects of the venous system. For example Suzuki and colleagues examined the effect of exercise training on venous density. Soleus muscle arteriolar and venular capillary densities in young (3 wks) and middle-age (54 wks) Wistar rats were evaluated before and after 6-wk exercise training program. The exercise protocols for the young training group were: 10-22.5 m/ min 60 min/day for 6 days a week, on 7O incline during the final 2 weeks; for the middle-aged training group, the protocols were: 10-20 m/min. 50 min/day for 6 days a week. Although it was greater in the younger group, arteriolar capillary density increased similarly in both training groups whereas the venular portion didn’t [468]. The same group of researchers examined time-course changes in soleus muscle arteriolar and venular capillary densities of young female Wisters. Running exercise training started at the age of 5 weeks and lasted for 5 weeks at 25 m/min on a 25O incline, 10-60 min/day, 5 days/week.

Morphological findings were obtained from the soleus and deep portions of the medial gastrocnemius muscles. In soleus, training resulted in increased capillary-to-fiber ratio,

213 total capillary density, and density of venular capillaries at Week 1. However values for total capillary and arteriolar and venular capillaries were significantly decreased after 4 weeks of training. In gastrocnemius muscles, training significantly increased the capillary-to-fiber ratio at Week 1 and from Week 3 onwards whereas total capillary density values showed no difference between control and training groups at all points of time. These findings suggest that exercise training in Wister rats results in greater arteriolar and venular capillarization, mainly in highly oxidative muscles, as early as 1 week. However the conflicting result between the two studies might indicate that training-induced adaptations might be time-, age- or gender-dependent. Additionally the reduction in capillary growth observed after the 4th week might suggest that other mechanisms, including vascular modulator adaptations, might be involved during a delayed phase of the exercise program. However since neither study have examined arterial or venous hemodynamics or any vascular function indicators a complete conclusion cannot be made. Finally the changes in venular capillary densities might explain the growth pattern of the capillary network associated with endurance training

[469].

Kenney and Armstrong investigated venous pooling following 40 mints, 3d/wk for 8-wk running exercise training program. Venous pooling was measured in the foot as volume drainage following “Trendeleburg” maneuver (VT). The “Trendeleburg” maneuver consisted of 45o leg elevation for 15 seconds while the subjects is laying down followed by setting up to a position of 90o flexion with the legs lower than the body. The investigators assisted the subject throughout the maneuver to avoid muscle contraction.

Moreover, to take the muscle pump contribution into account, volume drainage was

214 evaluated following 15soconds of dorsi-flexion/planter exercise while the foot is lower

than the body (VE) then subtracted from volume drainage following the Trendeleburg

maneuver (VT - VE). Exercise training resulted in increased VT, VT - VE, oxygen

consumption without changes in VE. Therefore the authors concluded that increased venous pooling is the result of hypervolemia associated while ignoring structural or functional changes in the venous system [470].

Hypervolemia, associated with exercise training, was also linked to changes in central venous pressures. Exercise program consisted of 10 wk of cycle exercise for 30 min/day, 4 days/wk at 75-80% of maximal O2 uptake and resulted in 20% increase in maximal O2 uptake, decreased resting heart rate, 9% increase in blood volume and increased central venous pressure. The increase in central blood pressure was associated with the increase in blood volume without changes in total effective venous capacitance, thus the authors attributed changes to blood volume but not to alterations in venous structure or function [471]. Louisy and coworkers also attributed changes in venous hemodynamics to hypervolemia. Leg-filling volumes during 30 head-up tilt (FV), half- emptying time (T1/2) and venous output at the 6th second of emptying during return to

horizontal position (VO6) were examined in endurance trained, strength trained and

sedentary subjects. Results showed higher values for FV and VO6 in endurance trained

subjects than strength trained and sedentary subjects. Whereas no significant differences

were observed between groups in T1/2. Therefore the authors concluded that endurance training seems to increase lower limb venous distensibility. Additionally since venous

emptying and venous return from the lower limbs were not affected increased venous

distensibility is probably a physical response to hypervolemia consecutive to chronic

215 aerobic training rather than alterations of the visco-elastic properties of walls

[472].

However other investigations reported venular structural and functional changes following exercise training. In a cross-sectional study, Wecht et al examined the difference in venous hemodynamic indices between individuals with spinal cords injury, sedentary and active individuals. Venous tone, venous capacitance, venous compliance, total venous outflow but not venous emptying rate were higher in the active group. The authors concluded that endurance exercise training enhance vascular function [473].

Miyachi et al. examined venous diameter following 6 wk one-legged cycle training at

80% peak oxygen consumption, 40 min/day for 4 days/wk followed by 6 wk of detraining. Arterial and venous cross-sectional areas increased by 16% and 46%, respectively and correlated with increase in one-leg peak oxygen consumption. These changes returned to baseline during detraining. The authors attributed part of the changes in oxygen consumption to arterial adaptations and improved blood delivery [474].

Involuntarily detraining also might affect venous function such as in morbidity diseases. Ikenouchi and colleagues found that venous distensibility is reduced and related to the severity of the disease in patients with heart failure [433, 475]. Our laboratory found that reduced venous distensibility and outflow in heart failure were related to exercise tolerance indicating inactivity influence confirming the effect of heart failure on venous function [476]. While venous changes might be a direct effect of heart failure, inactivity associated with the disease might indirectly alter venous function.

The effect of exercise training on venous function seems to be ignored despite evidences of the importance of the venous system in cardiovascular function. Therefore

216 more studies are certainly needed to examine the effect of exercise training modality, intensity, duration, frequency as well as training period length on different segments of the venous system of various species, particularly human.

217 SECTION 7. VASCULAR PATHOPHYSIOLOGY

The vessel diameter is modulated by cellular calcium ion level. Calcium ions increase in

vascular smooth cell as a result of a cascade of chemical reactions initiated by an agonist

(e.g. norepinephrine or endothelin-1) binding to a receptor located on the cell membrane.

The binding of agonists to vascular smooth muscle membrane receptors activates

phospholipase C (PLC) in a reaction coupled with guanine nucletide binding proteins (G

protein). Membrane Phospholipase C hydrolyzes phosphatidyl inositol bisphophate

(PIP2) to yield diacylglycerol (DAG) and inositol triphosphate (IP3). Diacylglycerol and

+2 +2 IP3 increase Ca by activating voltage-gated calcium channels and releasing Ca from

the sarcoplasmic reticulum, respectively. Thereafter, Ca+2 ions bind to calmodulin, which in turn binds to myosin light chain kinase. This activated Ca+2-calmodulin-myosin kinase

complex phosphorylates the light chains of myosin. The phophorylated myosin ATPase is

then activated by actin and the resulting cross-bridge cycling initiates contraction.

Relaxation on the other hand occurs when the myosin light chain phosphatases

dephosphorylate the light chains of myosin and the cytosolic Ca+2 is lowered by

sarcoplasmic reticulum uptake and by Ca+2 extrusion by the Ca pump and Na-Ca

exchanger [51, 52].

Agents elicit elevation in smooth muscle calcium ions and vasoconstriction include

catecholamines [39, 40], angiotensin II [41], arginine vasopressin [41, 42], and

endothelin-1, whereas metabolic byproduct [228, 477], nitric oxide, prostacyclin, and

endothelium hyperpolarization factor [54] inhibit the effect of these vasoconstrictors.

Therefore the endothelium-derived vasodilatory factors counteract the vasoconstrictor

effect induced mainly by the sympathetic nervous system and circulating hormones. In

218 vascular disease state these mechanisms are altered as such Ca+2 levels are constantly

elevated resulting in elevated vascular tone. Vascular disease is primarily characterized as

elevated circulating vasoconstricting agents and endothelium failure to inhibit the

contraction of smooth muscle cell. Additionally the vasculature undergoes structural

changes that result blood flow kinetics. The purpose of the section is to discuss the effect

of a variety of disease state on vascular function.

7.1. SYSTEMIC NEUROHORMONAL DYSFUNCTION IN DISEASES

Abnormalities in the neuroendocrine systems have been implicated in a verity of

cardiovascular and metabolic diseases. These abnormalities include mainly impaired

sympatho-vagal balance, elevated circulating catecholamines, and hyperactivated renin-

angiotensin-aldosterone hormonal system. Overstimulated neurohormonal systems

usually result in blood pressure rise followed by cardiac and vascular structural changes.

The increase in blood pressure is due to increased cardiac output, hypervolemia, and

increased vascular tone. This segment will focus on changes in the neuroendocrine

systems observed in heart failure and hypertension.

7.1.1. Heart failure The predominant quandary in heart failure is significant reduction in cardiac output resulting in underfilling of the arterial system. This underfilling causes activation of neurohormonal reflexes that stimulate multiple system responses. These neurohormonal reflexes are initiated by mechanoreceptors located at the high-pressure sides of the circulation including the left ventricle, carotid sinus, aortic arch, and renal afferent arterioles. Decreased activation of these receptors due to a decrease in arterial blood pressure, stroke volume, renal perfusion, or peripheral vascular resistance leads to an

increase in sympathetic nervous system, activation of the renin-angiotensin-aldosterone

219 system and release of arginine vasopressin as well as stimulation of thirst. The activation of these systems results in increased sodium and water retention, greater stroke volume, elevated cardiac pressure, fluid intake, and vasoconstriction. These responses are compensatory and mainly to maintain sufficient circulatory blood pressure and tissue perfusion, however unfortunately they may lead to further deterioration in heart function.

The activation of the sympathetic nervous system in heart failure causes increased cardiac contractility, tachycardia, arterial vasoconstriction thereby increased cardiac afterload. Sympathetic nervous system activation also indirectly triggers systemic vasoconstriction by stimulating the release of catecholamines from the adrenal medulla.

Furthermore vasoconstriction induced by increased sympathetic drive and circulating catecholamines causes reduction in renal blood flow, which stimulates the proximal convoluted tubule to activate the renin-angiotensin-aldosterone hormonal system, which also contributes to further vasoconstriction.

Plasma renin activity is increased and provides prognostic index in heart failure, additionally its activity is largely related to the severity of the disease [478]. The activated renin-angiotensin-aldosterone hormonal systems serve two compensatory purposes in the circulation. Aldosterone increases blood volume by increasing sodium and water retention while angiotensin II is a potent vasoconstrictor. This increase in body fluids is to increase the amount of blood pumped out of the left ventricle thereby increasing systemic blood pressure and tissue perfusion. Whereas angiotensin II further provokes the proximal tubular to increase aldosterone secretion in addition to its potent vasoconstrictive effect [479]. Angiotensin II also seems be involved in increased thirst sensation observed in heart failure as it may to trigger the central thirst center. However

220 angiotensin II seems to have a deleterious effect on the cardiac muscle as it causes cardiac mitogenesis resulting in cardiac remodeling that may lead to a decrease in the size of the capillary network thus predisposing the patient to ishcemic injury. Angiotensin- converting enzyme inhibitors seem attenuates the affects of angiotensin II as they may reduce the vasoconstrictor effect, reverse the myocytes proliferation, and inhibit thirst sensation [480-482].

Similar to renin-angiotensin-aldosterone hormonal system vasopressin actions include vasoconstriction and decrease in water excretion. The activities of vasopressin are elevated in heart failure contributing to the increased vasoconstriction and water retention. The stimulus for vasopressin release in heart failure is not completely known yet, though angiotensin II, osmolality and fluid volume have been implicated [483, 484].

Vasopressin appears to be the least influential vasoregulatory agent activated in heart failure though it is a very potent vasoconstrictor. It seems to be the last one activated among all of the vasoregulatory systems as its concentrations are related to the severity of the disease [485]. According to Creager et al., characterization of the contribution of the three vasoconstrictors in heart failure the sympathetic nervous system produce most of the vasoconstriction followed by renin-angiotensin system and last the arginine vasopressin [486]. Theoretically vasopressin antagonists could be useful drugs to inhibit its activities. In animal investigations as well as in healthy volunteers receptor blockers have shown promising results [487, 488]. However well controlled trials are still needed to verify the efficacy of vasopressin antagonists in heart failure.

The atrial natriuretic peptide is primarily produced by the atrial and triggered by stretching of myocardium cells. This peptide is a vasodilatory hormone but it is main

221 function is to increase sodium and water secretion from the kidney. Thus it counteracts

the effect of the other activated neurohormonal systems (i.e. sympathetic, renin-

angiotensin-aldosterone and arginine vasopressin systems). Atrial natriuretic peptide

inhibits renin-angiotensin, suppresses aldosterone and vasopressin release and prevents norepinephrine release from neurotransmitter sites [489, 490]. This hormone seems to be activated at the initial stages of heart failure as higher levels are evident in asymptomatic left ventricular dysfunction patients and before renin-angiotensin-aldosterone system is activated. Thus it appears to be essential in helping to maintain a compensated state and in inhibiting the progression of the disease at early stage [491]. Elevated atrial natriuretic factor is also found in symptomatic heart failure. In fact its level increases as heart failure worsen and is the highest in sever patients [485, 489]. However he beneficial effects of the hormone appear to be is attenuated in advanced patients as it is overwhelmed by the

increased activities of the other neurohormonal systems [492].

In patients with overt heart failure the compensatory mechanisms that initially

improve cardiac output are endless and become problematic and counterproductive as

they contribute to the progression of heart failure. The net result of neurohormonal

activation is intense vasoconstriction and expanded fluid volume as the balance is tipped in favor of vasoconstrictor and antidiuretic agents. Additionally neurohormonal activities

trigger cardiac and vascular remolding that exacerbate loading conditions on the failing

left ventricle and contribute to further decrease in myocardial contractility and the

development of overt and symptomatic failure. Antagonist and inhibitory pharmaceutical

agents seem to reduce the effect of these neurohormonal activities. Especially α- and β-

blockers and ACE-inhibitors have shown a variety of benefits including decrease

222 morbidity and mortality, improve quality of life and reduce the symptoms associated with

heart failure [493].

7.1.2. Hypertension Since a variety of systems are involved, the precise mechanism(s) responsible for

elevated blood pressure is hard to define (fig. 7-1). However the genesis of hypertension

includes one or all of the following alterations; abnormal volume expansion, increased

vascular resistance or greater cardiac output. Each one of these alterations is initiated by

abnormal neurohormonal activities and cardiovascular structural changes.

Abnormality in the reflex control of the sympathetic nervous system has been

reported. Mark et al., studied cardiopulmonary baroreceptors control of forearm vascular resistance in nine hypertensive patients (> 150/90 mmHg) and in seven normotensive.

Sympathetic nervous system activities and vascular resistance were measured while

reflex activation was provoked by applying negative pressure. The authors found

attenuation in cardiopulmonary baroreceptor reflexes characterized by increased

activation of the sympathetic nervous system and vascular resistance [494]. These

findings were confirmed by another study as abnormal arterial baroreceptor activities in

hypertensive patients were similar to those normal subjects following administrating

atropine and propranolol [495]. Impaired peripheral reflexes were also found in salt-

sensitive hypertensive patients and attributed to inadequate cardiopulmonary receptor

plasticity to increased salt intake [496]. These studies indicate abnormal sympatho-vagal

balance that contributes to vascular resistance and elevated blood pressure, additionally

these abnormalities are attributed to impaired peripheral reflexes [497]. Increased

sympathetic nervous system activation also indirectly triggers systemic vasoconstriction

by stimulating the release of catecholamines from the adrenal medulla. Furthermore

223 vasoconstriction induced by increased sympathetic drive and circulating catecholamines causes reduction in renal blood flow, which stimulates the proximal convoluted tubule to activate the renin-angiotensin-aldosterone hormonal system, which also contributes to further vasoconstriction and hypervolemia, thereby further elevation in blood pressure.

Abnormal sympatho-vagal balance also contributes to elevated blood pressure by increasing cardiac output. In borderline hypertension, the patients experience elevated cardiac output attributed to elevations in heart rate and stroke volume. Since blood volume is reduced in borderline patients the increase in stroke volume is attributed to alterations in autonomic nervous system balance characterized by increased sympathetic and attenuated parasympathetic activities. However with either aging, the disease or both, elevations in cardiac output tends to diminish reaching normal levels while vascular resistance remains elevated. The normalization of cardiac output appears to be the result of reductions in stroke volume secondary to cardiac structural changes. Continual elevations in blood pressure cause cardiac hypertrophy and decrease in cardiac compliance. Likewise prolonged blood pressure elevation leads to hypertrophy of the medial part of the vasculature leading to augmented vasoconstrictive response. The changes in vessel wall thickness cause the vessel wall to encroach even more on the vessel lumen [498].

Angiotensin II is a multifaceted hormone that contributes to hypertension in several ways. Angiotensin II is a very potent vasoconstrictor, can directly increase renal retention of salt and water and indirectly increase fluid retention by stimulating aldosterone secretion. Renin-angiotensin-aldosterone system is reportedly hyperactivated

224 in hypertensive patients. This system appears to be activated mainly as a result of

increased renal vascular resistance attributed to impaired sympatho-vagal balance [499].

7.2. ENDOTHELIUM DYSFUNCTION IN DISEASES

Vascular endothelium is a key element in maintaining balanced vasomotor regulation and

vascular homeostasis. In vascular diseases the endothelium function is compromised

leading to diminished vascular function. The main futures of endothelium dysfunction are

reduction in the efficacy of endothelium vasodilatory anticoagulant agents and increased

in the vasoconstriction capacity of the endothelium [500].

7.2.1. Atherosclerosis Endothelium dependent relaxation is impaired in atherosclerotic pigs and primates [501,

502], as well as in human coronary arteries [503]. In fact this impairment was related to

the stage of atherosclerotic development [503]. Moreover Ginsburg et al. found that the

contractile process of coronary arteries were related to the severity of the atherosclerotic

lesion not to the degree of the myocardial dysfunction [504]. In another study, vasodilation response to nitroglycerin and acetylcholine were reduced in proportion to local plaque load while this plaque load was associated with basal vascular tone. These results indicate that the vascular wall sensitivity to vasoactive agents is reduced in arteries with atherosclerotic lesion. However evident of endothelium dysfunction have also been observed in atherosclerosis as production of endothelium derived relaxing factor(s) are reduced in atherosclerotic coronary arteries obtained from patients with coronary artery disease [505].

Since atherosclerotic lesion is usually associated with other disorders including hypertension, hypercholesterolemia and diabetes mellitus it is difficult to separate the effect of atherosclerotic lesion on vascular wall function from the effect of other diseases.

225 It appears however the initial stage of vascular dysfunction in atherosclerosis is the result

of reduction in nitric oxide production, as this dysfunction appears to be essential for the

development of the atherosclerotic process. However some authors argue that as the

artery thickness and stiffness increase, it becomes harder for the endothelium-derived

relaxing factor(s) to reach the vascular smooth muscle cell. The prolong exposure to varies cardiovascular disease risk factors and the mechanical effect of sheer stress promote endothelium dysfunction. Subsequently larger and larger sections of the vessel wall become disinfected leading to less effective endothelium-derived relaxing factor(s) as well as increase in endothelium-vasoconstricting factor release. This process together with increased release of growth factors lead to further endothelium damage, wall proliferation and lesion formation initiating a vicious cycle of vessel wall deterioration

[112, 500].

7.2.2. Hypercholesterolemia/Hyperlipidemia Increased serum cholesterol levels appear to be associated with blunted vasoactivity.

After feeding rabbits 2% cholesterol diet for 13-15 wks, the inhibitory effect of

endothelium to neurogenic contraction was reduced whereas endothelium-dependent

relaxation to acetylcholine was similar to normally fed rats [506]. This inhibitory effect

was further worsen in denuded vessels [507] indicating endothelium-relaxing factor(s) is

still synthesized, released and has an effect on vascular smooth muscle in

hypercholesterolemia. However the oxidative stress of low-density protein seems to

deplete the reservoir of relaxing factors. Increased levels of low-density proteins activate

macrophages and favors formation and deposition of oxidative low-density protein in the

vascular wall. Nitric oxide seems to be either naturalized by oxidized low-density

lipoproteins or scavenged by macrophages [508, 509].

226 Endothelium-dependant relaxation has repeatedly been shown to be blunted following dietary-induced hypercholesterolemia. However this endothelium impairment is reversible by dietary manipulation. Morphological improvement of the atherosclerotic lesion, and restorations of endothelium-dependent relaxation were evident while the intima remained thickened after regression of atherosclerosis following dietary program aimed at improving lipid-lipoprotein profiles [510].

7.2.3. Diabetes mellitus Alterations in vascular wall are well documented in diabetes mellitus, leading to predisposition to other cardiovascular risk factors including hypertension and atherosclerosis. Compelling evidence strongly suggest endothelial dysfunction is manifested in this metabolic disorder. This dysfunction is characterized as blunting of endothelium-derived vasodilator effects and increased generation of vasoconstrictors such as angiotensin II, endothelin-1, and arachidonic acid products [511, 512]. Following

6 weeks of alloxan-induced diabetes mellitus in rabbits, acetylcholine relaxation was normal whereas endothelium inhibition of neurogenicaly-contracted vessels was nearly absent. Additionally neurogenic contraction in normal denuded vessels was similar to those diabetic with endothelium present indicating completely diminished vasodilatory effect of endothelium in diabetic rabbits [513].

Growth factors and cytokines agents appear to be elevated in diabetes resulting in acute increases in vascular tone, increased blood pressure and vascular remodeling that contributes to the microvascular, macrovascular, and renal complications in diabetes.

Superoxide products, overly produced in diabetics, seem to serve as signals that mediate many of the cellular biochemical reactions that result in these deleterious effects.

Diminished endothelium-dependent relaxation in diabetic rats observed in diabetic

227 vessels was significantly suppressed by pretreatment with superoxide dismutase [514].

These results indicate augmentations of superoxide products in diabetics and possible

treatments with antioxidant therapy [512].

7.2.4. Hypertension Endothelium-dependant relaxation was attenuated in genetically induced hypertensive

rats [515] as this attenuation progressively increased with age [516, 517]. The

vasodilatory response to acetylcholine was also altered in hypertensive patients [518],

independent of sodium intake [519]. Whether this attenuation is due to deficiency in

synthesis and release of endothelium-dependent relaxing factors or attenuated smooth

muscle cell responsiveness to these mediators is not clear yet. However both of these

mechanisms have been observed in human and animals with hypertension [520].

Figure LR7.1. Mechanism involved in vascular dysfunction in hypertension. Taken from [124].

Endothelin-1 may also be a factor contributing to vascular resistance in hypertensive, though its role is not clear yet. Plasma endothelin levels seem normal in most patients with essential hypertension however blood vessel wall from hypertensive animals are more sensitive and contract more profoundly in response to endothelin-1 [521].

228 The endothelium plays an obligatory role in controlling vascular function as it produces varies vasodilatory agents. It is essential to counterbalancing vasoconstriction induced by vasotonic agents released by either the autonomic nervous system or the endothelium cell. Vascular wall dysfunction observed in patients with cardiovascular risk factors is due to a decline in the inhibitory effect of the endothelium tissue. This decline might be due to either reduction in endothelium release of vasodilatory agents or failure of the vascular smooth muscle to respond appropriately to these agents, or a combination of both [522, 523] (fig 7-1).

7.3. SUMMARY

A Wide rage of changes occurs to the vascular system of patients with cardiovascular diseases including alterations in systemic and local neurohormonal stimulation, vascular responsiveness and vascular structure. The net result of these alterations usually is impaired blood flow kinetics in these patients and associated with increased fatigue and morbidity and reduced quality of life. Medications seem improve quality of life and reduce mortality and morbidity caused by the disease, however patients still experience symptoms associated with diseases.

