.: I! 3.1! I x it . . 3: :xi a‘lL. .31 a . N11. 31:: t‘fimmfii 2,... a3; .3: 7.: .fi . a. “any in...” a 5.14.4 . II...- {t‘fivIl :13: 93 u g r .5233 r. z... 6.1%!» . K, 2 . .. ‘ ti; .5. , . 31%;. g .. l... .3123: v $1.: . r,.tVl“'>’¢, .s .5 ‘ . .71. L... n. .5. .. 3.1.7.. 1 p «339 .9 .I 1 A ‘ A .L :55: .3» .32.“... _ sot affifizhfim f a 9!. . 1331. lerb§f5fflba .. u“ .! . . xiii‘ 1v}. h~rusts4 42‘: .14 i‘il. 5&3. 1.131% xix-i : ‘ ;|.. 2.}.(45: . Kraft!!! r H. let iii. a... 41.. .(52 5‘ . 12.215} .. “3.9:... , v9 . I 1. £3. a . L... A: i ,. :1. , LIBRARY w .4 Michigan State University This is to certify that the dissertation entitled FUNCTIONAL DIFFERENCES IN SYMPATHETIC NEURAL CONTROL MECHANISM OF ARTERIES AND VEINS AS PROBED BY CONTINUOUS AMPEROMETRY AND VIDEO MICROSCOPY presented by Jinwoo Park has been accepted towards fulfillment of the requirements for the Doctoral degree in Chemistry as MM; Major Professor’s Siénature Date MSU is an Alfinnative Action/Equal Opportunity Institution .'A ~.... -o-I-O-O-I-O-C-0-0-Q-C-o--a-a-O-I-C-I-u—O-O‘l— PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE IIAY 0 3 2003 F3 2T08 un' 2/05 p:/ClRC/DateDue.indd-p.1 FUNCTIONAL DIFFERENCES IN SYMPATHETIC NEURAL CONTROL MECHANISM OF ARTERIES AND VEINS AS PROBED BY CONTINUOUS AMPEROMETRY AND VIDEO MICROSCOPY By J inwoo Park A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2006 ABSTRACT FUNCTIONAL DIFFERENCES IN SYMPATHETIC NEURAL CONTROL MECHANISM OF ARTERIES AND VEINS AS PROBED BY CONTINUOUS AMPEROMETRY AND VIDEO MICROSCOPY By J inwoo Park Numerous studies have reported that norepinephrine (NE) levels in the plasma of hypertensive humans and animals are elevated due to increased sympathetic nerve activity. Little is known, however, about how the kinetics and mechanism of NE release from perivascular sympathetic nerves as well as the postjunctional response is altered in hypertension. The first goal of my research was to design and implement continuous amperometric monitoring method and video imaging to record real time NE overflow from sympathetic nerves innervating arteries and veins and the evoked contractile response of isolated mesenteric veins (MV) and arteries (MA). The research included fabrication and characterization of a new microelectrode material, boron-doped diamond. The second goal was to use these two techniques to probe the functional differences in sympathetic neural control of MA and MV, and how the control mechanisms are altered in salt-sensitive hypertension. The deoxycorticosterone acetate (DOCA)-salt animal model for salt sensitive hypertension was used. The results showed that the diamond microelectrode provided superior response performance as compared to a carbon fiber in terms of response precision and stability in tissue. Diamond exhibited resistance to fouling, least in part, because of the non-polar, low-oxygen, sp3-bonded carbon surface on which weak adsorption of polar biomolecules and other contaminants occurs. The results also confirmed that first there are fundamental differences in sympathetic neural control of MA and MV. These differences are: (i) the density and arrangement of sympathetic nerves in MA and MV, with MA having a higher density than MV, (ii) electrical stimulation-evoked NE overflow was exceeded in MV than MA in sham rats because NE release and clearance in MA were more strongly regulated by prejunctional aZ-adrenergic autoreceptor and norepinephrine transporter (NET), respectively compared with MV, iii) frequency-response vasoconstriction in MV was more sensitive than MA, iv) in small MA, ATP was the dominant neurotransmitter, whereas in small MV, NE was the transmitter mediating constriction. The neural control mechanism of MA and MV were altered in the DOCA-salt animal model for hypertension as compared to sham control. This was evidenced by these findings: (i) NE overflow increased in DOCA-salt rat MA compared to sham MA due to, at least in part, an altered function of the prejunctional a2-adrenergic autoreceptor. The elevated NE overflow caused an increase in NE uptake in DOCA-salt MA. (ii) ATP mediated the neurogenic constriction of MA from sham rats while NE mediated most neurogenic constriction in DOCA—salt MA. This may be due to an increased adrenergic component of neurogenic constriction in DOCA-salt MA. (iii) the maximum contractile response was decreased in DOCA-salt MV, though NE overflow was not different between sham and DOCA-salt MV. It may be due to the desensitization of Oil-adrenergic receptors or/and due to the altered vasculature in DOCA-salt MV. Taken together, the differences in neuroeffector transmission may contribute to the different hemodynamic function of arteries and veins. Furthermore, the altered neuroeffector transmission in hypertension may contribute to the elevation in blood pressure. Cepyright by .IINWOO PARK 2006 Dedicated to my family for their everlasting support and encouragement. Thank you for being the wind beneath my wings. ACKNOWLEDGEMENTS First of all, I would appreciate and thanks sincerely to my advisor Dr. Greg M. Swain and co-advisor Dr. James J. Galligan for their patience, great support. I respect and appreciate their quest for excellence, vast knowledge in science. This dissertation is the final state of five years work on my part. However, without your advice and support, I would have not completed my study. It has been great honor working with you. I anticipate having a life-long professional relationship with you. I would also like to acknowledge my