229 SECTION 8. VASCULAR FUNCTION ASSESSMENTS

Cardiac output distribution to body parts varies during different conditions and strongly

influence whole-body hemodynamics and homeostasis. Since skeletal muscle shares a

large portion of this distribution, especially during exercise, its essential to measure

skeletal muscle blood flow under various conditions for experimental and clinical

understanding. Therefore invasive and noninvasive techniques have been developed to

examine various vascular functions including arterial inflow, vascular resistance, venous

filling capacity, venous emptying rate, vessel diameter, vessel density, and volume

changes. While there are advantages and disadvantages for each technique, testing

environment, financial feasibility, staffing and purpose of testing are all deciding factors

to which of the techniques would be used. The purpose of this section is to provide an

overview of the common techniques available for examining vascular function.

8.1. PLETHYSMOGRAPHY

8.1.1. Principles The physiological base of venous occlusion plethysmography is measuring the changes in limb volume following manipulations of the arterial and/or venous systems pressures.

Arterial inflow, vascular resistance, blood pressure, venous reflux, venous capacitance, and venous outflow rate are some the vascular functions that can be measured using this technique. For example arterial inflow is measured by calculating the rate of change in limb blood volume following venous occlusion. Whereas the change in limb blood volume following maintaining venous occlusion for a period of time (2-10 minutes) allows for estimating maximum venous filling capacity and venous emptying rate can be estimated by calculating the change in limb blood volume following venous release [524,

525].

230 8.1.2. Technicalities Various modalities of this technique have been adapted to measure vascular function

including air-filled, water-filled and strain-gauge plethysmography. However while all

these modalities have provided equally sufficient estimations of various vascular

functions, recently due to its precision, simple operative procedures and small size,

strain-gauge plethysmography has gained a lot of popularity. The strain-gauge is a sensor placed around the widest part of the limb which converts the limb volume change to change in voltage. Mercury-in-silastic strain-gauge sensor has been most used among all sensors [524].

Plethysmography can be used to measure vascular function in the forearm, calf, finger and penis however forearm and calf are generally the most used model.

Measurements are usually obtained while the subject in supine position, although standing, sitting and exercise are conditions that have been used, with a rest period (20-30 minutes) prior to evaluation of vascular functions. Vascular function can be measured in the dominant and/or nondominant limb depending on the purpose of the study. When the forearm vascular functions are measured using mercury-in-silastic sensor, cuffs are strategically positioned around the participant’s upper arm and wrist, and the strain gauge sensor is placed around the forearm approximately 10 cm distal to the olecranon process.

The forearm would be extended and slightly supinated and supported by a styrofoam block. The wrist cuff is usually used to exclude the hand circulation for the forearm measurements whereas the upper arm cuff is used to occlude the venous or/and arterial vessels. Currently suprasystole (180-240 mmHg) pressures are used to exclude the hand and occlude the arterial system, and infradiastole (30-80 mmHg) pressure is used to

occlude the venous system [443, 476, 525-528].

231 Manipulations in the pressures of the upper and wrist cuffs allows examination of

the rate of change of limb volume thought to reflect vascular function measurements,

including arterial inflow, venous capacitance and venous outflow. These changes in limb

volume are depicted as volume graphs on the plethysmograph recorder. Forearm blood

inflow rate is derived from the change in the slope drawn on the forearm-volume graph

following the first 3-5 seconds of venous occlusion. Forearm venous capacitance is

calculated as the vertical distance (mm) representing the change in forearm-volume graph

following the completion of venous filling (2-10 minutes). Whereas forearm venous outflow rate is derived from the change in the slope drawn on the forearm-volume graph between at 0.5 and 2 second following the release of venous occlusion. Forearm vascular resistance is calculated using the following equation:

Vascular resistance = Mean arterial pressure/Blood inflow.

Vascular indices are usually expressed in ml•100ml tissue-1•min-1. Forearm

vasoreactivity can be examined following 5-10 minutes of forearm arterial occlusion achieved by inflating the cuff on the upper arm to 240mmHg [443, 476, 526-528].

8.1.3. Advantages and Disadvantages Low cost and noninvasive simple operative procedures are the advantages of venous occlusion plethysmography. This technique has usually been used to measure vascular function at rest, after the cessation of exercise, and during pharmaceutical intervention.

However heighten sensitivity to subject’s movements has traditionally been problematic and required the subject to be immobilized. While this technique has been refined and improved since the original model was introduced, vascular function assessment during exercise using plethysmography is still practically impossible although some researchers

232 have reported arterial inflow during exercise [395, 396]. Additionally applying sufficient

pressure to occlude the venous system without altering the arterial kinetics is the original

fundamental assumption has to be fulfilled. However occlusion pressures (30-80 mmHg)

used in various studies is arbitrary and lack physiological basis. Additionally venous

occlusion time required for maximum venous filling, in order to measure the venous

function, is not clear resulting in inaccurate interpretations of the current studies.

Therefore studies are needed to address methodological issues of the technique to provide more accurate vascular function measurements [524, 525].

8.2. ULTRASOUND

8.2.1. Principle The images created by ultrasonic flowmetery are based on stipulations made upon high

frequency sound (ultrasound) waves reflection off blood flowing in blood vessels. These

waves are usually transmitted in a beam form by a crystal transducer to an artery or vein

to record blood particle movements. Especially with advancements in Doppler ultrasound

technology the main estimates made of the reflecting waves are the velocity and direction

of flow in the vessel. Ultrasound Doppler technology allows for continues beat-to-beat

estimation of blood flow profile [529-532].

8.2.2. Technicalities Doppler echocardiography is among the most advanced modalities available currently

however M-mode, two-dimensional and contrast echocardiography are other modalities

can also provide valuable information. Doppler ultrasound technology was introduced

during the 1950’s however a lot of improvements have been made in order to provide

sufficient information, especially regarding exercise. Properly equipped Doppler probe

233 can provide continuous estimations of blood velocity and direction and compile a flow

profile while the subject is performing physical movements [532].

The angle between the direction of blood flow and the ultrasonic beam determines

the kind of information obtained from Doppler sonagrphy. Insonation angles below 60o

while using pulsed-wave mode seem to provide accurate parabolic velocity profiles at the

V=(Fr-Fo)*(c/2Fo)*(cosθ) Where V=velocity, θ= the angle of intonation, Fr=frequency of reflected sound beam, Fo= frequency of transmitted sound beam, c=speed of sound (1540).

BF=6*104*Vmean*A Where A=cross-sectional of the vessel, Vmean=weighted mean blood velocity (m/s), constant 6*104=convert m/s to L/min.

specific depth required. However perpendicular insonation angle is essential for diameter

measurements. The direction of flow is estimated such that the sound frequency

reflection is depicted positively on the computer monitor when blood is moving toward

the transducer and vise versa when blood is moving away. Whereas blood flow is a

function of volume and time (L/min.) and derived from estimating blood velocity and

vessel cross-sectional. The following standard Doppler equations are used to calculate

blood velocity and direction: [529-532]. The diameter of the vessel increases with each

pulsation and retreat during diastolic, therefore diameter measurements are related to the

systolic (1/3) and diastolic (2/3) time periods of the cardiac cycle.

8.2.3. Advantages and Disadvantages While the measurements performed noninvasively with fairly easy procedures,

continuous blood flow profile estimation with high resolution are useful attributes of ultrasound Doppler. Rapid changes in blood flow as a result of exercise, metabolic stress, orthostatic modifications and pharmacological infusion, are currently easily detected,

234 especially with the recent technological Doppler advancements [533-535]. These

measurements are considered highly stable and have sufficient within-day and between-

days reproducibility [536, 537]. However since maintaining appropriate and consistent

insonation angle is essential for accurate measurements, the measured limb and the vessel

being measured have to be in fixed position [532].

Relatively steady laminar flow is essential for accurate blood flow measurements,

when ultrasound is used. However blood streams through different sizes and shapes of

vessels and is in constant friction against the vessel walls, thus flow is described as

parabolic profile. Flow in large vessel is flatter than in small vessel. Additionally when

blood enters a vessel the flow profile is fairly flat however as it propels for long distance

the profile becomes more parabolic. When blood departs a narrow vessel into a larger

vessel, blood flow tends to changes from laminar to a turbulent, which might result in

inaccurate blood flow measurements. Due to friction against the walls blood would flow faster in the outside of a curvy vessel, resulting in change of profile with the vessel.

Friction between blood and vessel wall also causes blood velocity in the center, of even a straight vessel lumen, to be slower than near the edges of the wall, thereby flow profile also differs. These physiological concerns are largely avoided by using anatomical landmarks within the vasculature when measurements are obtained. Therefore major conduits in the body including brachial, femoral, aortic, carotid and pulmonary arteries have been used as models to examine vascular function [529-531].

Due to higher resolution, accurate estimations and easier assessments, measurements are usually obtained in larger conduits. Blood flow regulation and responses to various physiological, physical and pharmacological perturbations are

235 different in large, medium and small vessels. Therefore current ultrasonic studies should

be interpreted cautiously and maybe limited to only large vessels.

8.3. ISOTOPE CLEARANCE

8.3.1. Principle This method is based on locally tracing the clearance of an infused molecule into the

circulation while assuming clearance is flow-dependent. Obviously this molecule is

usually easily diffusible and radio-labeled [529-531].

8.3.2. Technicalities The tracer is injected intramuscularly to the area of interest and its clearance for that area

is externally observed. The rate of clearance is depicted as curves and assumed to

correspond to blood flow in vessels. Thus tracer is cleared faster during exercise than

during rest since blood flow is also faster. Usually catheter(s) is inserted into the muscle

to inject and measure drainage rate of the tracer mixture. Since they are not toxic and

easily diffuse, Thermodilution and 133xenon are the two most common tracers used in humans [529-531].

8.3.3. Advantages and Disadvantages The potential of measuring blood flow during exercise is an extremely important and

intriguing attribute of isotope clearance rate method. However its invasive nature and the

trauma associated with the measurement requires medical training and might actually

hinder exercise performance. Additionally in some cases isotope does not mix properly

with the blood, thus results might be erroneously interpreted. Adequately mixed isotope

and positioned catheter helps ensuring proper isotope/blood mixing and minimize errors

[529-531].

236 8.4. MAGNETIC RESONANCE IMAGING ANGIOGRAPHY

8.4.1. Principle It is a relatively recently adapted technique to measure blood flow and based on phase-

contrast magnetic resonance velocity imaging of hydrogen nuclei in blood moving through a magnetic-filed accumulate a phase shift proportional to their velocity. Moving

protons are phase-shifted in proportion to their velocity the circulation whereas stationary

protons do not change their phase [529, 530].

8.4.2. Technicalities Similar to Doppler ultrasound, blood velocity and vessel cross-sectional area are

measured to calculate blood flow. The following equation is used to calculate blood flow;

Blood flow (ml/min)Vmean * A * 60 2 Where Vmean (cm/s) is the mean velocity and A (cm ) is the cross-sectional area.

Deep vessels blood flow as well as local number of vessels can be evaluated using this

technique with extremely high quality images. The images can be obtained during the

entire cardiac cycle which provide information on the nature of the pulsatile flow. The

technique is usually used during rest and hyperemic conditions since its very sensitive to

movement. Magnetic resonance imaging has been validated and used to measure vascular

function in disease population and to detect vascular pathology [529, 530, 538].

8.4.3. Advantages and disadvantages High quality images and the noninvasive procedures are the most recognized advantages

of this technique. Blood flow can be measured in deep small individual vessels in human

as well as animals. Additionally this technique does not require infusing tracing elements

or determining the insonation angle, unlike isotope tracing or ultrasound respectively.

However a high price is tagged to each vascular evaluation additionally since it is a

237 magnetic procedure patients with any metal medical aids, such as pacemaker or orthopedic metal rods are excluded [529, 530].

Many direct as well as indirect techniques are available to measure blood flow including electromagnetic flowmetery, microdialysis ethanol outflow, near spectroscopy, positron emission tomography and laser Doppler. However since it is not expensive, simple to operate and measure blood flow during a variety of conditions, venous occlusion plethysmography is probably still considered the “gold standard” for blood flow measurements [530], however Doppler ultrasound has recently gained tremendous popularity. With recent advancements, relatively simple measurement procedures, and ability to measure blood flow even while the body is moving, Doppler ultrasound has become a favorite measurement tool of vascular function [529-531].

REFERENCES

1. Van De Graaff K.M., F.S.I., Lafleur K.M., Synopsis of Human Anatomy and Physiology. 1st ed. 1997, Dubuque, IA: Wm. C. Brown Publishers. 639.

2. Smith J, K.J., Circulatory Physiology: The Essentials. 2nd ed. 1990, Baltimore, MD: Williams & Wilkins. 345.

3. Guyton AC, Text Book of Medical Physiology. 8th ed. 1991, Philadelphis, PA: W. B. Saunders Company. 1014.

4. Levick JR, An introduction to cardiovascular physiology. 2nd edition ed. 1995, Oxford: Butterworth-Heinemann Ltd.

5. Burton AC, Relation of structure to function of the tissues of the wall of blood vessels. Physiological reviews, 1954. 34(4): p. 619-641.

6. Berne R, L.M., Cardiovascular Physiology. 7th ed. 1997, St. Louis, MO: Mosby. 324.

7. Mulvany MJ, The Seventh Heymans Memorial Lecture Ghent, February 18, 1995. Physiological aspects of small arteries. Archives internationales de pharmacodynamie et de therapie., 1996. 33(1): p. 1-31.

8. Carola R, H.J., Noback CR, Human Anatomy. 1992, New York, NY: McGraw-Hill Inc.

238 9. Donald DE, S.J., Autonomic regulation of the peripheral circulation. Annual review of physiology, 1980. 42: p. 429-439.

10. Bryant M, The heart and circulatory system: A review. Journal of Nephrology Nursing, 1984. 1(3): p. 139, 159.

11. Burton AC, Role of geometry, of size and shape, in the microcirculation. Federation proceedings, 1966. 25(6): p. 1753-1760.

12. Attinger EO, Wall properties of veins. IEEE transactions on bio-medical engineering, 1969. 16(4): p. 253-261.

13. Roddie IC, Modern views on physiology. 3. The venous system. Practitioner, 1969. 202(209): p. 441-449.

14. Davis MJ, Myogenic response gradient in an arteriolar network. American journal of physiology, 1993. 264(6 Pt 2): p. H2168-2179.

15. Rowell LB, Human Circulation: Regulation During Physical Stress. 1st ed. 1986, New York, NY: Oxford University Press Inc. 416.

16. Dawson TH, Engineering design of the cardiovascular system of mammals, ed. Noordergraaf A. 1991, Englewood Cliffs, NJ: Prentice Hall.

17. Pries AR, S.T., Gaehtgens P., Design principles of vascular beds. Circulation research., 1995. 77(5): p. 1017-1023.

18. Robergs R, R.S., Exercise Physiology: Exercise, Performance, and Clinical Application. 1st ed. 1997, St. Louis, MO: Mosby Inc. 840.

19. Hainsworth R, Vascular capacitance: its control and importance. Reviews of physiology, biochemistry and pharmacology, 1986. 105: p. 101-173.

20. Rothe CF, Physiology of venous return. An unappreciated boost to the heart. Archives of internal medicine, 1986. 146(5): p. 977-982.

21. Giannattasio C, F.M., Mangoni AA, Scandola L, Fraschini N, Mancia G, Evaluation of arterial compliance in humans. Clinical and experimental hypertension, 1996. 18(3-4): p. 347-362.

22. O'Rourke MF, B.H., Intorduction to arterial compliance and function. Journal of Hypertension, 1992. 10(Suppl 6): p. S3-S5.

23. Olsen H, L.T., Reduced venous compliance in lower limbs of aging humans and its importance for capacitance function. American Journal of Physiology, 1998. 275(44): p. H878-886.

239 24. Muntinga JH, S.J., Visser KR, Zijlstra WG, Age-related differences in elastic properties of the upper arm vascular bed in healthy adults. Journal of vascular research, 1997. 4(2): p. 137-147.

25. Bevan JA, L.I., Pressure and flow-dependent vascular tone. The FASEB journal, 1991. 5(9): p. 2267-2273.

26. Perloff D, G.C., Flack J, Frohlich ED, Hill M, McDonald M, Morgenstern BZ, Human blood pressure determination by sphygmomanometry. Circulation, 1993. 88(5 Pt 1): p. 2460-2470.

27. Milnor W R, Hemodynamics. 1st ed. 1982, Baltimore, MD: Williams & Wilkins. 390.

28. Ando J, K.A., Blood flow and vascular endothelial cell function. Frontiers of medical and biological engineering, 1993. 5(4): p. 245-264.

29. Barakat AI, Responsiveness of vascular endothelium to shear stress: potential role of ion channels and cellular cytoskeleton. International journal of molecular medicine, 1999. 4(4): p. 323-332.

30. Cokelet GR, Rheology and hemodynamics. 1980. 42: p. 311-824.

31. Smith WL, M.L., DeWitt DL, Prostaglandin and thromboxane biosynthesis. Pharmacology & therapeutics, 1991. 49(3): p. 153-179.

32. Guyton AC, R.T., Langston JB., Regulation of cardiac output and venous return. Clinacal Anesthesia, 1964. 3: p. 1-34.

33. Greenway CV, L.W., Blood volume, the venous system, preload, and cardiac output. Canadian journal of physiology and pharmacology, 1986. 64(4): p. 383-387.

34. Johnson PC, Myogenic nature of increase in intestinal vascular resistance with venous pressure elevation. Circulatory research, 1959. 7: p. 992-999.

35. Shoemaker JK, N.H., Hogeman CS, Sinoway LI, Blood flow dynamics in heart failure. Circulation, 1999. 99(23): p. 3002-3008.

36. Tretbar LL, Chronic venous insufficiency of the legs: pathogenesis of venous ulcers. Journal of Enterostomal Therapy, 1987. 14(3): p. 105-108.

37. Marieb EN, M.J., Human Anatomy. 1997, Melno Park, Ca: Benjamin/Cummings.

38. Karaki H, W.G., Calcium release in smooth muscle. Life Science, 1988. 42(2): p. 111-122.

240 39. Ambler SK, T.P., Mobilization of intracellular calcium by alpha 1-adrenergic receptor activation in muscle cell monolayers. The Journal of biological chemistry, 1986. 261(13): p. 5866-5871.

40. Takuwa Y, T.N., Rasmussen H, Measurement of cytoplasmic free Ca2+ concentration in bovine tracheal smooth muscle using aequorin. American Journal of Physiology, 1987. 253(6 Pt 1): p. C817-827.

41. Nabika T, V.P., Lovenberg W, Beaven MA, Increase in cytosolic calcium and phosphoinositide metabolism induced by angiotensin II and [Arg]vasopressin in vascular smooth muscle cells. The Journal of biological chemistry, 1985. 260(8): p. 4661-4670.

42. Doyle VM, R.U., Vasopressin induced production of inositol trisphosphate and calcium efflux in a smooth muscle cell line. Biochemical and biophysical research communications, 1985. 131(1): p. 469-476.

43. Doyle VM, C.J., Ruegg UT, Hoyer D, Serotonin increases the production of inositol phosphates and mobilises calcium via the 5-HT2 receptor in A7r5 smooth muscle cells. Naunyn-Schmiedeberg's archives of pharmacology, 1986. 333(2): p. 98-103.

44. Berne RM, K.R., Ely SW, Rubio R., Adenosine in the local regulation of blood flow: a brief overview. Federation proceedings, 1983. 42(15): p. 3136-3142.

45. Case RB, F.A., Wachter M, Kyriakidis G, Castellana F, Relative effect of CO2 on canine coronary vascular resistance. Circulation research, 1978. 43(3): p. 410-418.

46. Standen NB, Q.J., K+ channel modulation in arterial smooth muscle. Acta physiologica Scandinavica., 1998. 164(4): p. 549-557.

47. Smith GL, A.C., Crichton C, Wray S, A review of the actions and control of intracellular pH in vascular smooth muscle. Cardiovascular research, 1998. 38(2): p. 316- 331.

48. Furchgott RF, Endothelium-derived relaxing factor: discovery, early studies, and identification as nitric oxide. Bioscience Reports, 1999. 19(4): p. 235-251.

49. Morimoto S, K.S., Fukuo K, Koh E, Morita R, Kitano S, Miyashita Y, Imanaka S, Ogihara T, Participation of both intracellular free Ca2+ and protein kinase C in tonic vasoconstriction induced by prostaglandin F2 alpha. European journal of pharmacology, 1990. 188(6): p. 369-378.

50. Horowitz A, M.C., Laporte R, Morgan KG, Mechanisms of smooth muscle contraction. Physiological reviews, 1996. 76(4): p. 967-1003.

51. Takuwa Y, Regulation of vascular smooth muscle contraction. The roles of Ca2+, protein kinase C and myosin light chain phosphatase. Japanese heart journal., 1996. 37(6): p. 793-813.

241 52. Van Breemen C, C.C., Leyten P, Lodge N, Saida K, Regulation of cytoplasmic Ca2+ in vascular smooth muscle. Progress in clinical and biological research, 1986. 219: p. 157-168.

53. Jackson WF, Ion Channels and Vascular Tone. Hypertension, 1999. 35(prt 2): p. 173- 178.

54. Pearson PJ, V.P., Vasodilator and vasoconstrictor substances produced by the endothelium. Reviews of physiology, biochemistry and pharmacology., 1993. 122: p. 1- 67.

55. Furchgott RF, The Discovery of Endothelium-Dependent Relaxation. Circulation, 1993. 87(Suppl. V): p. V3-V8.

56. Furchgott RF, V.P., Endothelium-derived relaxing and contracting factors. The FASEB journal, 1989. 3(9): p. 2007-2018.

57. Collins P, G.T., Henderson AH, Lewis MJ, Endothelium-derived relaxing factor alters calcium fluxes in rabbit aorta: a cyclic guanosine monophosphate-mediated effect. Journal de physiologie, 1986. 381: p. 427-437.

58. Griffith TM, E.D., Collins P, Lewis MJ, Henderson AH, Endothelium derived relaxant factor. Journal of the Royal College of Physicians of London, 1985. 19(2): p. 74- 79.

59. Yanagisawa M, M.T., Endothelin, a novel endothelum-derived peptide. Biochemical pharmacology, 1989. 34(12): p. 1877-1883.

60. Korner PI, The central nervous system and physiological mechanisms of "optimal" cardiovascular control. The Australian journal of experimental biology and medical science, 1971. 49(4): p. 319-343.

61. Edis AJ, S.J., Autonomic control of the peripheral vascular system. Archives of internal medicine, 1970. 125(4): p. 716-724.

62. Reid JL, R.P., Peptides and central neural regulation of the circulation. Physiological reviews, 1987. 67(3): p. 725-749.

63. Allert JA, Autonomic control of the cardiovascular system. Seminars in veterinary medicine and surgery (small animal), 1990. 5(1): p. 17-23.

64. Cowley AW Jr, Long-term control of arterial blood pressure. Physiological reviews, 1992. 72(1): p. 231-300.

65. Spyer KM, Central nervous integration of cardiovascular control. The Journal of experimental biology, 1982. 100: p. 109-128.

242 66. McCall RB, Neurotransmitters involved in the central regulation of the cardiovascular system. Progress in drug research, 1996. 46: p. 43-113.

67. Coleridge HM, C.J., Cardiovascular afferents involved in regulation of peripheral vessels. Annual review of physiology, 1980. 42: p. 413-427.

68. Malliani A, Cariovascular sympathetic afferent fibers. Reviews of physiology, biochemistry and pharmacology, 1980. 94: p. 11-73.

69. Hirst GD, E.F., Sympathetic neuroeffector transmission in arteries and arterioles. Physiological reviews, 1989. 69(2): p. 546-604.

70. Hilton SM, S.K., Central nervous regulation of vascular resistance. Annual review of physiology, 1980. 42: p. 399-441.

71. Hole JW Jr, Essentionls of human anatomy and physiology. 5th ed, ed. Wheatley CH. 1995, Dubuque, IA: Wm. C. Brown Publishers. 586.

72. Rampin O, G.F., Central control of the cardiovascular and erection systems: possible mechanisms and interactions. The American journal of cardiology, 2000. 86(2A): p. 19F- 22F.

73. Harris MC, The endocrinology of cardiovascular control. The Journal of endocrinology, 1988. 117(3): p. 325-328.

74. Townend JN, Angiotensin II as a modulator of cardiovascular autonomic control. Cardiologia, 1996. 41(3): p. 217-225.

75. Gibbons GH, D.V., Farhi ER, Barger AC, Interaction of signals influencing renin release. Annual review of physiology, 1984. 46: p. 291-308.

76. Katzung BG, Basic & clinical pharmacology. 7th ed, ed. Katzung BG. 1998, Stamford, CT: Appleton & Lang. 1151.

77. Monos E, B.V., Nadasy G., Local control of veins: biomechanical, metabolic, and humoral aspects. Physiological Review, 1995. 75(3): p. 611-666.

78. Reid IA, Interactions between ANG II, sympathetic nervous system, and baroreceptor reflexes in regulation of blood pressure. American journal of physiology, 1992. 262(6 Pt 1): p. E763-778.

79. Lees KR, S.I., Reid JL, The clinical pharmacology of ACE inhibitors: evidence for clinically relevant differences? Clinical and experimental pharmacology & physiology. Supplement, 1992. 19: p. 49-53.

80. Chappell MC, A.A., Ferrario CM, Pathways of angiotensin-(1-7) metabolism in the kidney. Nephrology, dialysis, transplantation, 2001. 16( Suppl 1): p. 22-26.

243 81. Chappell MC, D.D., Yunis C, Ferrario CM, Differential actions of angiotensin-(1-7) in the kidney. Kidney international. Supplement, 1998. 68: p. S3-6.

82. Bickerton RK, B.J., Evidence for a central mechanism in angiotensin induced hypertension. Proceedings of the Society for Experimental Biology and Medicine., 1961. 106: p. 834-836.

83. Yu R, D.C., Neurogenic effects of angiotensin. Lancet, 1965. 2: p. 1276-1277.

84. Lowe RD, S.G., Effects of angiotensin on the autonomic nervous system. American heart journal, 1970. 79(4): p. 562-567.

85. Jenkins TA, A.A., Chai SY, Mendelsohn FA, Interactions of angiotensin II with central catecholamines. Clinical and experimental hypertension, 1995. 17(1-2): p. 267- 280.

86. Culman J, H.S., Qadri F, Edling O, Blume A, Lebrun C, Unger T, Angiotensin as neuromodulator/neurotransmitter in central control of body fluid and electrolyte homeostasis. Clinical and experimental hypertension, 1995. 17(1-2): p. 281-293.

87. Bennett T, G.S., Involvement of vasopressin in cardiovascular regulation. Cardiovascular research, 1985. 19(2): p. 57-68.

88. Weinstein H, B.R., Sachs H, Vasopressin in blood: effect of hemorrhage. Endocrinology, 1960. 66: p. 712-718.

89. De Lima J, C.H., Beaufils M, Ardaillou R, Effects of furosemide-induced plasma volume reduction on plasma antidiuretic hormone in normal and hypertensive subjects. Clinical nephrology, 1981. 15(5): p. 246-151.

90. Rundgren M, J.H., Appelgren B, Eriksson S, Leksell LG, Vasopressin release in response to acute hypotension induced at different time intervals in the conscious sheep. Acta physiologica Scandinavica, 1984. 121(4): p. 393-399.

91. Ledsome JR, Atrial receptors, vasopressin and blood volume in the dog. Life sciences, 1985. 36(14): p. 1315-1330.

92. Cowley AW Jr, Vasopressin and blood pressure regulation. Clinical physiology and biochemistry, 1988. 6(3-4): p. 150-162.

93. Mohama RE, J.W., Gilmore JP, Comparative reactivity of hamster cheek pouch microvessels to arginine vasopressin and angiotensin II. Microcirculation, endothelium, and lymphatics, 1984. 1(4): p. 397-413.

94. Liard JF, Vasopressin in cardiovascular control: role of circulating vasopressin. Clinical science, 1984. 67(5): p. 473-481.

244 95. Bartelstone HJ, N.P., Vasopressin potentiation of catecholamine action in dog, rat, cat and rat aortic strip. American journal of physiology, 1965. 208: p. 754-762.

96. Guyton AC, Circulatory physiology: cardiac output and its regulation. 1963, Philadelphia: Saunders.

97. Johnson PC, Autoregulation of blood flow. Circulation research, 1986. 59(5): p. 483- 495.

98. Bishop VS, Central nervous system regulation of cardiovascular homeostasis. New horizons, 1994. 2(4): p. 415-418.

99. Bishop VS, Carotid baroreflex control of blood pressure and heart rate in men during dynamic exercise. Journal of applied physiology, 1994. 77(2): p. 491-492.

100. Krieger EM, B.P., Negrao CE, Role of arterial baroreceptor function on cardiovascular adjustments to acute and chronic dynamic exercise. Biological research, 1998. 31(3): p. 273-279.

101. Marshall JM, Peripheral chemoreceptors and cardiovascular regulation. Physiological Review, 1994. 74(3): p. 543-594.

102. Vanhoutte PM, Endothelium and control of vascular function. State of the Art lecture. Hypertension, 1989. 13((6 Pt 2)): p. 658-67.

103. Vanhoutte PM, S.H., Endothelium-derived relaxing factor and coronary vasospasm. Circulation, 1989. 80(1): p. 1-9.

104. Furchgott RF, Z.J., The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature, 1980. 288(5789): p. 373-376.

105. Moncada S, V.J., Pharmacology and endogenous roles of prostaglandin endoperxides, thromoxane A2 and proctacyclin. Pharmacology Review, 1979. 30: p. 293.

106. Bolton TB, C.L., Endothelial-dependent relaxant actions of carbachol and substance P in arterial smooth muscle. British journal of pharmacology, 1986. 87(4): p. 713-723.

107. Feletou M, V.P., Endothelium-dependent hyperpolarization of canine coronary smooth muscle. British journal of pharmacology, 1988. 93(3): p. 515-524.

108. Komori K, V.P., Endothelium-derived hyperpolarizing factor. Blood Vessels, 1990. 27(2-5): p. 238-245.

109. Luscher TF, B.C., Dohi Y, Yang ZH,, Endothelium-derived contracting factors. Hypertension, 1992. 19(2): p. 117-130.

110. Doyle VM, C.J., Ruegg UT, Platelet-activating factor mobilizes intracellular calcium in vascular smooth muscle cells. FEBS letters, 1986. 197(1-2): p. 13-16.

245 111. Scott-Burden T, V.P., Regulation of smooth muscle cell growth by endothelium- derived factors. Texas Heart Institute journal, 1994. 21(1): p. 91-97.

112. Vanhoutte PM, S.-B.T., The endothelium in health and disease. Texas Heart Institute journal, 1994. 21(1): p. 62-67.

113. Bank AJ, R.T., Tschumperlin LK, Kraemer MD, Letourneau JG, Kubo SH, Endothelium-dependent vasodilation of peripheral conduit arteries in patients with heart failure. Journal of Cardiac Failure, 1994. 1(1): p. 35-43.

114. Cherry PD, F.R., Zawadzki JV, Jothianandan D, Role of endothelial cells in relaxation of isolated arteries by bradykinin. Proceedings of the National Academy of Sciences of the United States of America, 1982. 79(6): p. 79(6):2106-10.

115. Katz SD, S.M., Yuen J, LeJemtel TH, Impaired acetylcholine-mediated vasodilation in patients with congestive heart failure. Role of endothelium-derived vasodilating and vasoconstricting factors. Circulation, 1993. 88(1): p. 55-61.

116. Katz SD, B.L., Sabba C, Strom JA, Jondeau G, Galvao M, Solomon S, Nikolic SD, Forman R, LeJemtel TH, Impaired endothelium-mediated vasodilation in the peripheral vasculature of patients with congestive heart failure. Journal of the American College of Cardiology, 1992. 19(5): p. 918-925.

117. Jelliffe RW, Dilator. The Journal of pharmacology and experimental therapeutics, 1962. 135: p. 349-353.

118. Somlyo AP, S.A., Vascular smooth muscle. II. Pharmacology of normal and hypotensive vessels. Pharmacological reviews, 1970. 22(2): p. 249-353.

119. Toda N, The action of vasodilating drugs on isolated basilar, coronary and mesenteric arteries of the dog. The Journal of pharmacology and experimental therapeutics., 1974. 191(1): p. 139-46.

120. Steinsland OS, F.R., Kirpekar SM, Inhibition of adrenergic neurotransmission by parasympathomimetics in the rabbit ear artery. The Journal of pharmacology and experimental therapeutics, 1973. 184(2): p. 346-356.

121. Gruetter CA, B.B., McNamara DB, Gruetter DY, Kadowitz PJ, Ignarro L, Relaxation of bovine coronary artery and activation of coronary arterial guanylate cyclase by nitric oxide, nitroprusside and a carcinogenic nitrosoamine. Journal of cyclic nucleotide research, 1979. 5(3): p. 211-224.

122. Ignarro LJ, B.G., Byrns RE, Wood KS, Chaudhuri G, Endothelium-derived relaxing factor and nitric oxide possess identical pharmacologic properties as relaxants of bovine arterial and venous smooth muscle. The Journal of pharmacology and experimental therapeutics, 1988. 246(1): p. 218-226.

246 123. Fleming I, B.R., NO: the primary EDRF. Journal of molecular and cellular cardiology, 1999. 31(1): p. 5-14.

124. Shepherd JT, K.Z., Endothelium-derived vasoactive factors: I. Endothelium- dependent relaxation. Hypertension, 1991. 18(5 Suppl): p. III76-85.

125. Sakuma I, S.D., Gross SS, Nathan C, Levi R, Identification of arginine as a precursor of endothelium-derived relaxing factor. Proceedings of the National Academy of Sciences of the United States of America, 1988. 85(22): p. 8664-8667.

126. Palmer RM, A.D., Moncada S, Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature, 1988. 333(6174): p. 664-666.

127. Schmidt HH, N.H., Wittfoht W, Gerlach J, Prescher KE, Klein MM, Niroomand F, Bohme E, Arginine is a physiological precursor of endothelium-derived nitric oxide. European journal of pharmacology, 1988. 2(154): p. 213-216.

128. Ignarro LJ, K.P., The pharmacological and physiological role of cyclic GMP in vascular smooth muscle relaxation. Annual review of pharmacology and toxicology, 1985. 25: p. 171-191.

129. Ratz PH, G.M., Flaim SF, Simultaneous measurement of force and calcium uptake during acetylcholine-induced endothelium-dependent relaxation of rabbit thoracic aorta. Circulation research, 1987. 60(1): p. 31-38.

130. Popescu LM, P.C., Hinescu M, Nutu O, The mechanism of cGMP-induced relaxation in vascular smooth muscle. European journal of pharmacology, 1985. 107(3): p. 393-394.

131. Radomski MW, P.R., Moncada S, The anti-aggregating properties of vascular endothelium: interactions between prostacyclin and nitric oxide. British journal of pharmacology, 1987. 92(3): p. 639-646.

132. Collier JG, L.R., Robinson BF, Comparison of effects of tolmesoxide (RX71107), diazoxide, hydrallazine, prazosin, glyceryl trinitrate and sodium nitroprusside on forearm arteries and dorsal hand veins of man. British journal of clinical pharmacology, 1978. 5(1): p. 35-44.

133. Aellig WH, Pharmacological experiments in healthy volunteers with bopindolol, a long-acting beta-adrenoceptor blocking drug with partial agonist activity. British journal of clinical pharmacology, 1985. 19(6): p. 775-781.

134. Furuta T, S.T., A regional difference in endothelium-dependent relaxation responses to acetylcholine in the canine venous system. Japanese journal of pharmacology, 1987. 44(2): p. 207-210.

135. Lui AS, A.J., Difference between peripheral venous volume-pressure responses to sublingual nitroglycerine. Clinical pharmacology and therapeutics., 1992. 51(3): p. 171.

247 136. Marsden PA, G.M., Brenner BM, Endothelial cell biology in relation to current concepts of vessel wall structure and function. Journal of the American Society of Nephrology, 1991. 1(7): p. 931-948.

137. Gryglewski RJ, Prostacyclin and nitric oxide. Acta haematologica Polonica, 1994. 25(2 Suppl 2): p. 75-81.

138. Brandt R, D.-K.A., Korbut R, Gryglewski RJ, Nowak J, Release of prostacyclin from the human pulmonary vascular bed in response to cholinergic stimulation. Naunyn- Schmiedeberg's archives of pharmacology, 1984. 325(1): p. Naunyn Schmiedebergs Arch Pharmacol 1984 Jan;325(1):69-75.

139. Gryglewski RJ, Secretory function of vascular endothelium and its implication in the action of vasoactive drugs. Polish journal of pharmacology and pharmacy, 1990. 42(3): p. Pol J Pharmacol Pharm 1990 May;42(3):219-236.

140. Gryglewski RJ, S.J., Korbut R, Endogenous mechanisms that regulate prostacyclin release. Advances in prostaglandin and thromboxane research, 1980. 7: p. 777-787.

141. Pattison A, E.C., Bonner FW, Nicotinic acid enhances the production of 6- ketoprostaglandin F1 alpha in human whole blood in vitro. Research communications in chemical pathology and pharmacology, 1987. 55(3): p. 423-426.

142. Quadt JF, V.R., ten Hoor F, Prostacyclin production of the isolated pulsatingly perfused rat aorta. Journal of pharmacological methods, 1982. 7(4): p. 263-270.

143. Gryglewski RJ, P.B., Korbut R, Grodzinska L, Ocetkiewicz A, Corticosteroids inhibit prostaglandin release from perfused mesenteric blood vessels of rabbit and from perfused lungs of sensitized guinea pig. Prostaglandins, 1975. 10(2): p. Prostaglandins 1975 Aug;10(2):343-355.

144. Flower RJ, W.J., Parente W, Macrocortin and the mechanism of action glucocorticosteroids. Adv Inflammation Res, 1984. 7: p. 61-70.

145. Henriksen T, E.S., Carlander B, Injury to cultured endothelial cells induced by low density lipoproteins: protection by high density lipoproteins. Scandinavian journal of clinical and laboratory investigation, 1979. 39(4): p. 369-375.

146. Nolan RD, E.T., Inhibition of prostacyclin synthesis in cultured bovine aortic endothelial cells by vitamin K1. Biochemical pharmacology, 1986. 35(23): p. 4273-4281.

147. Siegel MI, M.R., Cuatrecasas P, Aspirin-like drugs interfere with arachidonate metabolism by inhibition of the 12-hydroperoxy-5,8,10,14-eicosatetraenoic acid peroxidase activity of the lipoxygenase pathway. Proceedings of the National Academy of Sciences of the United States of America, 1979. 76(8): p. 3774-3778.

148. Luscher TF, T.F., Endothelial regulation of vascular tone and growth. American journal of hypertension, 1993. 6(7 Pt 2): p. 283S-293S.

248 149. Sinzinger H, F.P., Prostacyclin (PGI2) contracts normal and varicose human saphenous veins. VASA, 1984. 13(3): p. 228-230.

150. Hiremath AN, P.R., Hoffman BB, Blaschke TF, Comparison of age-related changes in prostaglandin E1 and beta-adrenergic responsiveness of vascular smooth muscle in adult males. Journal of gerontology, 1989. 4(1): p. M13-17.

151. Collier JG, K.S., Robinson B, Somers K, Proceedings: Action of prostaglandins A2, B-1, E2 and F2 alpha on superficial hand veins of man. British journal of pharmacology, 1972. 44(2): p. 374P-375P.

152. Robinson BF, C.J., Karim SM, Somers K, Effect of prostaglandins A 1 , A 2 , B 1 , E 2 and F 2 on forearm arterial bed and superficial hand veins in man. Clinical science, 1973. 44(4): p. 367-376.

153. Kuriyama H, S.H., The effects of acetylcholine on the membrane and contractile properties of smooth muscle cells of the rabbit superior mesenteric artery. British journal of pharmacology, 1978. 64(4): p. 493-501.

154. Bolton TB, L.R., Takewaki T, Mechanisms of action of noradrenaline and carbachol on smooth muscle of guinea-pig anterior mesenteric artery. Journal of physiology, 1984. 351: p. 549-572.

155. Komori K, L.R., Vanhoutte PM, Nitric oxide, ACh, and electrical and mechanical properties of canine arterial smooth muscle. American journal of physiology, 1988. 255((1 Pt 2)): p. H207-212.

156. Huang AH, B.R., Bassenge E, Endothelium-dependent hyperpolarization of smooth muscle cells in rabbit femoral arteries is not mediated by EDRF (nitric oxide). Naunyn- Schmiedeberg's archives of pharmacology, 1988. 388(4): p. Naunyn Schmiedebergs Arch Pharmacol 1988 Oct;338(4):438-42.

157. Komori K, S.H., Electrical responses of smooth muscle cells during cholinergic vasodilation in the rabbit saphenous artery. Circulation research, 1987. 16(4): p. 586-593.

158. Nakashima M, M.J., Taylor AA, Vanhoutte PM, Endothelium-dependent hyperpolarization caused by bradykinin in human coronary arteries. The Journal of clinical investigation, 1993. 92(6): p. 2867-2871.

159. Twort C, V.B.C., Human airway smooth muscle in culture. Tissue and cell, 1988. 20(3): p. 339-344.

160. Beny JL, B.P., Huggel H, Effect of mechanical stimulation, substance P and vasoactive intestinal polypeptide on the electrical and mechanical activities of circular smooth muscles from pig coronary arteries contracted with acetylcholine: role of endothelium. Pharmacology, 1986. 33(2): p. 61-68.

249 161. Beny JL, B.P., Huggel H, Effects of substance P, calcitonin gene-related peptide and capsaicin on tension and membrane potential of pig coronary artery in vitro. Regulatory peptides, 1989. 25(1): p. 25-36.

162. Brayden JE, W.G., Endothelium-dependent dilation of feline cerebral arteries: role of membrane potential and cyclic nucleotides. Journal of cerebral blood flow and metabolism, 1989. 9(3): p. 256-263.

163. Nakashima M, V.P., Age-dependent decrease in endothelium-dependent hyperpolarizations to endothelin-3 in the rat mesenteric artery. Journal of cardiovascular pharmacology, 1993. 22(Suppl 8): p. S352-354.

164. Nakashima M, V.P., Endothelins and membrane potential of vascular smooth muscle of the canine coronary artery. Journal of cardiovascular pharmacology, 1993. 22(Suppl 8): p. J Cardiovasc Pharmacol 1993;22 Suppl 8:S229-231.

165. Nakashima M, V.P., Endothelin-1 and -3 cause endothelium-dependent hyperpolarization in the rat mesenteric artery. American journal of physiology, 1993. 265(6 Pt 2): p. H2137-2141.

166. Chen G, S.H., Calcium dependency of the endothelium-dependent hyperpolarization in smooth muscle cells of the rabbit carotid artery. Journal of physiology, 1990(421): p. 521-534.

167. Cohen RA, V.P., Endothelium-dependent hyperpolarization. Beyond nitric oxide and cyclic GMP. Circulation, 1995. 92(11): p. 3337-3349.

168. Nagao T, I.S., Vanhoutte PM, Calmodulin antagonists inhibit endothelium- dependent hyperpolarization in the canine coronary artery. British journal of pharmacology, 1992. 107(2): p. 382-386.

169. Illiano S, N.T., Vanhoutte PM, Calmidazolium, a calmodulin inhibitor, inhibits endothelium-dependent relaxations resistant to nitro-L-arginine in the canine coronary artery. British journal of pharmacology, 1992. 107(2): p. 387-392.

170. Chen G, S.H., Some electrical properties of the endothelium-dependent hyperpolarization recorded from rat arterial smooth muscle cells. Journal of physiology, 1989. 410: p. 91-106.

171. Chen G, S.H., Weston AH, Acetylcholine releases endothelium-derived hyperpolarizing factor and EDRF from rat blood vessels. British journal of pharmacology, 1988. 95(4): p. 1165-1174.

172. Taylor SG, S.J., Weston AH, Baker JR, Endothelium-dependent effects of acetylcholine in rat aorta: a comparison with sodium nitroprusside and cromakalim. British journal of pharmacology, 1988. 94(3): p. 853-863.

250 173. Chen G, Y.Y., Miwa K, Suzuki H, Hyperpolarization of arterial smooth muscle induced by endothelial humoral substances. American journal of physiology, 1990. 260((6 Pt 2)): p. H1888-1892.

174. Nelson MT, P.J., Worley JF, Standen NB, Calcium channels, potassium channels, and voltage dependence of arterial smooth muscle tone. American journal of physiology, 1990. 259(1 Pt 1): p. C3-18.

175. Feletou M, V.P., The third pathway: endothelium-dependent hyperpolarization. Journal of physiology and pharmacology, 1999. 50(4): p. 525-534.

176. Bossu JL, E.A., Feltz A, Tanzi F, Aunis D, Thierse D, Voltage-gated Ca entry in isolated bovine capillary endothelial cells: evidence of a new type of BAY K 8644- sensitive channel. Pflugers Archiv : European journal of physiology, 1992. 420(2): p. 200-207.

177. Chen G, Y.Y., Miwa K, Suzuki H, Hyperpolarization of arterial smooth muscle induced by endothelial humoral substances. American journal of physiology, 1991. 260((6 Pt 2)): p. H1888-1892.

178. Rand VE, G.C., Endothelium-dependent relaxation to acetylcholine in the rabbit basilar artery: importance of membrane hyperpolarization. British journal of pharmacology, 1992. 106(1): p. 143-150.

179. Jackson WF, K.A., Dambacher T, Busse R, Prostacyclin-induced vasodilation in rabbit heart is mediated by ATP-sensitive potassium channels. American journal of physiology, 1993. 264(1 Pt 2): p. H238-243.

180. Clapp LH, T.S., Hall S, Baloch M, Evidence that Ca2+-activated K+ channels play a major role in mediating the vascular effects of iloprost and cicaprost. European journal of pharmacology, 1998. 356(2-3): p. 215-224.

181. Hecker M, B.A., Bauersachs J, Busse R, Characterization of endothelium-derived hyperpolarizing factor as a cytochrome P450-derived arachidonic acid metabolite in mammals. Journal of physiology, 1994. 481(Pt 2): p. 407-414.

182. Campbell WB, G.D., Pratt PF, Harder DR, Identification of epoxyeicosatrienoic acids as endothelium-derived hyperpolarizing factors. Circulation research, 1994. 78(3): p. 415-423.

183. Nelson MT, Q.J., Physiological roles and properties of potassium channels in arterial smooth muscle. American journal of physiology, 1995. 268(4 Pt 1): p. 268(4 Pt 1):C799-822.

184. Prior HM, W.N., Quinn K, Beech DJ, Yates MS, K(+)-induced dilation of a small renal artery: no role for inward rectifier K+ channels. Cardiovascular research, 1998. 37(3): p. 780-790.

251 185. Corriu C, F.M., Canet E, Vanhoutte PM, Inhibitors of the cytochrome P450-mono- oxygenase and endothelium-dependent hyperpolarizations in the guinea-pig isolated carotid artery. British journal of pharmacology, 1996. 117(4): p. 607-610.

186. Urakami-Harasawa L, S.H., Nakashima M, Egashira K, Takeshita A, Importance of endothelium-derived hyperpolarizing factor in human arteries. The Journal of clinical investigation, 1997. 100(11): p. 2793-2799.

187. Graier WF, H.S., Hoebel BG, Kukovetz WR, Kostner GM, Mechanisms of L-NG nitroarginine/indomethacin-resistant relaxation in bovine and porcine coronary arteries. British journal of pharmacology, 1996. 119(6): p. 1177-1186.

188. Shimokawa H, Y.H., Fujii K, Owada MK, Nakaike R, Fukumoto Y, Takayanagi T, Nagao T, Egashira K, Fujishima M, Takeshita A, The importance of the hyperpolarizing mechanism increases as the vessel size decreases in endothelium-dependent relaxations in rat mesenteric circulation. Journal of cardiovascular pharmacology, 1996. 28(5): p. 703- 711.

189. Mantelli L, A.S., Ledda F, Roles of nitric oxide and endothelium-derived hyperpolarizing factor in vasorelaxant effect of acetylcholine as influenced by aging and hypertension. Journal of cardiovascular pharmacology, 1995. 25(4): p. 595-602.

190. Najibi S, C.C., Palacino JJ, Cohen RA, Enhanced role of potassium channels in relaxations to acetylcholine in hypercholesterolemic rabbit carotid artery. American journal of physiology, 1994. 266(5 Pt 2): p. 266(5 Pt 2):H2061-7.

191. Fukao M, H.Y., Kanno M, Sakuma I, Kitabatake A, Alterations in endothelium- dependent hyperpolarization and relaxation in mesenteric arteries from streptozotocin- induced diabetic rats. British journal of pharmacology, 1997. 121(7): p. 1383-1391.

192. De Mey JG, V.P., Heterogeneous behavior of the canine arterial and venous wall. Importance of the endothelium. Circulation research, 1982. 51(4): p. 439-447.

193. Pearson PJ, L.P., Schaff HV, Vanhoutte PM, Augmented endothelium-dependent constriction to hypoxia early and late following reperfusion of the canine coronary artery. Clinical and experimental pharmacology & physiology, 1996. 23(8): p. 634-641.

194. Katusic ZS, S.J., Endothelium-derived vasoactive factors: II. Endothelium- dependent contraction. Hypertension, 1991. 18(5 Suppl): p. III86-92.

195. Hickey KA, R.G., Paul RJ, Highsmith RF, Characterization of a coronary vasoconstrictor produced by cultured endothelial cells. American journal of physiology, 1985. 248(5 Pt 1): p. C550-556.

196. Gillespie MN, O.J., McMurtry IF, O'Brien RF, Sustained coronary vasoconstriction provoked by a peptidergic substance released from endothelial cells in culture. The Journal of pharmacology and experimental therapeutics, 1986. 236(2): p. 339-343.

252 197. Inoue A, Y.M., Kimura S, Kasuya Y, Miyauchi T, Goto K, Masaki T, The human endothelin family: three structurally and pharmacologically distinct isopeptides predicted by three separate genes. Proceedings of the National Academy of Sciences of the United States of America, 1989. 86(8): p. 2863-2867.

198. Yanagisawa M, K.H., Kimura S, Tomobe Y, Kobayashi M, Mitsui Y, Yazaki Y, Goto K, Masaki T, A novel potent vasoconstrictor peptide produced by vascular endothelial cells. Nature, 1988. 332(6163): p. 411-415.

199. Emori T, H.Y., Ohta K, Kanno K, Eguchi S, Imai T, Shichiri M, Marumo F, Cellular mechanism of endothelin-1 release by angiotensin and vasopressin. Hypertension, 1991. 18(2): p. 165-170.

200. Boulanger C, L.T., Release of endothelin from the porcine aorta. Inhibition by endothelium-derived nitric oxide. The Journal of clinical investigation, 1990. 85(2): p. 587-590.

201. Prins BA, H.R., Nazario B, Pedram A, Frank HJ, Weber MA, Levin ER, Prostaglandin E2 and prostacyclin inhibit the production and secretion of endothelin from cultured endothelial cells. The Journal of biological chemistry, 1994. 269(16): p. 11938- 11944.

202. Emori T, H.Y., Imai T, Eguchi S, Kanno K, Marumo F, Cellular mechanism of natriuretic peptides-induced inhibition of endothelin-1 biosynthesis in rat endothelial cells. Endocrinology, 1993. 133(6): p. 2474-2480.

203. Imai T, H.Y., Emori T, Marumo F, Heparin has an inhibitory effect on endothelin-1 synthesis and release by endothelial cells. Hypertension, 1993. 21(3): p. 353-358.

204. Hosoda K, N.K., Hiroshi-Arai, Suga S, Ogawa Y, Mukoyama M, Shirakami G, Saito Y, Nakanishi S, Imura H, Cloning and expression of human endothelin-1 receptor cDNA. FEBS letters, 1991. 287(1-2): p. 23-26.

205. Sakamoto A, Y.M., Sakurai T, Takuwa Y, Yanagisawa H, Masaki T, Cloning and functional expression of human cDNA for the ETB endothelin receptor. Biochemical and biophysical research communications, 1991. 178(2): p. 656-663.

206. Moreland S, M.D., Delaney CL, Lee VG, Hunt JT, Venous smooth muscle contains vasoconstrictor ETB-like receptors. Biochemical and biophysical research communications, 1992. 184(1): p. 100-106.

207. Summner MI, C.T., Mundin JW, White DG, Watts IS, Endothelin ETA and ETB receptors mediate vascular smooth contraction. British journal of pharmacology, 1992. 107: p. 858-860.

208. Warner TD, M.J., de Nucci G, Vane JR, Endothelin-1 and endothelin-3 release EDRF from isolated perfused arterial vessels of the rat and rabbit. Journal of cardiovascular pharmacology, 1989. 13( Suppl 5): p. S85-88.

253 209. Filep JG, F.-F.E., Rousseau A, Sirois P, Fournier A, Vascular responses to endothelin-1 following inhibition of nitric oxide synthesis in the conscious rat. British journal of pharmacology, 1993. 110(3): p. 1213-1221.

210. Van Renterghem C, V.P., Barhanin J, Schmid-Alliana A, Frelin C, Lazdunski M, Molecular mechanism of action of the vasoconstrictor peptide endothelin. Biochemical and biophysical research communications, 1988. 157(3): p. 977-985.

211. Highsmith RF, P.D., Rapoport RM, Endothelial cell-derived vasoconstrictors: mechanisms of action in vascular smooth muscle. Journal of cardiovascular pharmacology, 1989. 13(Suppl 5): p. S36-44.

212. Liu CY, S.M., Attenuation of endothelin-1-induced calcium response by tyrosine kinase inhibitors in vascular smooth muscle cells. American journal of physiology, 1996. 270(6 Pt 1): p. C1825-1833.

213. Golfman LS, H.T., Beamish RE, Dhalla NS, Role of endothelin in heart function in health and disease. The Canadian journal of cardiology, 1993. 9(7): p. 635-653.

214. Bobik A, G.A., Millar JA, Mitchell A, Grinpukel S, Growth factor activity of endothelin on vascular smooth muscle. American journal of physiology, 1990. 258(3 Pt 1): p. C408-415.

215. Haynes WG, W.D., Endothelin as a regulator of cardiovascular function in health and disease. Journal of hypertension, 1998. 16(8): p. 1081-1098.

216. Kedzierski RM, Y.M., Endothelin system: the double-edged sword in health and disease. Annual review of pharmacology and toxicology, 2001. 41: p. 851-876.

217. Miller VM, K.K., Burnett JC Jr, Vanhoutte PM., Differential sensitivity to endothelin in canine arteries and veins. American journal of physiology. Heart and circulatory physiology., 1989. 257(4 Pt 2): p. H1127-1131.

218. Cocks TM, B.A., Dib M, Sudhir K, Angus JA, Endothelin is blood vessel selective: studies on a variety of human and dog vessels in vitro and on regional blood flow in the conscious rabbit. Clinical and experimental pharmacology & physiology, 1989. 16(4): p. 243-246.

219. McNamara DB, L.J., Minkes RK, Saroyan RM, Kvamme P, Rowe N, Webb WR, Fox L, Kerstein MD, Mills NL, et al, Comparative effects of endothelin (ET-1) and U46619 on human saphenous vein and gastroepiploic artery, sources of human autologous grafts. Molecular and cellular biochemistry, 1992. 117(1): p. 81-85.

220. Clarke JG, B.N., Larkin SW, Webb DJ, Davies GJ, Maseri A, Endothelin is a potent long-lasting vasoconstrictor in men. The American journal of physiology, 1989. 257(6 Pt 2): p. H2033-2035.

254 221. Haynes WG, C.J., Cockcroft JR, Webb DJ, Pharmacology of endothelin-1 in vivo in humans. Journal of cardiovascular pharmacology, 1991. 17(Suppl 7): p. J Cardiovasc Pharmacol 1991;17 Suppl 7:S284-286.

222. Haynes WG, W.D., Venoconstriction to endothelin-1 in humans: role of calcium and potassium channels. The American journal of physiology, 1993. 265(5 Pt 2): p. H1676- 1681.

223. Haefeli WE, L.L., Kiowski W, Luscher TF,, In vevo properties of endothelin-1 and - 3 in hand veins and its reversal by bradykinin and verapamil. Hypertension, 1993. 22(4): p. 434.

224. Haefeli WE, L.L., Kiowski W, Noll G, Luscher TF, Comarative action of enodtheline-1 and -3 and alpha adrenergic agonists in human veins in vivo. European heart journal supplements, 1993. 14(Suppl): p. 310.

225. Nakao K, I.Y., Oike M, Kitamura K, Kuriyama H, Mechanisms of endothelin- induced augmentation of the electrical and mechanical activity in rat portal vein. Pflugers Archiv, 1990. 415(5): p. 526-532.

226. Yang Z, B.E., von Segesser L, Stulz P, Turina M, Luscher TF, Different mobilization of calcium in endothelin-1-induced contractions in human arteries and veins: effects of calcium antagonists. Journal of cardiovascular pharmacology, 1990. 16(4): p. 654-660.

227. Olsson RA, Local factors regulating cardiac and skeletal muscle blood flow. Annual review of physiology, 1981. 43: p. 385-395.

228. Borgstrom P, G.S., Integrated myogenic and metabolic control of vascular tone in skeletal muscle during autoregulation of blood flow. Microvascular research, 1987. 33(3): p. 353-376.

229. Eklof B, L.N., Nilsson L, Norberg K, Siesjo BK, Blood flow and metabolic rate for oxygen in the cerebral cortex of the rat. Acta physiologica Scandinavica, 1973. 88(4): p. 587-589.

230. Rumsey WL, S.C., Nuutinen EM, Robiolio M, Wilson DF, Cellular energetics and the oxygen dependence of respiration in cardiac myocytes isolated from adult rat. The Journal of biological chemistry, 1990. 265(26): p. 15392-15402.

231. Chance B, O.N., Sugano T, Mayevsky A, Basic principles of tissue oxygen determination from mitochondrial signals. Advances in experimental medicine and biology, 1973. 37a: p. 277-292.

232. Richardson RS, N.E., Leigh JS, Wagner PD, Lactate efflux from exercising human skeletal muscle: role of intracellular PO2. Journal of applied physiology, 1998. 85(2): p. 627-634.

255 233. Richardson RS, N.E., Kendrick KF, Leigh JS, Wagner PD, Myoglobin O2 desaturation during exercise. Evidence of limited O2 transport. The Journal of clinical investigation, 1995. 96(4): p. J Clin Invest 1995 Oct;96(4):1916-26.

234. De Mey JG, V.P., Anoxia and endothelium-dependent reactivity of the canine femoral artery. Journal of physiology, 1983. 335: p. 65-74.

235. Rubanyi GM, V.P., Hypoxia releases a vasoconstrictor substance from the canine vascular endothelium. Journal of physiology, 1985. 364: p. 45-56.

236. Katusic ZS, V.P., Anoxic contractions in isolated canine cerebral arteries: contribution of endothelium-derived factors, metabolites of arachidonic acid, and calcium entry. Journal of cardiovascular pharmacology, 1986. 8(Suppl 8): p. S97-101.

237. Iqbal A, V.P., Flunarizine inhibits endothelium-dependent hypoxic facilitation in canine coronary arteries through an action on vascular smooth muscle. British journal of pharmacology, 1988. 95(3): p. 789-794.

238. Vanhoutte PM, A.-S.W., Boulanger C, Janssen PA, Katusic ZS, Komori K, Miller VM, Schini VB, Vidal M, Does endothelin-1 mediate endothelium-dependent contractions during anoxia? Journal of cardiovascular pharmacology, 1989. 13(Suppl 5): p. S124-142.

239. Vanhoutte PM, K.Z., Endothelium-derived contracting factor: endothelin and/or superoxide anion? Trends in pharmacological sciences, 1988. 9(7): p. 229-230.

240. Graser T, V.P., Hypoxic contraction of canine coronary arteries: role of endothelium and cGMP. American journal of physiology, 1991. 261(6 Pt 2): p. H1769-1777.

241. Muramatsu M, I.Y., Shimizu K, Asano H, Toki Y, Miyazaki Y, Okumura K, Hashimoto H, Ito T, Hypoxia-elicited contraction of aorta and coronary artery via removal of endothelium-derived nitric oxide. American journal of physiology, 1992. 263(5 Pt 2): p. H1339-1347.

242. Pearson PJ, L.P., Evora PR, Schaff HV, Endothelium-dependent response of human internal mammary artery to hypoxia. American journal of physiology, 1993. 264(2 Pt 2): p. H376-380.

243. Kobayashi K, N.J., Control of maximum rates of glycolysis in rat cardiac muscle. Circulation research, 1979. 44(2): p. 166-175.

244. Bunger R, P.B., Yaffe S, Energy utilization and pyruvate as determinants of pyruvate dehydrogenase in norepinephrine-stimulated heart. European journal of physiology, 1983. 397(3): p. 214-219.

245. Bunger R, P.B., Parallel stimulation by Ca2+ of inotropism and pyruvate dehydrogenase in perfused heart. American journal of physiology, 1984. 247(1 Pt 1): p. C45-52.

256 246. Case RB, G.H., The response of canine coronary vascular resistance to local alterations in coronary arterial PCO2. Circulation research, 1976. 39(4): p. 558-66.

247. Case RB, F.A., Wachter M, Measurement of myocardial PCO2 with a microelectrode: its relation to coronary sinus PCO2. American journal of physiology, 1979. 236(1): p. H29-34.

248. Murray PA, B.F., Sparks HV, The role of potassium in the metabolic control of coronary vascular resistance of the dog. Circulation research, 1979. 44(6): p. 767-780.

249. Bunger R, H.R., Querengasser A, Gerlach E, Studies on potassium induced coronary dilation in the isolated guinea pig heart. European journal of physiology, 1976. 363(1): p. 27-31.

250. Sparks HV Jr, B.F., The peripheral circulation: local regulation. Annual review of physiology, 1978. 40: p. 67-92.

251. Johansson B, J.O., Cell volume as a factor influencing electrical and mechanical activity of vascular smooth muscle. Acta physiologica Scandinavica, 1968. 72(4): p. 456- 468.

252. Haddy FJ, Ionic control of vascular smooth muscle cells. Kidney international. Supplement, 1988. 25: p. S2-8.

253. Shepherd JT, V.P., Local modulation of adrenergic neurotransmission in blood vessels. Journal of cardiovascular pharmacology, 1985. 7(Suppl 3): p. S167-178.

254. Verhaeghe RH, L.R., McGrath MA, Shepherd JT, Vanhoutte PM, Metabolic modulation of neurotransmitter release--adenosine, adenine nucleotides, potassium, hyperosmolarity, and hydrogen ion. Federation proceedings, 1978. 37(2): p. 208-211.

255. Muldoon SM, J.W., Verbeuren TJ, Vanhoutte PM, Alpha-adrenergic blocking properties of droperidol on isolated blood vessels of the dog. British journal of anaesthesia, 1977. 49(3): p. Br J Anaesth 1977 Mar;49(3):211-216.

256. Axelsson J, W.B., Johansson B, Jonsson O, Influence of the ionic environment on spontaneous electrical and mechanical activity of the rat portal vein. Circulation research, 1967. 21(5): p. 609-618.

257. Lorenz RR, V.P., Inhibition of adrenergic neurotransmission in isolated veins of the dog by potassium ions. Journal of physiology, 1975. 246(2): p. 479-500.

258. Muldoon SM, V.T., Vanhoutte PM, Effects of amide-linked local anesthetics on adrenergic neuroeffector junction in cutaneous veins of dog. The Journal of pharmacology and experimental therapeutics, 1976. 196(3): p. 723-736.

259. Olsson RA, D.C., Khouri EM, Patterson RE, Evidence for an adenosine receptor on the surface of dog coronary myocytes. Circulation research, 1976. 39(1): p. 93-8.

257 260. Schrader J, N.S., Gerlach E, Evidence for a cell surface adenosine receptor on coronary myocytes and atrial muscle cells. Studies with an adenosine derivative of high molecular weight. European journal of physiology, 1977. 369(3): p. 251-257.

261. Olsson RA, K.E., Bedynek JL Jr, McLean J, Coronary vasoactivity of adenosine in the conscious dog. Circulation research, 1979. 45(4): p. 468-478.

262. Juhran W, V.E., Dietmann K, Schaumann W, Pharmacological effects on coronary reactive hyperemia in conscious dogs. Naunyn-Schmiedebergs Archiv fur Pharmakologie, 1971. 269(1): p. 32-47.

263. Schwartz GG, M.P., Greenfield JC Jr, Hyperemic response of the coronary circulation to brief diastolic occlusion in the conscious dog. Circulation research, 1982. 50(1): p. 28-37.

264. Merrill GF, H.F., Dabney JM, Adenosine, theophylline, and perfusate pH in the isolated, perfused guinea pig heart. Circulation research, 1978. 42(2): p. 225-229.

265. Verhaeghe RH, Action of adenosine and adenine nucleotides on dogs' isolated veins. American journal of physiology, 1977. 233(1): p. H114-121.

266. Verhaeghe RH, V.P., Shepherd JT, Inhibition of sympathetic neurotransmission in canine blood vessels by adenosine and adenine nucleotides. Circulatory research, 1977. 40(2): p. 208-215.

267. Sabouni MH, B.G., Kotake AN, Douglas FL, Mustafa SJ, Effects of CGS-15943A on the relaxations produced by adenosine analogs in human blood vessels. European journal of pharmacology, 1990. 187(3): p. 525-530.

268. Ford GA, H.B., Vestal RE, Blaschke TF, Age-related changes in adenosine and beta-adrenoceptor responsiveness of vascular smooth muscle in man. British journal of clinical pharmacology, 1992. 33(1): p. 83-87.

269. Takahashi Y, I.N., Lue TF, Tanagho EA, Effects of adenosine on canine penile erection. The Journal of urology, 1992. 148(4): p. 1323-1325.

270. Mathie RT, A.B., Ralevic V, Burnstock G, Adenosine-induced dilatation of the rabbit hepatic arterial bed is mediated by A2-purinoceptors. British journal of pharmacology, 1991. 103(1): p. 1103-1107.

271. Wood BE, O.C.S., Leff P, Adenosine 5'-(2-fluorodiphosphate) is not a selective P2Y purinoceptor agonist in the rabbit jugular vein. The Journal of pharmacology and experimental therapeutics, 1992. 262(2): p. 566-569.

272. Olsson RA, B.R., Metabolic control of coronary blood flow. Progress in cardiovascular diseases, 1987. 29(5): p. 369-387.

258 273. Sun D, M.E., Kaley G, Koller A, Characteristics and origin of myogenic response in isolated mesenteric arterioles. American journal of physiology, 1992. 263(5 Pt 2): p. H1486-1491.

274. Sun D, K.G., Koller A, Characteristics and origin of myogenic response in isolated gracilis muscle arterioles. American journal of physiology, 1994. 266(3 Pt 2): p. H1177- 1183.

275. Jones AW, S.S., Geisbuhler BB, Stimulation of phospholipase D activity and phosphatidic acid production by norepinephrine in rat aorta. American journal of physiology, 1993. 246(3 Pt 1): p. C609-616.

276. Ostroumoff A, Experiments onthe inhibitory nerves of skin blood vessels. European journal of physiology, 1868. 12: p. 219-277.

277. Gaskell WH, On the tonicity of the heart and blood vessels. Journal of physiology, 1881. 3: p. 48-75.

278. Bayliss WM, On the local reactions of the arterial wall to changes of internal pressure. Journal of physiology, 1902. 28: p. 220-231.

279. Folkow B, Intravascular pressure as a factor regulating the tone of the small vessels. Acta physiologica Scandinavica, 1949. 17: p. 389-310.

280. Folkow B, A study of the factors influencing the tone of denervated blood vessels performed at various pressures. Acta physiologica Scandinavica, 1952. 27: p. 99-117.

281. Selkurt EE, J.P., Effect of acute elevation of portal venous pressure on mesenteric blood volume, interstitital fluid volume and hemodynamics. Circulatroy research, 1958. 6: p. 592-599.

282. Johnson PC, Autoregulatory responses of cat mesenteric arterioles measured in vivo. Circulatory research, 1968. 22(2): p. 199-212.

283. Wiederhielm CA, Effects of temperature and transmural pressure on contractile activity of vascular smooth muscle. Bibliotheca anatomica, 1967. 9: p. 321-327.

284. Sparks HV, Effect of quick stretch on isolated vascular smooth muscle. Circulatory research, 1964. 1(Suppl): p. I254-260.

285. Duling BR, G.R., Dacey RG Jr, Damon DN, Methods for isolation, cannulation, and in vitro study of single microvessels. American journal of physiology, 1981. 241(1): p. H108-116.

286. Harder DR, Pressure-dependent membrane depolarization in cat middle cerebral artery. Circulatory research, 1984. 55(2): p. 197-202.

259 287. Osol G, H.W., Myogenic properties of cerebral blood vessels from normotensive and hypertensive rats. American journal of physiology, 1985. 249(5 Pt 2): p. H914-921.

288. Jackson PA, D.B., Myogenic response and wall mechanics of arterioles. American journal of physiology, 1989. 257(4 Pt 2): p. H1147-1155.

289. Kuo L, D.M., Chilian WM, Myogenic activity in isolated subepicardial and subendocardial coronary arterioles. American journal of physiology, 1988. 255(6 Pt 2): p. H1558-1562.

290. Thorin-Trescases N, B.T., Hyman N, Penar PL, Walters CL, Bevan RD, Bevan JA, Diameter dependence of myogenic tone of human pial arteries. Possible relation to distensibility. Stroke, 1997. 28(12): p. 2486-2492.

291. Osol G, L.I., Cipolla M, Protein kinase C modulates basal myogenic tone in resistance arteries from the cerebral circulation. Circulatory research, 1991. 68(2): p. 359- 367.

292. Watanabe J, K.A., Horiguchi S, Keitoku M, Satoh S, Takishima T, Shirato K, Modification of myogenic intrinsic tone and [Ca2+]i of rat isolated arterioles by ryanodine and cyclopiazonic acid. Circulatory research, 1993. 73(3): p. 465-472.

293. Mellander S, Functional aspects of myogenic vascular control. Journal of hypertension. Supplement, 1989. 7(4): p. S21-S30.

294. Mellander S, M.M., Bjornberg J, Andersson LO., Autoregulation of capillary pressure and filtration in cat skeletal muscle in states of normal and reduced vascular tone. Acta physiologica Scandinavica, 1987. 129(3): p. 337-351.

295. Jarhult J, M.S., Autoregulation of capillary hydrostatic pressure in skeletal muscle during regional arterial hypo- and hypertension. Acta physiologica Scandinavica, 1974. 91(1): p. 32-41.

296. Osol G, Mechanotransduction by vascular smooth muscle. Journal of vascular research, 1995. 32(5): p. 275-292.

297. Nakayama K, Calcium-dependent contractile activation of cerebral artery produced by quick stretch. American journal of physiology, 1982. 242(5): p. H760-H768.

298. Meininger GA, D.M., Cellular mechanisms involved in the vascular myogenic response. American journal of physiology, 1992. 263(3 Pt 2): p. H647-H659.

299. Harder DR, Pressure-dependent membrane depolarization in cat middle cerebral artery. Circulation research, 1984. 55(2): p. 197-202.

300. Davis MJ, D.J., Hood JD, Stretch-activated single-channel and whole cell currents in vascular smooth muscle cells. American journal of physiology, 1992. 262(4 Pt 1): p. C1083-C1088.

260 301. Wellner MC, I.G., Properties of stretch-activated channels in myocytes from the guinea-pig urinary bladder. Journal of physiology, 1993. 466: p. 213-227.

302. Setoguchi M, O.Y., Abe I, Fujishima M, Stretch-activated whole-cell currents in smooth muscle cells from mesenteric resistance artery of guinea-pig. Journal of physiology, 1997. 501( Pt 2): p. 343-353.

303. Barclay JK, A delivery-independent blood flow effect on skeletal muscle fatigue. Journal of Applied Physiology, 1986. 61(3): p. 1084-1090.

304. Bevegard BS, S.J., Regulation of the circulation during exercise in man. Physiological review, 1967. 47(2): p. 178-213.

305. Segal SS, K.D., Coordination of blood flow control in the resistance vasculature of skeletal muscle. Medicine and science in sports and exercise, 1995. 27(8): p. 1158-1164.

306. Rowell LB, Muscle blood flow in humans: how high can it go? Medicine and science in sports and exercise, 1988. 5(5 Suppl): p. S97-103.

307. Saltin B, R.G., Koskolou MD, Roach RC, Skeletal muscle blood flow in humans and its regulation during exercise. Acta physiologica Scandinavica, 1998. 162(3): p. 421- 436.

308. Rowell LB, Neural control of muscle blood flow: importance during dynamic exercise. Clinical and experimental pharmacology & physiology, 1997. 24(2): p. 117- 125.

309. Freyschuss U, Cardiovascular adjustment to somatomotor activation. The elicitation of increments in heart rate, aortic pressure and venomotor tone with the initiation of muscle contraction. Acta physiologica Scandinavica. Supplementum, 1970. 342: p. 1-63.

310. Iellamo F, L.J., Raimondi G, Peruzzi G, Baroreflex control of sinus node during dynamic exercise in humans: effects of central command and muscle reflexes. American journal of physiology, 1997. 272(3 Pt 2): p. H1157-1164.

311. Hollander AP, B.L., Cardiac acceleration in man elicited by a muscle-heart reflex. Journal of applied physiology, 1975. 38(2): p. 272-278.

312. Friedman DB, P.C., Mitchell JH, Cardiovascular responses to voluntary and nonvoluntary static exercise in humans. Journal of applied physiology, 1992. 73(5): p. 1982-1985.

313. Seals DR, C.P., Taylor JA, Autonomic mediation of the pressor responses to isometric exercise in humans. Journal of applied physiology, 1988. 64(5): p. 2190-2196.

314. Seals DR, Influence of muscle mass on sympathetic neural activation during isometric exercise. Journal of applied physiology, 1989. 67(5): p. 1801-1806.

261 315. Mark AL, V.R., Nerhed C, Wallin BG, Microneurographic studies of the mechanisms of sympathetic nerve responses to static exercise in humans. Circulation research, 1985. 57(3): p. 461-469.

316. Turner DL, Cardiovascular and respiratory control mechanisms during exercise: an integrated view. The Journal of experimental biology, 1991. 160: p. 309-340.

317. Winchester PK, W.J., Mitchell JH, Cardiovascular responses to static exercise in patients with Brown-Sequard syndrome. Journal of physiology, 2000. 527( Pt 1): p. 193- 202.

318. Mitchell JH, Neural control of the circulation during exercise. Medicine and science in sports and exercise, 1990. 22(2): p. 141-154.

319. Smith EE, G.A., Manning RD, White RJ, Integrated mechanisms of cardiovascular response and control during exercise in the normal human. Progress in cardiovascular diseases, 1976. 18(6): p. 421-444.

320. Folkow B, S.R., Wright DL, Loci of neurogenic and metabolic effects on precapillary vessels of skeletal muscle. Acta physiologica Scandinavica, 1971. 81(4): p. 459-471.

321. Thompson LP, M.D., Blood flow and oxygen consumption in skeletal muscle during sympathetic stimulation. American journal of physiology, 1983. 245(1): p. H66-71.

322. Clausen JP, Effect of physical training on cardiovascular adjustments to exercise in man. Physiological reviews, 1977. 57(4): p. 779-815.

323. O'Leary DS, R.L., Scher AM, Baroreflex-induced vasoconstriction in active skeletal muscle of conscious dogs. American journal of physiology, 1991. 260(1 Pt 2): p. H37-41.

324. Kjellimer I, On the competition between metabolic vasodilatation and neurogenic vasoconstriction in skeletal muscle. Acta physiologica Scandinavica, 1965. 63: p. 450- 459.

325. Vanhoutte PM, V.T., Webb RC, Local modulation of adrenergic neuroeffector interaction in the blood vessel well. Physiological review, 1981. 16(1): p. 151-247.

326. Cohen RA, W.R., Endothelium inhibits norepinephrine release from adrenergic nerves of rabbit carotid artery. American journal of physiology, 1988. 254(5 Pt 2): p. H871-878.

327. Anderson KM, F.J., Differential sensitivity of arteriolar alpha 1- and alpha 2- adrenoceptor constriction to metabolic inhibition during rat skeletal muscle contraction. Circulation research, 1991. 69(1): p. 174-184.

262 328. Paulsen W, B.D., Ko P, Cunningham DA, Persaud JA, Left ventricular function in marathon runners: echocardiographic assessment. Journal of applied physiology, 1981. 51(4): p. 881-886.

329. Miles DS, C.M., Bomze JP, Cardiovascular responses to upper body exercise in normals and cardiac patients. Medicine and science in sports and exercise, 1989. 21(5 Suppl): p. S126-131.

330. Pendergast DR, Cardiovascular, respiratory, and metabolic responses to upper body exercise. Medicine and science in sports and exercise, 1989. 21(5 Suppl): p. S121-125.

331. Delp MD, L.M., Regulation of skeletal muscle perfusion during exercise. Acta physiologica Scandinavica, 1998. 162(3): p. 411-419.

332. Mohrman DE, R.R., Relation of blood flow to VO2, PO2, and PCO2 in dog gastrocnemius muscle. American journal of physiology, 1988. 255(5 Pt 2): p. H1004- 1010.

333. Hilton SM, H.O., Marshall JM, Possible mediators of functional hyperaemia in skeletal muscle. Journal of physiology, 1978. 282: p. 131-147.

334. Mackie BG, T.R., Blood flow to different skeletal muscle fiber types during contraction. American journal of physiology, 1983. 245(2): p. H265-275.

335. Armstrong RB, L.M., Rat muscle blood flows during high-speed locomotion. Journal of applied physiology, 1985. 59(4): p. 1322-1328.

336. Laughlin MH, A.R., Muscular blood flow distribution patterns as a function of running speed in rats. American journal of physiology, 1982. 243(2): p. H296-306.

337. Johnson JM, R.L., Forearm skin and muscle vascular responses to prolonged leg exercise in man. Journal of applied physiology, 1975. 39(6): p. 920-924.

338. Laughlin MH, A.R., Muscle blood flow during locomotory exercise. Exercise and sport sciences reviews, 1985. 13: p. 95-136.

339. Armstrong RB, D.M., Goljan EF, Laughlin MH, Distribution of blood flow in muscles of miniature swine during exercise. Journal of applied physiology, 1987. 62(3): p. 1285-1298.

340. Franke WD, S.G., Schmid PG 3rd, Effects of intense exercise training on endothelium-dependent exercise-induced vasodilatation. Clinical Physiology, 1998. 18(6): p. 521-528.

341. Skinner NS Jr, C.J., Interactions between oxygen, potassium, and osmolality in regulation of skeletal muscle blood flow. Circulation research, 1971. 28(Suppl 1): p. Circ Res 1971 Jan;28:Suppl 1:73-85.

263 342. Burcher E, G.D., Effects of exercise metabolites on adrenergic vasoconstriction in the gracilis muscle of the dog. The Journal of pharmacology and experimental therapeutics, 1975. 192(1): p. 149-156.

343. Dodd LR, J.P., Antagonism of vasoconstriction by muscle contraction differs with alpha-adrenergic subtype. Amercian journal of physiology, 1993. 264(3 Pt 2): p. H892- 900.

344. Anrep GV, v.S.E., The blood flow through skeletal muscle in relation to its contraction. Journal of physiology, 1935. 85: p. 375-399.

345. Radawski DP, H.W., Haddy FJ, Role of vasoactive substances in active hyperemia in skeletal muscle. Proceedings of the Society for Experimental Biology and Medicine, 1975. 148(1): p. 270-276.

346. Hellsten Y, M.D., Radegran G, Saltin B, Bangsbo J., Adenosine concentrations in the interstitium of resting and contracting human skeletal muscle. Circulation, 1998. 98(1): p. 6-8.

347. Hellsten Y, F.U., Adenosine formation in contracting primary rat skeletal muscle cells and endothelial cells in culture. Journal of physiology, 1997. 504(Pt 3): p. 695-704.

348. Deussen A, M.G., Schrader J, Contribution of coronary endothelial cells to cardiac adenosine production. Pflugers Archiv, 1986. 406(6): p. 608-614.

349. Nees S, H.V., Becker BF, Bock M, Des Rosiers Ch, Gerlach E, The coronary endothelium: a highly active metabolic barrier for adenosine. Basic research in cardiology, 1985. 80(5): p. 515-529.

350. Cunha RA, S.A., Adenosine and adenine nucleotides are independently released from both the nerve terminals and the muscle fibres upon electrical stimulation of the innervated skeletal muscle of the frog. Pflugers Archiv, 1993. 424(5-6): p. 503-510.

351. Radegran G, S.B., Muscle blood flow at onset of dynamic exercise in humans. American Jjournal of Physiology, 1998. 274(1 Pt 2): p. H314-322.

352. Sollevi A, Cardiovascular effects of adenosine in man; possible clinical implications. Progress in neurobiology, 1986. 27(4): p. 319-349.

353. Radegran G, S.B., Nitric oxide in the regulation of vasomotor tone in human skeletal muscle. American journal of physiology, 1999. 276(6 Pt 2): p. H1951-1960.

354. Poucher SM, The role of the A(2A) adenosine receptor subtype in functional hyperaemia in the hindlimb of anaesthetized cats. Journal of applied physiology, 1996. 492(Pt 2): p. 495-503.

264 355. Tabaie HM, S.J., Haddy FJ, Reduction of exercise dilation by theophylline. Proceedings of the Society for Experimental Biology and Medicine, 1977. 154(1): p. 93- 97.

356. Persson MG, O.A., Lindbom L, Hedqvist P, Gustafsson LE, Role of adenosine in functional hyperemia in skeletal muscle as indicated by pharmacological tools. Naunyn- Schmiedeberg's archives of pharmacology, 1991. 343(1): p. 52-57.

357. Proctor KG, Reduction of contraction-induced arteriolar vasodilation by adenosine deaminase or theophylline. American journal of physiology, 1984. 247(2 Pt 2): p. H195- 205.

358. Radegran G, H.Y., Adenosine and nitric oxide in exercise-induced human skeletal muscle vasodilatation. Acta physiologica Scandinavica, 2000. 168(4): p. 575-591.

359. Schwartz LM, M.J., Adenosine and active hyperemia in soleus and gracilis muscle of cats. American journal of physiology, 1990. 259(4 Pt 2): p. H1295-1304.

360. Tominaga S, S.T., Nakamura T, Evaluation of roles of potassium, inorganic phosphate, osmolarity, pH, pCO2, pO2, and adenosine or AMP in exercise and reactive hyperemias in canine hindlimb muscles. The Tohoku journal of experimental medicine, 1973. 109(4): p. 347-363.

361. Stowe DF, O.T., Anderson DK, Haddy FJ, Scott JB, Interaction of O2 and CO2 in sustained exercise hyperemia of canine skeletal muscle. American journal of physiology, 1975. 229(1): p. 28-33.

362. Mohrman DE, Lack of influence of potassium or osmolality on steady-state exercise hyperemia. American journal of physiology, 1982. 242(6): p. H949-954.

363. Wilson JR, K.S., Krishna GG, Contribution of potassium to exercise-induced vasodilation in humans. Journal of applied physiology, 1994. 77(6): p. 2552-2557.

364. Lash JM, B.H., Perivascular and tissue PO2 in contracting rat spinotrapezius muscle. Amerrican journal of physiology, 1987. 252(6 Pt 2): p. H1192-1202.

365. Scott JB, R.M., Radawski D, Haddy FJ., Role of osmolarity, K+, H+, Mg++, and O2 in local blood flow regulation. American journal of physiology, 1970. 218(2): p. 338-345.

366. Gray SD, Effect of hypertonicity on vascular dimensions in skeletal muscle. Microvascular research, 1971. 3(1): p. 117-124.

367. Scott JB, R.D., Role of hyperosmolarity in the genesis of active and reactive hyperemia. Circulation research, 1971. 28(Suppl 1): p. 26-32.

368. Stainsby WN, B.J., Effect of infusions of osmotically active substances on muscle blood flow and systemic blood pressure. Circulation research, 1971. 28(Suppl 1): p. 33- 38.

265 369. Lash JM, Contribution of arterial feed vessels to skeletal muscle functional hyperemia. Journal of applied physiology, 1994. 76(4): p. 1512-1519.

370. Supinski GS, B.H., Guanciale A, Kelsen SG, Effect of alterations in muscle fiber length on diaphragm blood flow. Journal of applied physiology, 1986. 60(5): p. 1789- 1796.

371. Hobbs SF, M.D., Effects of blood pressure on force production in cat and human muscle. Journal of applied physiology, 1987. 63(2): p. 834-839.

372. Gray SD, S.N., Resistance to blood flow in leg muscles of dog during tetanic isometric contraction. American journal of physiology, 1967. 213(3): p. 677-682.

373. Gray SD, C.E., Staub NC, Site of increased vascular resistance during isometric muscle contraction. American journal of physiology, 1967. 213(3): p. 683-689.

374. Mohrman DE, S.H., Myogenic hyperemia following brief tetanus of canine skeletal muscle. American journal of physiology, 1974. 227(3): p. 531-535.

375. Meininger GA, M.C., Fehr KL, Bohlen HG, Myogenic vasoregulation overrides local metabolic control in resting rat skeletal muscle. Circulation research, 1987. 60(6): p. 861-870.

376. Lash JM, S.A., Pressure dependence of baroreceptors-mediated vasoconstriction in rat skeletal muscle. Journal of applied physiology, 1991. 70(6): p. 2551-2558.

377. Rowell LB, O.L.D., Reflex control of the circulation during exercise: chemoreflexes and mechanoreflexes. Journal of applied physiology, 1990. 69(2): p. 407-418.

378. Seals DR, V.R., Regulation of muscle sympathetic nerve activity during exercise in humans. Exercise and sport sciences reviews, 1991. 19: p. 313-349.

379. Pohl U, D.W.C., Gloe T, Large arterioles in the control of blood flow: role of endothelium-dependent dilation. Acta physiologica Scandinavica, 2000. 168(4): p. 505- 510.

380. Hartley LH, M.J., Hogan RP, Jones LG, Kotchen TA, Mougey EH, Wherry FE, Pennington LL, Ricketts PT, Multiple hormonal responses to graded exercise in relation to physical training. Journal of applied physiology, 1972. 33(5): p. 602-606.

381. Hickner RC, F.J., Ehsani AA, Kohrt WM, Role of nitric oxide in skeletal muscle blood flow at rest and during dynamic exercise in humans. American Journal of Physiology, 1997. 273(1 Pt 2): p. H405-410.

382. Vanhoutte PM, M.V., Alpha 2-adrenoceptors and endothelium-derived relaxing factor. The American journal of medicine, 1989. 87(3C): p. 1S-5S.

266 383. Miller VM, V.P., Endothelial alpha 2-adrenoceptors in canine pulmonary and systemic blood vessels. European journal of pharmacology, 1985. 118(1-2): p. 123-129.

384. Niebauer J, C.J., Cardiovascular effects of exercise: role of endothelial shear stress. Journal of American College of Cardiology, 1996. 28(7): p. 1652-1660.

385. Koller A, K.G., Endothelial regulation of wall shear stress and blood flow in skeletal muscle microcirculation. American journal of physiology, 1991. 260(3 Pt 2): p. H862- 868.

386. Rowell LB, S.B., Kiens B, Christensen NJ, Is peak quadriceps blood flow in humans even higher during exercise with hypoxemia? American journal of physiology, 1986. 251(5 Pt 2): p. H1038-1044.

387. Segal SS, B.J., Intracellular recording and dye transfer in arterioles during blood flow control. American journal of physiology, 1992. 263(1 Pt 2): p. H1-7.

388. Beny JL, C.J., An electron-microscopic study of smooth muscle cell dye coupling in the pig coronary arteries. Role of gap junctions. Circulation research, 1992. 70(1): p. 49- 55.

389. Pierzga JM, S.S., Spatial relationships between neuromuscular junction and microvessel in hamster cremaster muscle. Microvascular research, 1994. 48: p. 50-67.

390.Segal SS, Communication among endothelial and smooth muscle cells coordinates blood flow control during exercise. News of physiological science, 1992. 7: p. 152-156.

391. Siato Y, E.A., Lockard V, Hester RL, Role of the venular endothelium in control of arteriolar diameter during functional hyperemia. American journal of physiology, 1994. 267: p. H1227-1231.

392. Maeda S, M.T., Goto K, Matsuda M., Alteration of plasma endothelin-1 by exercise at intensities lower and higher than ventilatory threshold. Journal of applied physiology, 1994. 77(3): p. 1399-1402.

393. Maeda S, M.T., Sakane M, Saito M, Maki S, Goto K, Matsuda M., Does endothelin- 1 participate in the exercise-induced changes of blood flow distribution of muscles in humans? Journal of applied physiology, 1997. 82(4): p. 1107-1111.

394. Bevegard BS, S.J., Reaction in man of resistance and capacity vessels in forearm and hand to leg exercise. Journal of applied physiology, 1966. 21(1): p. 123-132.

395. Bevegard BS, S.J., Changes in tone of limb veins during supine exercise. Journal of applied physiology, 1965. 20(1): p. 1-8.

396. Bevegard BS, S.J., Effect of local exercise of forearm muscles on forearm capacitance vessels. Journal of applied physiology, 1965. 20(5): p. 968-974.

267 397. Seaman RG, W.R., Zechman FW, Goldey JA., Venous reactivity during static exercise (handgrip) in man. Journal of applied physiology, 1973. 35(6): p. 858-860.

398. Robinson BF, W.A., Effect on forearm arteries and veins of attenuation of the cardiac response to leg exercise. Clinical science, 1968. 35(1): p. 143-152.

399. Detry JM, W.C., Rowell LB, Increased forearm venous compliance during exercise with nitroglycerin. Acta cardiologica, 1974. 29(1): p. 31-43.

400. Katz ML, C.A., Kerr RP, Caputo GC, Variability of venous-hemodynamics with daily activity. Journal of vascular surgery, 1994. 19(2): p. 361-365.

401. Greenfield ADM, P.G., On the capacity and distensibility of the blood vessels of the human forearm. The Journal of physiology, 1956. 131: p. 290.

402. Glover WE, G.A., Kidd BSL, Whelan RF, Reactions of the capacity blood vessels of the human hand and forearm to vaso-active substances infused intra-arterially. The Journal of physiology, 1958. 140: p. 113.

403. Coles DR, K.B., Moffat W, Distensibility of blood vessels of the human calf determined by local application of subatmospheric pressures. Journal of applied physiology, 1957. 10: p. 471.

404. Marshall JM, T.H., Direct observations of muscle arterioles and venules following contraction of skeletal muscle fibres in the rat. The Journal of physiology, 1984. 350: p. 447-459.

405. Glick MR, G.J., Gascho JA, Endothelium-derived nitric oxide reduces baseline venous tone in awake instrumented rats. American Journal of Physiology, 1993. 265(1 Pt 2): p. H47-51.

406. Saetre T, S.O., Scholz T, Carlsen H, Nordsletten L, Fahlstrom E, Aasen AO, Nitric oxide synthase inhibition reduces venous return in porcine endotoxemia. American Journal of Physiology 1996 Oct;271(4 Pt 2):H1325-32, 1996. 271(4 Pt 2): p. H1325- 1332.

407. Koller A, D.G., Kaley G., Flow-induced responses in skeletal muscle venules: modulation by nitric oxide and prostaglandins. American journal of physiology, 1998. 275(3 Pt 2): p. H831-836.

408. Shepherd JT, Behavior of resistance and capacity vessels in human limbs during exercise. Circulation research., 1967. 20(Suppl 1): p. I70-82.

409. Saltin B, Physiological adaptation to physical conditioning. Old problems revisited. Acta medica Scandinavica. Supplementum, 1986. 711: p. 11-24.

410. Grimby G, S.B., Physiological effects of physical training. Scandinavian journal of rehabilitation medicine, 1971. 3(1): p. 6-14.

268 411. Clausen JP, Circulatory adjustments to dynamic exercise and effect of physical training in normal subjects and in patients with coronary artery disease. Progress in cardiovascular diseases, 1976. 18(6): p. 459-495.

412. Saltin B, B.G., Mitchell JH, Johnson RL Jr, Wildenthal K, Chapman CB, Response to exercise after bed rest and after training. Circulation, 1968. 38(5 Suppl): p. VII1-78.

413. Rowell LB, Human cardiovascular adjustments to exercise and thermal stress. Physiological review, 1974. 54(1): p. 75-159.

414. Delp MD, Differential effects of training on the control of skeletal muscle perfusion. Medicine and Science in Sports and Exercise, 1998. 30(3): p. 361-374.

415. Andersen P, H.J., Capillary supply of the quadriceps femoris muscle of man: adaptive response to exercise. Journal of applied physiology, 1977. 270(3): p. 677-690.

416. Clausen JP, K.K., Rasmussen B, Trap-Jensen J, Central and peripheral circulatory changes after training of the arms or legs. American journal of physiology, 1973. 225(3): p. 675-682.

417. Ekblom B, K.A., Soltysiak J, Physical training, bradycardia, and autonomic nervous system. Scandinavian journal of clinical and laboratory investigation, 1973. 32(3): p. 251-256.

418. Tipton CM, Training and bradycardia in rats. American journal of physiology, 1965. 209(6): p. 1089-1094.

419. Winder WW, H.J., Hickson RC, Ehsani AA, McLane JA., Time course of sympathoadrenal adaptation to endurance exercise training in man. Journal of applied physiology, 1978. 45(3): p. 370-374.

420. Pelliccia A, M.B., Spataro A, Proschan MA, Spirito P, The upper limit of physiologic cardiac hypertrophy in highly trained elite athletes. The New England journal of medicine, 1991. 324(5): p. 295-301.

421. Cox ML, B.J.r., Dudley GA, Exercise training-induced alterations of cardiac morphology. Journal of applied physiology, 1986. 61(3): p. 926-931.

422. Ehsani AA, H.J., Hickson RC, Rapid changes in left ventricular dimensions and mass in response to physical conditioning and deconditioning. The American journal of cardiology, 1978. 42(1): p. 52-56.

423. Gillen CM, L.R., Mack GW, Tomaselli CM, Nishiyasu T, Nadel ER, Plasma volume expansion in humans after a single intense exercise protocol. Journal of applied physiology, 1991. 71(5): p. 1914-1920.

424. Hopper MK, C.A., Coyle EF, Exercise stroke volume relative to plasma-volume expansion. Journal of applied physiology, 1988. 64(1): p. 404-408.

269 425. Coyle EF, H.M., Coggan AR, Effects of detraining on cardiovascular responses to exercise: role of blood volume. Journal of applied physiology, 1986. 60(1): p. 95-99.

426. Saltin B, N.K., Costill DL, Stein E, Jansson E, Essen B, Gollnick D, The nature of the training response; peripheral and central adaptations of one-legged exercise. Acta physiologica Scandinavica, 1976. 96(3): p. 289-305.

427. Clausen JP, T.-J.J., Lassen NA, The effects of training on the heart rate during arm and leg exercise. Scandinavian journal of clinical and laboratory investigation, 1970. 26(3): p. 295-301.

428. Sinoway LI, M.T., Minotti JR, Zelis R., Enhanced maximal metabolic vasodilation in the dominant forearms of the tennis players. Journal of Applied Physiology, 1986. 61(2): p. 673-678.

429. Sinoway LI, Effect of conditioning and deconditioning stimuli on metabolically determined blood flow in humans and implications for congestive heart failure. American Journal of Cardiology, 1988. 62(8): p. 45E-48E.

430. Green DJ, C.N., Fox C, Rankin JM, Taylor RR, Modification of forearm resistance vessels by exercise training in young men. Journal of applied physiology, 1994. 77(4): p. 1829-1833.

431. Armstrong RB, H.D., Delp MD, Blood flow distribution in rat muscles during preexercise anticipatory response. Journal of applied physiology, 1989. 67(5): p. 1855- 1861.

432. Huonker M, H.M., Keul J, Structural and functional adaptations of the cardiovascular system by training. International journal of sports medicine, 1996. 17(Suppl 3): p. S164-172.

433. Saito M, M.H., Miyamura M, Effects of physical training on the calf and thigh blood flows. Japanese journal of physiology, 1980. 30(6): p. 955-959.

434. Bond V Jr, W.P., Adams RG, Johnson AT, Vaccaro P, Tearney RJ, Millis RM, Franks BD, Bassett DR Jr., Lower leg high-intensity resistance training and peripheral hemodynamic adaptations. Canadian journal of applied physiology, 1996. 21(3): p. 209- 117.

435. Proctor DN, M.J., Dietz NM, Minson CT, Joyner MJ, Reduced submaximal leg blood flow after high-intensity aerobic training. Journal of applied physiology, 2001. 91(6): p. 2619-2627.

436. Klausen K, S.N., Clausen JP, Hartling O, Trap-Jensen J, Central and regional circulatory adaptations to one-leg training. Journal of applied physiolgoy, 1982. 52(4): p. 976-983.

270 437. Musch TI, H.G., Ordway GA, Longhurst JC, Mitchell JH, Training effects on regional blood flow response to maximal exercise in foxhounds. Journal of applied physiology, 1987. 62(4): p. 1724-1732.

438. Musch TI, T.J., Hilty MR, Effects of high-intensity sprint training on skeletal muscle blood flow in rats. Journal of applied physiology, 1991. 71(4): p. 1387-1395.

439. Armstrong RB, L.M., Exercise blood flow patterns within and among rat muscles after training. American journal of physiology, 1984. 246(1 Pt 2): p. H59-68.

440. Shoemaker JK, P.S., Green HJ, Hughson RL., Faster femoral artery blood velocity kinetics at the onset of exercise following short-term training. Cardiovascular research, 1996. 31(2): p. 278-286.

441. Yasuda Y, M.M., Cross transfer effects of muscular training on blood flow in the ipsilateral and contralateral forearms. European journal of applied physiology and occupational physiology, 1983. 51(3): p. 321-329.

442. Sinoway LI, S.J., Wilson J, McLaughlin D, Musch T, Zelis R, A 30-day forearm work protocol increases maximal forearm blood flow. Journal of applied physiology, 1987. 62(3): p. 1063-1067.

443. Alomari MA, W.M., Prisby RD, Lee CM, Wood RH,, Modification of forearm vascular function following short-term handgrip exercise training. International journal of sports medicine, 2001. 22(5): p. 361-365.

444. Silber D, M.D., Sinoway L, Leg exercise conditioning increases peak forearm blood flow. Journal of applied physiology, 1991. 71(4): p. 1568-1573.

445. Hashimoto M, Effects of exercise on plasma lipoprotein levels and endothelium- dependent vasodilatation in young and old rats. European journal of applied physiology and occupational physiology, 1990. 61(5-6): p. 440-445.

446. Delp MD, M.R., Laughlin MH, Exercise training alters endothelium-dependent vasoreactivity of rat abdominal aorta. Journal of applied physiology, 1993. 75(3): p. 1354-1363.

447. Delp MD, M.R., Laughlin MH, Exercise training alters aortic vascular reactivity in hypothyroid rats. American journal of physiology, 1995. 268(4 Pt 2): p. H1428-1435.

448. Chen HI, L.H., Physical conditioning can modulate endothelium-dependent vasorelaxation in rabbits. Arteriosclerosis and thrombosis, 1993. 13(6): p. 852-856.

449. Sun D, H.A., Koller A, Kaley G, Short-term daily exercise activity enhances endothelial NO synthesis in skeletal muscle arterioles of rats. Journal of applied physiology, 1994. 7(5): p. 2241-2247.

271 450. Delp MD, L.M., Time course of enhanced endothelium-mediated dilation in aorta of trained rats. Medicine and science in sports and exercise, 1997. 29(11): p. 1454-1461.

451. Green DJ, F.D., O'Driscoll JG, Blanksby BA, Taylor RR, Endothelium-derived nitric oxide activity in forearm vessels of tennis players. Journal of applied physiology, 1996. 81(2): p. 943-948.

452. Green DJ, O.D.J., Blanksby BA, Taylor RR, Effect of casting on forearm resistance vessels in young men. Medicine and science in sports and exercise, 1997. 29(10): p. 1325-1331.

453. Kingwell BA, T.B., Cameron JD, Jennings GL, Dart AM, Enhanced vasodilation to acetylcholine in athletes is associated with lower plasma cholesterol. American journal of physiology, 1996. 270(6 Pt 2): p. H2008-2013.

454. Kingwell BA, S.B., Jennings GL, Dart AM, Four weeks of cycle training increases basal production of nitric oxide from the forearm. American Journal of Physiology, 1997. 272((3 Pt 2)): p. H1070-1077.

455. Lewis TV, D.A., Chin-Dusting JP, Kingwell BA., Exercise training increases basal nitric oxide production from the forearm in hypercholesterolemic patients. Arteriosclerosis, thrombosis, and vascular biology, 1999. 19(11): p. 2782-2787.

456. Clarkson P, M.H., Mullen MJ, Donald AE, Powe AJ, Bull T, Jubb M, World M, Deanfield JE, Exercise training enhances endothelial function in young men. Journal of the American College of Cardiology, 1999. 33(5): p. 1379-1385.

457. Taddei S, V.A., Ghiadoni L, Salvetti G, Salvetti A, Endothelial dysfunction in hypertension. Journal of nephrology, 2000. 13(3): p. 205-210.

458. Hambrecht R, F.E., Weigl C, Gielen S, Hamann C, Kaiser R, Yu J, Adams V, Niebauer J, Schuler G, Regular physical exercise corrects endothelial dysfunction and improves exercise capacity in patients with chronic heart failure. Circulation, 1998. 98(24): p. 2709-7915.

459. Higashi Y, S.S., Kurisu S, Yoshimizu A, Sasaki N, Matsuura H, Kajiyama G, Oshima T, Regular aerobic exercise augments endothelium-dependent vascular relaxation in normotensive as well as hypertensive subjects: role of endothelium-derived nitric oxide. Circulation, 1999. 100(11): p. 1194-1202.

460. Higashi Y, S.S., Sasaki N, Nakagawa K, Ueda T, Yoshimizu A, Kurisu S, Matsuura H, Kajiyama G, Oshima T, Daily aerobic exercise improves reactive hyperemia in patients with essential hypertension. Hypertension, 1999. 33(1 Pt 2): p. 591-597.

461. Engelke KA, H.J., Proctor DN, Dietz NM, Joyner MJ, Contribution of nitric oxide and prostaglandins to reactive hyperemia in human forearm. Journal of applied physiology, 1996. 81(4): p. 1807-1814.

272 462. Bank AJ, S.R., Mullen K, Chuang PP, Effects of short-term forearm exercise training on resistance vessel endothelial function in normal subjects and patients with heart failure. Journal of Cardiac Failure, 1998. 4(3): p. 193-201.

463. Wiegman DL, H.P., Joshua IG, Miller FN, Decreased vascular sensitivity to norepinephrine following exercise training. Journal of applied physiology, 1981. 51(2): p. 282-287.

464. Tanabe K, Y.A., Suzuki N, Yokoyama Y, Osada N, Nakayama M, Akashi Y, Seki A, Samejima H, Oya M, Murabayashi T, Omiya K, Itoh H, Miyake F, Murayama M, Physiological role of endothelin-1 in nonworking muscles during exercise in healthy subjects. Japanese circulation journal, 2000. 64(1): p. 64(1):27-31.

465. Jones AW, R.L., Magliola L, Endothelin-1 sensitivity of porcine coronary arteries is reduced by exercise training and is gender dependent. Journal of applied physiology, 1999. 87(3): p. 1172-1177.

466. Laughlin MH, S.W., McAllister RM, Garverick HA, Jones AW, Interaction of gender and exercise training: vasomotor reactivity of porcine skeletal muscle arteries. Journal of applied physiology, 2001. 90(1): p. 216-227.

467. Maeda S, M.T., Kakiyama T, Sugawara J, Iemitsu M, Irukayama-Tomobe Y, Murakami H, Kumagai Y, Kuno S, Matsuda M,, Effects of exercise training of 8 weeks and detraining on plasma levels of endothelium-derived factors, endothelin-1 and nitric oxide, in healthy young humans. Life sciences, 2001. 69(9): p. 1005-1016.

468. Suzuki J, G.M., Batra S, Koyama T, Effects of treadmill training on the arteriolar and venular portions of capillary in soleus muscle of young and middle-aged rats. Acta physiologica Scandinavica, 1997. 159(2): p. 113-121.

469. Suzuki J, K.T., Uruma T, Koyama T, Time-course changes in arteriolar and venular portions of capillary in young treadmill-trained rats. Acta physiologica Scandinavica, 2001. 171(1): p. 77-86.

470. Kenney WL, A.C., The effect of aerobic conditioning on venous pooling in the foot. Medicine and science in sports and exercise, 1987. 19(5): p. 474-479.

471. Convertino VA, M.G., Nadel ER, Elevated central venous pressure: a consequence of exercise training-induced hypervolemia? American journal of physiology, 1991. 260(2 Pt 2): p. R273-277.

472. Louisy F, J.J., Guezennec CY, Filling and emptying characteristics of lower limb venous network in athletes. Study by postural plethysmography. International journal of sports medicine, 1997. 18(1).

473. Wecht JM, d.M.R., Weir JP, Bauman WA, Grimm DR, Effects of autonomic disruption and inactivity on venous vascular function. American journal of physiology., 2000. 278(2): p. H515-520.

273 474. Miyachi M, T.H., Yamamoto K, Yoshioka A, Takahashi K, Onodera S, Effects of one-legged endurance training on femoral arterial and venous size in healthy humans. Journal of applied physiology, 2001. 90(6): p. 2439-2444.

475. Ikenouchi H, I.M., Sato H, Momomura S, Serizawa T, Sugimoto T, Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Japanese heart journal, 1991. 32(1): p. 17-34.

476. Welsch MA, A.M., Parish TP, Wood RH, Kalb D, Influence of venous function on exercise tolerance in chronic heart failure. Journal of cardiopulmonary rehabilitation, 2002.

477. Haddy FJ, S.J., Metabolic factors in peripheral circulatory regulation. Federation proceedings, 1975. 34(11): p. 2001-11.

478. Levine TB, F.G., Goldsmith SR, Simon AB, Cohn JN, Activity of the sympathetic nervous system and renin-angiotensin system assessed by plasma hormone levels and their relation to hemodynamic abnormalities in congestive heart failure. The American journal of cardiology, 1982. 49(7): p. 1659-66.

479. Ardaillou R, C.D., Glomerular effects of angiotensin II: a reappraisal based on studies with non-peptide receptor antagonists. Journal of hypertension. Supplement, 1993. 11(3): p. S43-47.

480. Kubo SH, The role of neurohormonal activation in the pathophysiology and prognosis of patients with congestive heart failure. Coronary artery disease, 1993. 4(1): p. 4-15.

481. Sica DA, Pharmacotherapy in congestive heart failure: angiotensin II and thirst: therapeutic considerations. Congestive heart failure, 2001. 7(6): p. 325-328.

482. Pratt NG, Neurohormonal response to ventricular failure: pharmacologic management. The Journal of cardiovascular nursing., 1993. 8(1): p. 49-62.

483. Francis GS, Neuroendocrine activity in congestive heart failure. The American journal of cardiology., 1990. 66(11): p. 33D-39D.

484. Goldsmith SR, Baroreflex loading maneuvers do not suppress increased plasma arginine vasopressin in patients with congestive heart failure. Journal of the American College of Cardiology., 1992. 19(6): p. 1180-1184.

485. Benedict CR, J.D., Weiner DH, Bourassa MG, Bittner V, Kay R, Kirlin P, Greenberg B, Kohn RM, Nicklas JM, et al, Relation of neurohumoral activation to clinical variables and degree of ventricular dysfunction: a report from the Registry of Studies of Left Ventricular Dysfunction. SOLVD Investigators. Journal of the American College of Cardiology, 1994. 23(6): p. 1410-1420.

274 486. Creager MA, F.D., Cutler SS, Kohlmann O, Ryan TJ, Gavras H, Contribution of vasopressin to vasoconstriction in patients with congestive heart failure: comparison with the renin-angiotensin system and the sympathetic nervous system. Journal of the American College of Cardiology, 1986. 7(4): p. 758-765.

487. Ohnishi A, O.Y., Okahara R, Fujihara H, Inoue T, Yamamura Y, Yabuuchi Y, Tanaka T., Potent aquaretic agent. A novel nonpeptide selective vasopressin 2 antagonist (OPC-31260) in men. The Journal of clinical investigation, 1993. 92(6): p. 2653-2639.

488. Shimizu K, Aquaretic effects of the nonpeptide V2 antagonist OPC-31260 in hydropenic humans. Kidney international., 1995. 48(1): p. 220-226.

489. Fyhrquist F, T.I., Atrial natriuretic peptide in congestive heart failure. The American journal of cardiology, 1988. 62(2): p. 20A-24A.

490. Flores JS, Sympahoinhibitory effects of atrial natriuretic factor in normal humans. Circulation, 1990. 80: p. 1860-1873.

491. Middlekauff HR, Mechanisms and implications of autonomic nervous system dysfunction in heart failure. Current opinion in cardiology, 1997. 12(3): p. 265-275.

492. Opie LH, Compensation and overcompensation in congestive heart failure. American heart journal., 1990. 120(6 Pt 2): p. 1552-1557.

493. Schrier RW, A.W., Hormones and hemodynamics in heart failure. The New England journal of medicine, 1999. 341(8): p. 577-585.

494. Mark AL, K.R., Augmentation of cardiopulmonary baroreflex control of forearm vascular resistance in borderline hypertension. Hypertension, 1982. 4(1): p. 39-46.

495. Volpe M, T.B., Ricciardelli B, Vigorito C, de Luca N, Rengo F, Condorelli M, The autonomic nervous tone abnormalities in the genesis of the impaired baroreflex responsiveness in borderline hypertensive subjects. Clinical science, 1982. 62(6): p. 581- 588.

496. Trimarco B, L.G., Ricciardelli B, De Luca N, Rendina V, Condorelli G, Volpe M, Salt-induced plasticity in cardiopulmonary baroreceptor reflexes in salt-resistant hypertensive patients. Hypertension, 1991. 18(4): p. 483-493.

497. Lembo G, M.C., Lanni F, Barbato E, Vecchione C, Fratta L, Trimarco B., Systemic hypertension and coronary artery disease: the link. American journalo of cardiology, 1998. 82(3A): p. 2H-7H.

498. Julius S, Hemodynamic and neurohumoral evidence of multifaceted pathophysiology in human hypertension. Journal of cardiovascular pharmacology, 1990. 15(Suppl 5): p. S53-58.

275 499. Ferrario CM, The renin-angiotensin system: importance in physiology and pathology. Journal of cardiovascular pharmacology, 1990. 15( Suppl 3): p. S1-5.

500. Vanhoutte PM, Endothelial dysfunction and vascular disease. Verhandelingen - Koninklijke Academie voor Geneeskunde van Belgie, 1998. 60(3): p. 251-266.

501. Freiman PC, M.G., Heistad DD, Armstrong ML, Harrison DG, Atherosclerosis impairs endothelium-dependent vascular relaxation to acetylcholine and thrombin in primates. Circulation research., 1986. 58(6): p. 783-789.

502. Shimokawa H, V.P., Impaired endothelium-dependent relaxation to aggregating platelets and related vasoactive substances in porcine coronary arteries in hypercholesterolemia and atherosclerosis. Circulation research., 1989. 64(5): p. 900-914.

503. Zeiher AM, D.H., Wollschlager H, Just H, Modulation of coronary vasomotor tone in humans. Progressive endothelial dysfunction with different early stages of coronary atherosclerosis. Circulation, 1991. 83(2): p. 391-401.

504. Ginsburg R, B.M., Davis K, Dibiase A, Billingham ME, Quantitative pharmacologic responses of normal and atherosclerotic isolated human epicardial coronary arteries. Circulation, 1984. 69(2): p. 430-440.

505. Bossaller C, H.G., Yamamoto H, Williams C, Wells S, Henry PD, Impaired muscarinic endothelium-dependent relaxation and cyclic guanosine 5'-monophosphate formation in atherosclerotic human coronary artery and rabbit aorta. The Journal of clinical investigation, 1987. 79(1): p. 170-174.

506. Tesfamariam B, W.R., Cohen RA, Augmented adrenergic contractions of carotid arteries from cholesterol-fed rabbits due to endothelial cell dysfunction. Journal of cardiovascular pharmacology, 1989. 6(13): p. 820-825.

507. Wu CC, B.D., Role of endothelium in the response to endothelin in hypertension. Hypertension, 1990. 16(6): p. 677-681.

508. Howes LG, K.H., Plasma lipoproteins, cardiovascular reactivity and the sympathetic nervous system. Journal of autonomic pharmacology, 1989. 9(4): p. 293-301.

509. Chin JH, A.S., Hoffman BB, Inactivation of endothelial derived relaxing factor by oxidized lipoproteins. The Journal of clinical investigation, 1992. 89(1): p. 10-18.

510. Harrison DG, A.M., Freiman PC, Heistad DD, Restoration of endothelium- dependent relaxation by dietary treatment of atherosclerosis. The Journal of clinical investigation, 1987. 80(5).

511. Kassab E, M.S.I., Sower J R, Vascular complications in diabetes and their prevention. Vascular medicine, 2001. 6(4): p. 249-255.

276 512. Taylor AA, Pathophysiology of hypertension and endothelial dysfunction in patients with diabetes mellitus. Endocrinology and metabolism clinics of North America, 2001. 30(4): p. 983-997.

513. Cohen RA, T.B., Weisbrod RM, Zitnay KM, Adrenergic denervation in rabbits with diabetes mellitus. American journal of physiology, 1990. 259(1 Pt 2): p. H55-61.

514. Hattori Y, K.H., Abe K, Kanno M, Superoxide dismutase recovers altered endothelium-dependent relaxation in diabetic rat aorta. American journal of physiology, 1991. 261(4 Pt 2): p. H1086-1094.

515. Winquist RJ, B.P., Baskin EP, Wallace AA, Decreased endothelium-dependent relaxation in New Zealand genetic hypertensive rats. Journal of hypertension, 1984. 2(5): p. 541-545.

516. Dohi Y, T.M., Buhler FR, Luscher TF, Activation of endothelial L-arginine pathway in resistance arteries. Effect of age and hypertension. Hypertension, 1990. 16(2): p. 170- 179.

517. Hongo K, N.T., Kassell NF, Sasaki T, Lehman M, Vollmer DG, Tsukahara T, Ogawa H, Torner J, Effects of aging and hypertension on endothelium-dependent vascular relaxation in rat carotid artery. Stroke, 1988. 19(7): p. 892-897.

518. Cardillo C, K.C., Quyyumi AA, Cannon RO 3rd, Panza JA, Selective defect in nitric oxide synthesis may explain the impaired endothelium-dependent vasodilation in patients with essential hypertension. Circulation research., 1998. 97(9): p. 851-856.

519. Miyoshi A, S.H., Fujiwara M, Masai M, Iwasaki T, Impairment of endothelial function in salt-sensitive hypertension in humans. American journal of hypertension, 1997. 10(10 Pt 1): p. 1083-1090.

520. De Artinano AA, G.V., Endothelial dysfunction and hypertensive vasoconstriction. Pharmacological research, 1999. 40(2): p. 113-124.

521. Nava E, L.T., Endothelium-derived vasoactive factors in hypertension: nitric oxide and endothelin. Journal of hypertension. Supplement, 1995. 13(2): p. S39-48.

522. Cines DB, P.E., Buck CA, Loscalzo J, Zimmerman GA, McEver RP, Pober JS, Wick TM, Konkle BA, Schwartz BS, Barnathan ES, McCrae KR, Hug BA, Schmidt AM, Stern DM, Endothelial cells in physiology and in the pathophysiology of vascular disorders. Blood, 1998. 91(10): p. 3527-3561.

523. Cohen RA, Pathways controlling healthy and diseased arterial smooth muscle. The American journal of cardiology, 1993. 72(8): p. 39C-47C.

524. Joyner MJ, D.N., Shepherd JT, From Belfast to Mayo and beyond: the use and future of plethysmography to study blood flow in human limbs. Journal of applied physiology, 2001. 91(6): p. 2431-2441.

277 525. Wilkinson IB, W.D., Venous occlusion plethysmography in cardiovascular research: methodology and clinical applications. British journal of clinical pharmacology, 2001. 52(6): p. 631-646.

526. Cramer M, B.K., Strandness Jr. DE, The detection of proximal deep vein thrombosis by stain gauge plethysmography through the use of an outflow/capacitance discriminant line. Bruit, 1983. VII.

527. Cramer M, L.Y., Beach K, Martin D, Strandness DE Jr., Standardization of venous flow measurements by strain gauge plethysmography in normal subjects. Bruit, 1983. VII: p. 33-39.

528. Rueckert PA, H.P., Comparison of arterial occlusion and ischaemic exercise for the study of vasodilatation in the human calf. Clinical Science, 1995. 88(6): p. 643-649.

529. Radegran G, Limb and skeletal muscle blood flow measurements at rest and during exercise in human subjects. The Proceedings of the Nutrition Society, 1999. 58(4): p. 887-898.

530. McCully KK, P.J., The application of blood flow measurements to the study of aging muscle. The journals of gerontology, 1995. 50A(Special issue): p. 130-136.

531. LeJemtel TH, T.M., Jondeau G, Direct and indirect assessment of skeletal muscle blood flow in patients with congestive heart failure. Journal of molecular and cellular cardiology, 1996. 28(11): p. 2249-2254.

532. Feigenbaum H, Echocardigraphy. Fifth ed. 1994, Malvern: Lea & Febiger. 695.

533. Kimmerly DS, S.J., Hypovolemia and neurovascular control during orthostatic stress. American journal of physiology., 2002. 282(2): p. H645-655.

534. Van Beekvelt MC, S.J., Tschakovsky ME, Hopman MT, Hughson RL, Blood flow and muscle oxygen uptake at the onset and end of moderate and heavy dynamic forearm exercise. American journal of physiology, 2001. 280(6): p. R1741-1747.

535. Lee F, S.J., McQuillan PM, Kunselman AR, Smith MB, Yang QX, Smith H, Gray K, Sinoway LI, Effects of forearm bier block with bretylium on the hemodynamic and metabolic responses to handgrip. American journal of physiology, 2000. 279(2): p. H586- 593.

536. Shoemaker JK, P.Z., Hughson RL, Forearm blood flow by Doppler ultrasound during test and exercise: tests of day-to-day repeatability. Medicine and science in sports and exercise, 1996. 28(9): p. 1144-1149.

537. Welsch MA, A.J., Geaghan GP, Stability and reproducibility of brachial artery flow mediated dilation. Medicine and science in sports and exercise, 2002. 34(6).

278 538. Quick HH, D.J., Ladd ME, MR imaging of the vessel wall. European radiology, 2002. 12(4): p. 889-900.

279 APPENDIX B: PROPOSAL

280 CHAPTER 1. INTRODUCTION

1.1. SPECIFIC AIMS

Selye’s theory of stress-response-adaptation (3) suggests an exercise stress, when applied to the vascular system, results in an increased delivery of blood flow to active skeletal

muscle, and if the stress is repeated over time contributes to a series of vascular

adaptations. These vascular changes are aimed at increasing vascular conductance and

oxygen and nutrient delivery to working skeletal muscle during subsequent exercise

bouts (7, 10, 13, 16, 31, 33). In contrast to the extensive literature describing the

structural and functional changes in the arterial system following training, training-

induced alterations in the venous system are less clear. Given the importance of the

venous circulation in regards to metabolic waste removal, its enormous potential to act as

a volume reservoir and its contribution to the end-diastolic volume, changes in this

segment of the vasculature could have a profound impact on the exercise response. To

this extent, Louisy et al. (23) reported greater venous distensibility in endurance-trained

subjects compared to healthy controls and Miyachi et al., (26) reported an increase in

peripheral (femoral) venous size following one-legged training. Both investigators

suggested the veno-adaptations could be associated with increases in blood volume

following endurance training, and that the increased blood volume may be distributed on

the venous side of the vascular space, since the veins represent a more compliant

reservoir. Recognizing the potential influence of systemic factors such as blood volume

in several previous studies, the purpose of this proposal is to examine the effects of 4-

weeks of handgrip exercise training on arterial reactivity and venous function,

particularly venous capacitance and venous outflow.

281 1.2. HYPOTHESES

This study will test two specific hypotheses regarding the influence of physical training

on arterial (Hypothesis I) and veno-dynamics (Hypothesis II). Finally, depending on the

outcome of the first two hypotheses, the third hypothesis will test the timecourse of the expected changes in vascular dynamics.

Hypothesis I Four-weeks of handgrip exercise training will result in an increase in arterial reactivity

following 5-minutes of arm occlusion;

Hypothesis II Four-weeks of handgrip exercise training will result in an increase in venous capacitance

and venous outflow at rest and following 5-minutes of arm occlusion;

1.3. BACKGROUND AND SIGNIFICANCE

1.3.1. Previous work The effects of exercise training on hemodynamic and hematologic factors have been

extensively investigated. One of the early studies examined the effect of exercise training

on regional vascular function reported a decrease in blood flow. The authors attributed

lowered leg blood flow to a more “effective” metabolic process (30). However Sinoway and coworkers reported increased vascular reactivity in the dominant arm of tennis

players (34) and following handgrip exercise training (35). These findings were further

confirmed in a number of studies and attributed to alterations in vascular smooth muscle

and/or vascular endothelium structure and function, however the influence of cardiac and

skeletal muscle adaptations were not totally excluded (3, 17).

282 One of the most predictable hemodynamic changes following “aerobic” training is

an increase in blood volume. Consequently, venous return is higher and stroke volume is

elevated at each absolute level of work. Recognizing that the venous system is a reservoir

for ~65% of total blood volume and is repeatedly challenged during exercise, it is

reasonable to assume that this system would be affected by exercise training. However

few studies examined the effect of exercise training on venous function, therefore

conclusions regarding circulatory adaptations have for the most ignored the possible

contribution of the venous system (11, 18, 29). The following paragraphs will provide a

brief review of the available literature regarding venous adaptations following exercise

training.

Wecht et al. examined the difference in venous hemodynamics between

individuals with spinal cords injury, sedentary and active individuals. Venous tone,

venous capacitance, venous compliance, and total venous outflow were higher in the

active group, leading the authors to conclude that endurance exercise training enhanced

venous function (41). Louisy and coworkers examined venous hemodynamics in

endurance trained, strength trained and sedentary subjects and reported higher values for leg-filling volumes and venous emptying in the endurance trained subjects. The authors

concluded that endurance training increased lower limb venous hemodynamics

subsequent to hypervolemia rather than alterations of the visco-elastic properties of the

vein walls (23).

Miyachi et al. examined venous diameter following 6 weeks of one-legged cycle training at 80% peak oxygen consumption, 40 min/day for 4 days/week followed by 6 weeks of detraining. Arterial and venous cross-sectional areas increased by 16% and

283 46%, respectively and correlated with an increase in one-leg peak oxygen consumption.

These changes returned to baseline during detraining. The authors attributed part of the

changes in oxygen consumption to arterial adaptations and improved blood delivery.

Furthermore, they suggested the veno-adaptations could be associated with increases in

blood volume following endurance-training, and that the increased blood volume may be

distributed on the venous side of the vascular space, since the veins represent a more

compliant reservoir (26).

Suzuki and colleagues examined the effects of exercise training on venous

capillary density of young (3 weeks) and middle-age (54 weeks) (37). The 6-week

exercise-training program resulted in increased arteriolar capillary density in both groups

without changes in venular density. In a subsequent time-course study by the same group,

the effect of exercise training on venular density was examined in young female rats (5

weeks). Vascular density indices (capillary-to-fiber ratio, total capillary density, and

density of venular capillaries) of soleus and medial gastrocnemius muscles increased

following only 1 week of training. However values for total capillary, arteriolar, and

venular capillaries were significantly decreased after 4 weeks of training in the soleus

muscle and after 3 weeks in the medial gastrocnemius. These results suggest that 1-week of exercise training in Wistar rats results in greater arteriolar and venular capillarization, mainly in highly oxidative muscles and indicate that training-induced adaptations might be time-, age- or gender-dependent. However since neither arterial or venous hemodynamics nor any vascular function indicators have not examined, a complete

conclusion were not made (37, 38).

284 Kenney and Armstrong investigated venous pooling following an 8-week

exercise-training program (40 min/day, 3day/week). Venous pooling was measured in the

foot as volume drainage following the “Trendelenburg” maneuver (VT). The

“Trendelenburg” maneuver consists of 45º leg elevation for 15 seconds with the subjects in a supine position, followed by sitting up to a position of 90º flexion with the legs lower than the body. The investigators assisted the subject throughout the maneuver to avoid muscle contraction. Moreover, to account for the muscle pump contribution, volume drainage was evaluated following 15 seconds of dorsi-flexion/plantar exercise while the foot is lower than the body (VE) then subtracted from volume drainage following the

Trendelenburg maneuver (VT-VE). Exercise training resulted in increased VT, VT-VE, oxygen consumption without changes in VE. The authors concluded that the increased venous pooling was the result of hypervolemia. However, the possible structural and/or functional changes in the venous system following exercise training were not addressed

(21). Hypervolemia associated with exercise training, has also been linked to changes in central venous pressures as reported by Convertino et al. Convertino used an exercise program consisting of 10 week of cycle exercise for 30 min/day, 4 days/week at 75-80% to examine changes in central venous pressure. Training resulted in a 20% increase in maximal O2 uptake, decreased resting heart rate, 9% increase in blood volume and increased central venous pressure. The increase in central blood pressure was associated with the increase in blood volume without changes in total effective venous capacitance,

thus the authors attributed the observed changes in central venous pressure to an increase

in blood volume but not to alterations in venous structure or function (12).

285 Detraining has also been shown to induce changes in the venous system (20, 22,

32, 42). Louisy and Guezennec examined the effect of 30-days head-down bedrest on

lower-body venous hemodynamics and reported a progressive increase in venous

compliance. The study was not designed to reveal the mechanisms involved, however

reduction in the blood volume and subsequent dilation in the venous system and

decreased pressure in the proximity of the lower veins were stipulated. Involuntarily

detraining induced by morbidity diseases might also affect venous function (22).

Ikenouchi and colleagues found that venous distensibility is reduced and related to the

severity of the disease in patients with heart failure (20, 32). In our laboratory venous capacitance and outflow were lowered and related to exercise intolerance in heart failure, confirming the relationship between inactivity and venous dysfunction (42). While venous changes might be a direct effect of heart failure, inactivity associated with the disease might indirectly alter venous function.

In summary, the aforementioned studies suggest higher levels of fitness and/or exercise training may alter venous hemodynamic. The changes in venous hemodynamics appears to be directly related to changes in blood volume and associated with changes in exercise capacity. In contrast detraining reveres these adaptations, which may be further affected by systemic diseases such as heart failure. However, none of the above- mentioned studies have clearly addressed the issue whether exercise improves measures of venous function and the manner in which this relates to muscle performance.

Barclay et al. demonstrated the importance of metabolite removal to muscle fatigue, indicating the essential role of venous hemodynamics to performance. In this study, muscle fatigue rate in dogs was measured in 5 experimental groups subjected to

286 different perfusion conditions by manipulating O2 contents and/or muscle blood flow with a perfusion pump. The first group served as a control group and breathed room air while the muscles were self-perfused (self-perfused). Group 2, breathed 14% O2 while

increasing blood flow across the muscle bed with the perfusion pump (low O2/ high flow)

and examined the effect of hyperfusion on muscle fatigue. In group 3, the dogs breathed air saturated with 100% O2 while blood flow was at reduced perfusion rate (high O2/low

flow) and examined the effect of hypofusion on muscle fatigue. In group 4, blood flow

pump-perfused while perfusion pressure was kept constant at the animal’s blood pressure

until a steady pressure was attained. This procedure kept the plasma flow constant and

identified the effect of blood and plasma flows at constant level established by the

th animal’s blood pressure (constant flow). Finally, the 5 group breathed 100% O2 and the

blood was diluted with a volume-for-volume exchange of dextran solution while

maintaining constant plasma flow by increasing blood flow with the perfusion pump to

identify the effect of hyperfusion with a constant plasma flow (diluted/high flow). The

self-perfused group fatigued the fastest while the pump-perfused groups had slower

fatigue rate regardless of O2 delivery or plasma flow rates. The authors argued that

increased blood flow across the muscle bed resulted in changed cellular environment

and/or metabolic pattern. Additionally they concluded that the attenuated fatigue

associated with hyperfusion was independent of O2 and/or substrate delivery and is

attributed to improved metabolite removal out of the muscle bed (8). This suggests that

venous health is particular important in exercise performance subsequent to ensuring

adequate metabolite removal.

287 More recently, Tschakovsky and Hughson also examined the importance of venous emptying in the vasodilatory capacity and muscle function. Continuous forearm venous volume and blood flows were obtained following arm elevation while either venous drainage is prevented or allowed (Fig 1.1).

Figure PR1.1. Schematic illustration of the timing of arm position relative to heart level during the prolonged and the acute arm elevation protocols. This timing was the same for both venous emptying and venous congestion trials. Solid bars indicate transition period between arm positions. Taken from (40).

As demonstrated in Fig 1.2, venous emptying protocol resulted in marked decreased blood volume associated with transit increase in blood flow during acute arm elevation while the opposite was observed during venous occlusion protocol. The hyperemic respond was attributed to transit vasodilatory effect consequent to reduced forearm venous volume. Additionally, this increase in forearm blood flow was linked to venoarteriolar reflex vasoconstriction with reduction in venous volume and was

288 suggested to simulate hyperemic response subsequent to venous emptying during muscle contractions (40).

Figure PR1.2. Hemodynamic response to acute forearm elevation. Data are means ± with values at 1-s intervals. A: forearm blood flow; B: changes in forearm volume from below heart baseline; C: forearm venous pressure (antecubital vein). Forearm began below heart (time = 10 s to 0 s). Arm above heart from 2 s to 6 s. Venous emptying allowed during arm elevation ( ). Venous emptying prevented during arm elevation ( ). Solid bars indicate the 2-s transition periods between arm positions; n = 9 except for venous pressure, venous emptying allowed n = 4, venous emptying prevented n = 3. Taken from (40).

Thus, these above-mentioned studies suggest a critical role of venous outflow in

exercise performance. Consequently, studies are needed to examine the effects of acute

and chronic exercise training, as well as the modality, intensity, duration, and frequency

289 of training, on different segments of the venous system. Additionally investigations

examining the effects of a localized exercise program on venous alterations are certainly

warranted.

1.3.2. Progress report / Preliminary studies In preparation of this proposal a series of studies have been conducted to develop the

appropriate methodologies and understand the rationale for this proposal. These studies

focused on determining the reproducibility of several outcome measures pertinent to this

study, examined the ability of vascular measures to differentiate between various

populations, and evaluated the influence of interventions on changing measures of

vascular function. These projects will be summarized in chronological order in the

following section.

The first study examined the effect of 4 weeks of low (25% of MVC) and high

(75% of MVC) intensity handgrip (HG) exercise training on arterial function. Forearm blood flow (Binf) and vascular resistance (VR) were evaluated in twenty-eight young men at rest, following forearm occlusion (OCC) and following forearm occlusion combined with HG exercise (HGOCC) using mercury strain gauge plethysmography

(MSGP) (3). The 4-week program consisted of non-dominant HG exercise performed 5 d/week for 20-min at either low or high intensity. Both training intensities resulted in improved vasodilatory responsiveness in the trained arm only with no change in muscular strength (see figures 1.3 & 1.4 and table 1.1).

290 ‡ † § 45 § § * § 40 ‡ * † * Pre * 35 * * * Post 30 * (ml.100ml-1.min-1) 25 20 15 10 5 0 Rest OCC HGOCC Rest OCC HGOCC Forearm Blood Flow Dominant Arm Non-dominant Arm

Figure PR1.3. Forearm Binf changes in the low group following HG exercise training. Values are means ± SD. *p<0.05 vs. rest; §p<0.05 vs. OCC ‡p<0.05 vs. Pre Non- dominant; †p<0.0 (3).

† § ‡ * 45 § § § ‡ † 40 * * * * * 35 * * (ml.100ml-1.min-1)

30 25 Pre 20 Post 15 10 5 0 Rest OCC HGOCC Rest OCC HGOCC

Forearm Blood Flow Dominant Arm Non-dominant Arm

Figure PR1.4. Forearm Binf changes in the high group following HG exercise training. Values are means ± SD. *p<0.05 vs. rest; §p<0.05 vs. OCC ‡p<0.05 vs. Pre Non- dominant (3).

291

Table PR1.1. Dominant and non-dominant forearm VR changes in the low and high groups at rest, following OCC and following HGOCC (3). Dominant Non-dominant

Pre Post Pre Post

Low group (15)

VRrest 47.85 ± 13 42.2 ± 13.4 42.5 ± 15.4 44.6 ±17.0

VROCC 3.3 ± 1* 3.5 ± 0.5* 3.5 ± 0.6* 3.5 ± 0.8*

VRHGOCC 2.5 ± 0.5* 2.8 ± 0.4* 3 ± 0.4* 2.9 ± 0.7*

High group (13)

VRrest 44.5 ± 19 45.7 ± 22.7 39.4 ± 15 34.4 ± 8.8

VROCC 3.1 ± 0.5∗ 3.1 ± 0.5∗ 3.7 ± 0.6∗ 3.3 ±0.7∗

VRHGOCC 2.7 ± 0.3∗ 2.77 ± 0.4∗ 3.3 ± 0.7∗ 2.5 ± 0.7∗

Values are means ± SD. LO, training intensity = 25% of maximum voluntary contraction (MVC); HI, training intensity = 75% of MVC; VRrest, forearm vascular resistance at rest; VROCC, vascular resistance following five minute of occlusion; VRHGOCC, vascular resistance following OCC coupled with 3-min. of HG of exercise. ∗p < 0.05 vs. rest.

Table PR1.2. Independent t-test comparisons between CHF and Control groups (42). CHF Control p-value (n=20) (n=9) Binf (ml•100ml tissue-1 •min-1) Rest 3.10 ± 1.20 3.43 ± 1.24 0.50 Post Occlusion 15.25 ± 6.31 22.30 ± 6.70 0.018 VR (U) Rest 36.90 ± 14.87 30.40 ± 10.46 0.24 Post Occlusion 7.80 ± 3.07 4.64 ± 1.38 0.007 Vcap (ml•100ml tissue-1 •min-1) Rest 1.88 ± 0.44 2.17 ± 0.22 0.068 Post Occlusion 1.43 ± 0.50 2.04 ± 0.40 0.003 Vout (ml•100ml tissue-1 •min-1) Rest 20.50 ± 6.40 22.90 ± 10.78 0.46 Post Occlusion 24.50 ± 9.42 33.11 ± 9.97 0.04

292 In a subsequent study forearm Binf, VR, venous capacitance (Vcap) and outflow (Vout)

and maximum walking distance (MWD) were evaluated in heart failure patients and age-

matched older adults (42). Arterial and venous function indices and MWD were different

between the groups (table 1.2). Additionally, correlation analyses revealed significant

associations between vascular indices and exercise tolerance measurements (figures 1.2

& 1.3), confirming impaired forearm vasoreactivity and indicating that venous function

might contribute to exercise intolerance in heart failure.

800 700 r =0.57; p =001 600 500 400

MWD (m) 300 200 100

.25 .75 1.25 1.75 2.25 2.75 Vcap (ml·100ml tissue-1·min-1)

Figure PR1.5. Relationship between forearm venous capacitance and maximum walking distance (42).

Interestingly many of the available protocols to examine venous function using

MSGP have failed to address many physiological and methodological concerns that might directly or indirectly affect the precision of the measurements (4, 5). A particular important factor to consider is the selection of the pressure and time of venous occlusion.

For example Hokanson et al. and Cramer et al. have recommended 50mmHg for venous occlusion pressure, regardless of the pressure of the arterial system (14, 15, 19). While this pressure is arguably sufficient to prevent venous emptying, in some cases it may also

293 change the kinetics of the arterial system. In a more recent study, Wecht et al. used a

pressure “just below” diastolic blood pressure as a criterion for venous occlusion

pressure. However the authors did not quantify the precise pressure used to occlude the

venous system, thereby interpretations of their findings were ambiguous (41).

800 700 r =0.67; p =0001 600 500 400

MWD (m) 300 200 100

10 15 20 25 30 35 40 45 50 55 60 Vout(ml·100ml tissue-1·min-1)

Figure PR1.6. Relationship between forearm venous outflow and maximum walking distance (42).

Furthermore, the aforementioned studies (14, 15, 19) have also adapted two minutes of venous occlusion to evaluate the venous system, however it is not clear whether this time is sufficient to achieve maximum venous filling. Thus the next phase involved examining the effect of different venous occlusion times and pressures on Vcap and Vout. Venous indices were examined this in ten young individuals using two distinct protocols. In protocol one venous occlusion pressures of 7, 14, 21, 28 and 35 mmHg below resting diastolic blood pressure were used with a fixed occlusion time of four minutes.

294 In protocol two venous occlusion times of 5 and 10 minutes were used with a

fixed venous pressure of 50mmHg. Figures 1.5 and 1.6 depict a stepwise decrease in

Vcap (p ≤ 0.0001) and Vout (p ≤ 0.0001) with greatest values following venous occlusion

pressure 7mmHg below diastolic blood pressure. Forearm Vcap ranged between 3.7±0.7

and 2.2±0.5 ml•100ml tissue-1•min-1 and Vout ranged between 32.7±4.5 and 17.3±4.0

ml•100ml tissue-1•min-1.

6 1 = 07mmHg < DBP 2 = 14mmHg < DBP 5 3 = 21mmHg < DBP 4 = 28mmHg < DBP 5 = 35mmHg < DBP 4 ‡ ‡ ‡ * § 3 *

2 Vcap (ml·100ml tissue-1·min-1)

1 12345 Pressures (mmHg)

Figure PR1.7. Venous capacitance using different pressures. Values are in means ± SD. ‡p<0.007 vs. 1; *p<0.01 vs. 2; §p<0.03 vs. 3 (36). Furthermore, 10 minutes of venous occlusion resulted in greater Vcap (p < 0.001) and Vout (p < 0.026) values as compared to 5 minutes (Figure 1.7). Finally, the effects of altering occlusion times and pressures on resting arterial inflow and vascular resistance were minimal and not statistically significant. These findings indicate the importance of recognizing the possible influence of venous occlusion times and pressures on venous capacitance and outflow. Additionally these technical aspects are particularly relevant when considering repeated measures designs during which blood pressure may be altered

(36).

295

60 1 = 07mmHg < DBP 2 = 14mmHg < DBP 3 = 21mmHg < DBP 50 4 = 28mmHg < DBP 5 = 35mmHg < DBP 40

‡ ‡ ‡ § * 30 * †

Vout (ml·100ml tissue-1·min-1) 20

10 12345 Pressures (mmHg)

Figure PR1.8. Venous outflow using different pressures. Values are in means ± SD. ‡p<0.02 vs. 1; *p<0.02 vs. 2; §p<0.05 vs. 3; †p<0.06 vs. 4 (36).

5 40

4 30 ‡

3 20 Vout (ml·100ml tissue-1·min-1) Vcap (ml·100ml tissue-1·min-1)

2 10 5Minutes 10Minutes 5Minutes 10Minutes Time Time

Figure PR1.9. Venous capacitance and outflow following 5 & 10 minute venous occlusion Values are in means ± SD. ‡p<0.026 vs. 10Minutes (36).

296 With attempts made to investigate the role of the venous system for understanding cardiovascular physiology, it is imperative to have a reliable investigative model for venous function. Our laboratory (24) as well as others (28) has reported adequate reproducibility and reliability of the plethysmographic technique to measure arterial inflow. In contrast, few studies have examined the reproducibility and reliability of plethysmography to measure venous function. Thus another aspect of the research process involved examination of the reproducibility of venous and arterial indices.

Forearm Binf, VR, Vcap and Vout were evaluated at rest and following OCC and

HGOCC on three different occasions within 10 days and expressed in ml•100ml tissue-

1•min-1. As revealed in table 1.3, between-days interclass correlation coefficients for arterial and venous indices at rest and following OCC and HGOCC ranged between 0.7 and 0.97 (2). The values for the resting Binf and VR are well within expected range, additionally no significant difference were found between the means values of the three visits. However a slight day-to-day difference might result in low statistical reproducibility (34, 35). These results indicate adequate reliability of MSGP technique for measuring arterial as well as venous function on different days under controlled conditions.

Table PR1.3. Between-days interclass correlation coefficients for forearm arterial and venous indices (2). Condition Binf VR Vcap Vout Rest 0.29 0.3 0.78 0.94 OCC 0.81 0.7 0.7 0.91 HGOCC 0.95 0.95 0.8 0.97 Binf=Blood inflow, VR=Vascular resistance, Vcap=Venous capacitance, Vout=Venous outflow

297 Finally the effects of functional ability on arterial and venous measures have been examined. This is extremely relevant as this research is ultimately trying to help clinicians understand why individuals may be intolerant to exercise. Linking a pathologic condition to a functional limitation will eventually result in the development of more appropriate strategies to preserve function and avoid disability. Consequently, several studies have attempted to link performance measures with arterial and venous function.

Most recently, HG strength on venous function, forearm Vcap and Vout were evaluated in 13 apparently healthy males at rest and following OCC using MSGP. Maximum HG strength was determined as the average of 3 maximum consecutive contraction trials using a hand-dynamometer and expressed in kilogram. Subsequently, the group was divided into above (HI) and below (LO) average HG strength according to the Canadian

Society of Exercise Physiology classification. As depected in figure 1.8 Pearson correlation coefficient showed an association between HG strength and forearm venous indices at rest {Vcap: r=0.72; p=0.006; Vout: r=0.72; p=0.006} and following OCC

{Vcap: r=0.56; p=0.058; Vout: r=0.73; p=0.007} (2).

7 70

60 6

50 5 40

4 30

Vcap (ml.100ml-1.min-1) r =0.72; p =0.006 r =0.72; p =0.006 3 Vout (ml.100ml-1.min-1) 20 30 80 130 180 30 80 130 180 Strength (kg) Strength (kg)

Figure PR1.10. Relationship between HG strength & resting forearm venous indices (2).

298 Finally, Student t-test comparison between the HI and LO HG strength groups

revealed difference in Vcap and Vout at rest and post OCC (table 1.4). These results

suggest that MSGP is useful technique to differentiate between populations. Furthermore,

HG strength appears to contribute to venous system hemodynamics implying that muscular conditioning might be important for venous health.

Table PR1.4. Differences in venous indices between low and high strength groups (2) Group HG Strength (kg) Vcap at rest Vcap post OCC Vout at rest Vout post OCC LO 80.55±16.2 4.5±0.8 3.4±0.7 38.7±9.2 38.0±7.5 HI 136.8±20.7 5.7±0.5 4.2±0.7 54.3±7.6 54.0±8.2 Alpha 0.0001 0.002 0.065 0.01 0.007

In conclusion, the completed research has determined that assessment of arterial

and venous function using plethysmography is reproducible and reliable, allows

differentiation between populations, and arterial function is modifiable with training. The

final aspect of this research is to examine the influence of localized exercise training on

venous measures.

1.5. DISCUSSION OF EXPECTED RESULTS

Due to the relatively small muscle size involved in the exercise-training program,

physiological adaptations are expected to be limited. Arm ergometer test, measures of

autonomic nervous system-HRV, body composition are expected to remain unchanged

throughout the exercise-training program (33, 34). Additionally as previously reported

arm circumference as well as MVC are expected to increase by 2 (1) and 6-8% (1, 3),

respectively.

299 Vascular indexes As previously reported, HG exercise training is expected to result in an increase in arterial reactivity following 5-minutes of arm occlusion (3, 17, 34, 35). The effect of localized exercise training program on venous function have not been evaluated.

However few studies have examined the effect of exercise-training programs that involve large muscle groups and trigger systemic adaptations, including hypervolemia. Therefore increase in venous capacitance and venous outflow were attributed to systemic changes while ignoring possible changes in the venous function (21, 23, 41). Given the importance of the venous system in overall cardiovascular physiology (8, 40) and since the venous system will be challenged with exercise training, improvement are expected in venous capacitance and venous outflow.

300 CHAPTER 2. METHODS

2.1. PARTICIPANTS AND DESIGN

Fifteen apparently healthy sedentary young (ages 18-30) adult males will be recruited to participate in the study. Individuals with manifestations of cardiovascular, metabolic, orthopedic or neurological disease or are on any medication (e.g., digitalis), which could affect the results of this study will be excluded.

The proposed study is randomized, prospective, controlled and will be executed in

4 weeks. The study involves once a week evaluation session and 5-time-a-week exercise training sessions. Prior to and at the end of each week, participants will undergo HG strength and vascular function evaluations in the dominant and non-dominant arms. The primary outcomes of this study include: (1) Vasoreactivity, defined as the change in forearm blood flow from rest to the peak blood flow response following 5 minutes of forearm occlusion (34, 35); (2) Venous capacitance; defined as the changes in limb blood volume following 6 minutes of venous occlusion; and (3) Venous outflow; defined as the changes in limb blood volume following the release of venous occlusion (15, 19).

Additionally, handgrip strength, arm exercise tolerance, and hemodynamic measures (i.e. blood pressure, heart rate and heart rate variability) will also be assessed. Following the completion of the initial assessment, each subject will be asked to report to the exercise physiology laboratory 5 days/week to perform 20 minutes of forearm flexion exercise at a cadence of 1 contraction every 4 seconds starting at intensity of 60%. The non-dominant arm will be trained while the dominant arm will serve as control.

301 2.2. STUDY MEASUREMENTS:

2.2.1. Vascular function assessments Vascular function indices will be obtained in the dominant and non-dominant forearms using MSGP at rest and following OCC (34) at the beginning of the study, and at the end of each week throughout the study period. This technique is noninvasive based on the assumption that alterations of pressures in strategically placed cuffs allows examination of the rate of change of limb volume thought to reflect venous function indices including

Binf, VR, Vcap and Vout (15, 19). Prior to the experiment, blood pressure cuffs will be positioned around the participant’s upper arm and wrist, and a mercury-in-silastic strain gauge placed around the forearm approximately 10 cm distal to the olecranon process (3).

Resting forearm vascular function measures will be obtained following 15 minutes of supine rest. Immediately before the measurements, hand circulation will be occluded for 1 minute by inflating the cuff at the wrist to 240mmHg. Forearm Binf will be obtained using an upper arm venous occluding pressure of 7mmHg below diastolic blood pressure.

Forearm Vcap will be measured after an additional 6 minutes of venous occlusion and

Vout following the release of pressure (2, 36). Subsequently, forearm vasoreactivity is examined following 5 minutes of forearm arterial occlusion achieved by inflating the cuff on the upper arm to 240mmHg. Forearm vascular indices will then be determined as described above. The exact same procedures will then be performed on the other arm.

Prior to, during and following each procedure blood pressure and heart rate will be obtained.

2.2.2. Autonomic nervous function assessments: Indices of autonomic function will be assessed using measures of heart rate variability.

These measures will be obtained at the same time as the vascular assessments described

302 above. The ECG electrodes will be placed on the chest and interfaced with a Biopac

MP100 and its companion software Acqknowledge (model MP100A, Biopac Inc., Santa

Barbara, CA) to allow for continuous data acquisition. A period of five minutes (at rest

and during forearm occlusion) of ECG data will be analyzed for mean heart period (mean

R-R interval), standard deviation (SDNN), low frequency power (LF) and high frequency

power (HF) using the guidelines set forth by the Task Force of the European Society of

Cardiology and the North American Society of Pacing and Electrophysiology (9, 39).

These variables will be obtained at the beginning of the study, and at the end of each

week of the study period.

2.2.3. Arm ergometer test Arm exercise tolerance will be evaluated using a graded arm ergometer exercise test at the beginning and at the end of the study period (6, 25). The test will be performed by

initially turning the arm ergometer at a speed of 50 rpm and a resistance of 0.5kp for two

minutes while the participant in a seated position. The resistance will then be increased

by 0.25kp every two minutes until the subject reaches exhaustion, requests to stop, or

exhibits signs or symptoms, which require test termination. Prior to, during, and

following the exercise test heart rate, blood pressure measurements and ratings of

perceived exertion will also be obtained at the end of each minute throughout the test using a standard sphygmomanometer and the Borg's perceived exertion scale,

respectively.

2.2.4. Forearm circumference and HG strength assessments Participants forearm circumferences and HG maximal voluntary contraction will be evaluated at baseline and at the end of each week throughout the study period. Forearm circumference will be examined using a weighted measuring tape 10 cm distal to the

303 midpoint between the lateral epicondyle and olecranon process. Handgrip strength will be evaluated using a handgrip dynamometer (Lafayette instruments) with the participant upright, but slightly bent forward. The test involves an all-out gripping effort for 3 seconds, without movement of the arm. The average of three consecutive trials will be used as the measure of strength.

2.2.5. Physical activity questionnaire The participants will be asked to complete the Aerobic Center Longitudinal Study

Physical Activity Questionnaire. This Questionnaire is designed to conduct physical activity pattern, including leisure and household, during the past 3 months. Information regarding walking, stair climbing, jogging, swimming, weightlifting, aerobic dance, and household activities are acquired (27).

2.2.6. Body composition Three-site skinfold method will be used to calculate body percents of fat and fat-free masses. The recommended body sites for this method are chest, abdominal and thigh.

Additionally dominant and non-dominant forearm skinfolds will be obtained at the upper

1/3 of the forearm length (43).

2.2.7. Computer assessment The Biopac system and Acknowledge computer software will be utilized to capture and digitize vascular function indices received from the plethysmograph. The data will be collected at 500 Hz. then resampled at 4 Hz.

2.3. DATA REDUCTION AND ANALYSIS

Resting Binf will be recorded at a paper speed of 5 mm/sec. and values will be derived from the slope drawn at a best-fit tangent using the first 2-3 pulses. Calculations will be made as a function of 60 seconds divided by the horizontal distance (mm) needed for the

304 slope to rise vertically from baseline to the top of the recording paper and multiplied by the full chart range (15, 19). Forearm Binf following OCC will be recorded at a paper speed of 25 mm/sec. Analyses will be performed using a slope drawn at a best-fit tangent to the curves of the first 3-pulse flows post cuff release. Blood flows will then be calculated from 60 seconds multiplied by the paper speed divided by the horizontal distance (mm) needed for the volume slope to increase by 20 mm vertically (2, 3).

Forearm VR will then be calculated by dividing mean arterial pressure by the Binf.

Forearm Vcap will calculated as the vertical distance (mm) representing the change in forearm-volume graph following 6 minutes of venous filling (15). Forearm Vout will be derived from a tangent line representing the vertical drop in the volume-graph from the excursion line and drawn at 0.5 second and 2 second following the release of the venous occlusion pressure (15). All vascular indices will be expressed in ml•100ml tissue-1•min-

1.

2.4 STATISTICAL ANALYSIS

All statistical analysis will be completed using the SPSS statistical program (version

11.00). To examine the influence of 4-weeks of handgrip exercise training on arterial

reactivity (Hypothesis I) a 2 (Arms) * 2 (Pre and Post training) ANOVA will be used. To examine the influence of 4-weeks of handgrip exercise training on venous capacitance and venous outflow at rest and following 5-minutes of arm occlusion (Hypothesis II) a

second 2 (Arms) * 2 (Pre and Post training) ANOVA will be used. Depending on the

outcome of the 2 * 2 ANOVA Hypothesis III will be examined with a 1 (trained arm) * 5

(visits) repeated measure ANOVA to examine the time-course of the arterial and venous

changes. Finally a paired t-test will be used to compare between pre and post training

305 measures of HRV indices, body composition, and physical activity questioner. Alpha will

be set a-priori al p<0.05

REFERENCES

1.Allen J. Time course of vascular adaptations to short-term handgrip exercise training. Med Sci Sports Exerc, In Press.

2.Alomari M, Reyes R, Wood R and Welsch M. Short-term reliability of forearm venous indices using strain gauge plethysmography: influence of handgrip strength on venous outflow and capacitance. SEACSM, Atlanta, Ga, 2003.

3.Alomari MA, Welsch MA, Prisby RD, Lee CM and Wood RH. Modification of forearm vascular function following short-term handgrip exercise training. Int J Sports Med 22: 361-365., 2001.

4.Altenkirch HU, Fransson L and Koch G. Assessment of arterial and venous circulation in upper and lower extremities by venous occlusion strain gauge plethysmography. Normal values and reproducibility. Vasa 18: 140-145, 1989.

5.Altenkirch HU, Koch G and Koralewski HE. Variability and reproducibility of arterial and venous circulation parameters in the forearm and calf measured at one-week intervals. Vasa 19: 21-25, 1990.

6.Balady GJ, Weiner DA, Rose L and Ryan TJ. Physiologic responses to arm ergometry exercise relative to age and gender. J Am Coll Cardiol 16: 130-135, 1990.

7.Bank AJ, Shammas RA, Mullen K and Chuang PP. Effects of short-term forearm exercise training on resistance vessel endothelial function in normal subjects and patients with heart failure. J Card Fail 4: 193-201., 1998.

8.Barclay JK. A delivery-independent blood flow effect on skeletal muscle fatigue. J Appl Physiol 61: 1084-1090, 1986.

9.Berntson GG, Bigger JT, Jr., Eckberg DL, Grossman P, Kaufmann PG, Malik M, Nagaraja HN, Porges SW, Saul JP, Stone PH and van der Molen MW. Heart rate variability: origins, methods, and interpretive caveats. Psychophysiology 34: 623-648, 1997.

10.Chiba M, Nakamura M, Kanaya Y, Kobayashi N, Ueshima K, Kawazoe K and Hiramori K. Improvement in lower limb vasodilatory reserve and exercise capacity in patients with chronic heart failure due to valvular heart disease. Eur Heart J 18: 1931- 1936., 1997.

306 11.Clausen JP. Effect of physical training on cardiovascular adjustments to exercise in man. Physiol Rev 57: 779-815, 1977.

12.Convertino VA, Mack GW and Nadel ER. Elevated central venous pressure: a consequence of exercise training-induced hypervolemia? Am J Physiol 260: R273-277, 1991.

13.Copeland SR, Mills MC, Lerner JL, Crizer MF, Thompson CW and Sullivan JM. Hemodynamic effects of aerobic vs resistance exercise. J Hum Hypertens 10: 747-753., 1996.

14.Cramer M, Beach KW and Strandness Jr DE. The detection of proximal deep vein thrombosis by stain gauge plethysmography through the use of an outflow/capacitance discriminant line. Bruit VII, 1983.

15.Cramer M, Langlois Y, Beach K, Martin D and Strandness Jr DE. Standardization of venous flow measurements by strain gauge plethysmography in normal subjects. Bruit VII: 33-39, 1983.

16.Goodman JM, Pallandi DV, Reading JR, Plyley MJ, Liu PP and Kavanagh T. Central and peripheral adaptations after 12 weeks of exercise training in post-coronary artery bypass surgery patients. J Cardiopulm Rehabil 19: 144-150., 1999.

17.Green DJ, Cable NT, Fox C, Rankin JM and Taylor RR. Modification of forearm resistance vessels by exercise training in young men. J Appl Physiol 77: 1829-1833, 1994.

18.Guyton AC, Richardson TQ and Langston JB. Regulation of cardiac output and venous return. Clin Anesth 3: 1-34, 1964.

19.Hokanson DE, Sumner DS and Strandness Jr. DE. An electrically calibrated plethysmograph for direct measurement of limb blood flow. IEEE transactions on bio- medical engineering 22: 21-25, 1975.

20.Ikenouchi H, Iizuka M, Sato H, Momomura S, Serizawa T and Sugimoto T. Forearm venous distensibility in relation to severity of symptoms and hemodynamic data in patients with congestive heart failure. Japanese heart journal 32: 17-34, 1991.

21.Kenney WL and Armstrong CG. The effect of aerobic conditioning on venous pooling in the foot. Med Sci Sports Exerc 19: 474-479, 1987.

22.Louisy F and Guezennec CY. Hemodynamics of leg veins during a 30 days bed rest-- effect of lower body negative pressure (LBNP). Physiologist 32: S86-87., 1989.

307 23.Louisy F JJ, Guezennec CY. Filling and emptying characteristics of lower limb venous network in athletes. Study by postural plethysmography. International journal of sports medicine 18, 1997.

24.Mahankali B, Allen J, Mann A, Graves A and Welsch MA. Reproducibility of forearm blood flow using mercury-in silastic strain gauge plethysmography. Medicine and science in sports and exercise 29: S258(abstract), 1997.

25.Milani J, Fernhall B and Manfredi T. Estimating oxygen consumption during treadmill and arm ergometry activity in males with coronary artery disease. J Cardiopulm Rehabil 16: 394-401, 1996.

26.Miyachi M TH, Yamamoto K, Yoshioka A, Takahashi K, Onodera S. Effects of one- legged endurance training on femoral arterial and venous size in healthy humans. Journal of applied physiology 90: 2439-2444, 2001.

27.Pereira MA, FitzerGerald SJ, Gregg EW, Joswiak ML, Ryan WJ, Suminski RR, Utter AC and Zmuda JM. A collection of Physical Activity Questionnaires for health-related research. Med Sci Sports Exerc 29: S1-205., 1997.

28.Roberts DH, Tsao Y and Breckenridge AM. The reproducibility of limb blood flow measurements in human volunteers at rest and after exercise by using mercury-in-Silastic strain gauge plethysmography under standardized conditions. Clin Sci (Lond) 70: 635- 638, 1986.

29.Rowell LB. Human cardiovascular adjustments to exercise and thermal stress. Physiol Rev 54: 75-159, 1974.

30.Saito M, Matsui H and Miyamura M. Effects of physical training on the calf and thigh blood flows. Jpn J Physiol 30: 955-959, 1980.

31.Saltin B. Physiological adaptation to physical conditioning. Old problems revisited. Acta Med Scand Suppl 711: 11-24, 1986.

32.Sato H, Ikenouchi H, Aoyagi T, Matsui H, Mochizuki T, Momomura S, Serizawa T, Iizuka M and Sugimoto T. Pathophysiology and evaluation of severity of congestive heart failure on the basis of venous characteristics. Japanese circulation journal 53: 141- 145, 1989.

33.Seleye H. The Stress of Life. New York: McGraw-Hill, 1976.

34.Sinoway LI, Musch TI, Minotti JR and Zelis R. Enhanced maximal metabolic vasodilatation in the dominant forearms of tennis players. J Appl Physiol 6: 673-678, 1986.

308 35.Sinoway LI, Shenberger J, Wilson J, McLaughlin D, Musch T and Zelis R. A 30-day forearm work protocol increases maximal forearm blood flow. J Appl Physiol 62: 1063- 1067, 1987.

36.Solomito AN, Marsh L, Alomari M and Welsch MA. Technical considerations for the assessment of venous function: influence of occlusion time and pressure. SEACSM, Atlanta, GA, 2003.

37.Suzuki J, Gao M, Batra S and Koyama T. Effects of treadmill training on the arteriolar and venular portions of capillary in soleus muscle of young and middle-aged rats. Acta Physiol Scand 159: 113-121, 1997.

38.Suzuki J, Kobayashi T, Uruma T and Koyama T. Time-course changes in arteriolar and venular portions of capillary in young treadmill-trained rats. Acta Physiol Scand 171: 77-86, 2001.

39.Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology. Heart rate variability: standards of measurement, physiological interpretation and clinical use. Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology. Circulation 93: 1043-1065., 1996.

40.Tschakovsky ME, Shoemaker JK and Hughson RL. Vasodilation and muscle pump contribution to immediate exercise hyperemia. Am J Physiol 271: H1697-1701, 1996.

41.Wecht JM, de Meersman RE, Weir JP, Bauman WA and Grimm DR. Effects of autonomic disruption and inactivity on venous vascular function. Am J Physiol Heart Circ Physiol 278: H515-520, 2000.

42.Welsch MA, Alomari M, Parish TR, Wood RH and Kalb D. Influence of venous function on exercise tolerance in chronic heart failure. J Cardiopulm Rehabil 22: 321- 326., 2002.

43.Williams DP, Going SB, Milliken LA, Hall MC and Lohman TG. Practical techniques for assessing body composition in middle-aged and older adults. Med Sci Sports Exerc 27: 776-783, 1995.

309 APPENDIX C: LETTERS OF PERMISSIONS

310 Sent by: [email protected] To: [email protected] cc: Subject: RE: Urgent_Permission request_JCR, dissertation Permission is granted to reproduce the requested material for use in your academic thesis/dissertation. Permission is granted provided a prominent credit line is placed stating the original source and copyright owner, Lippincott Williams & Wilkins.

Sincerely,

Arnetta G. Queen Permissions Editor

-----Original Message----- From: O'Brien, David Sent: Wednesday, September 03, 2003 1:34 PM To: Queen, Arnetta Subject: FW: Urgent_Permission request_JCR, dissertation Importance: High

-----Original Message----- From: Mahmoud A Alomari [mailto:[email protected]] Sent: Tuesday, September 02, 2003 5:00 PM To: [email protected]; [email protected] Subject: Urgent_Permission request_JCR

Dear Mr. O?Brien I am completing a doctoral dissertation at Louisiana State University entitled " ACUTE AND CHRONIC VASCULAR RESPONSES TO EXERCISE TRAINING ". Since I was a co-author, I would like your permission to reprint in my dissertation following article:

Welsch MA, Alomari M, Parish TR, Wood RH and Kalb D. Influence of venous function on exercise tolerance in chronic heart failure. J Cardiopulm Rehabil 22: 321-326., 2002.

The requested permission extends to any future revisions and editions of my dissertation, including non-exclusive world rights in all languages, and to the prospective publication of my dissertation by UMI Company. These rights will in no way restrict republication of the material in any other form by you or by others authorized by you. Your signing of this letter will also confirm that you own the copyright to the above-described material.

If these arrangements meet with your approval, please sign this letter and email me the permission for the use of this article.

311

Since the permission has to be included in my dissertation, I would prefer an electronic copy of the permission however a fax would also serve the purpose. I have been trying to communicate with your company regarding this permission, either by email or using your online permission form, since August/28. However I havnt received any communication and your online permission form has not been working. Therefore, a quick response would be highly appreciated ?.. I have to meet the graduate school deadlines (September/3/2003) and they need to review the dissertation before I can submit it. Thank you very much.

Sincerely,

Mahmoud Alomari,

PERMISSION GRANTED FOR THE USE REQUESTED ABOVE:

Signature:

Name:

Date: ______

312 Sent by: "[email protected]"

Subj: Reprint permission Date: 8/29/2003 2:05:45 AM Pacific Daylight Time From: [email protected] To: [email protected] Sent from the Internet (Details)

Dear Mr. Alomari,

PERMISSION GRANTED provided that full credit is given to the original publication. ACUTE AND CHRONIC VASCULAR RESPONSES TO EXERCISE TRAINING, Int J Sports Med 22: 361-365, 2001. GEORG THIEME VERLAG.

Kind regards Georg Thieme Publishers

Margaretha Schoo Team 1

Ignat Mayzeles, Manager International Reprint Corporation (IRC) 287 East H Street Benicia, CA 94510, US Phone: (707) 746-8740 Fax: (707) 746-1643

313

314 VITA

Mahmoud Alomari was born in 1968 in Habaka, a little village of Irbid in Jordan. He received his elementary through high school education in United Arab Emirates until

1986. Subsequently Mahmoud attended the Yarmouk University, at his native Irbid, and received a Bachelor of Science in physical education in 1990. In an effort to expand his horizon, he moved to the United States to attend the Minnesota (previously named

Mankato) State University, Mankato, Minnesota, and finished a Master of Science degree in exercise physiology/cardiac rehabilitation in 1995. While pursuing his masters’, he received the International Student Endowment Scholarship for two consecutive years and served as teacher assistant for one year.

Mahmoud started pursuing a doctorate in exercise physiology in 1997 at the

Louisiana State University, Baton Rouge, Louisiana, and finished in 2003. During his tenure he served as research/teacher assistant at the Department of Kinesiology for four years and research associate at the Baton Rouge General Hospital for two years.

Additionally, in 2003 he was recipient of the Corbett Graduate Summer Research Award from the Department of Kinesiology and the Southeast American Collage of Sports

Medicine Student Research Award. Mahmoud will assume his responsibilities as a

faculty member in the Department of Allied Health Science at the Jordan University of

Science and Technology, Irbid, Jordan, starting October 2003.

315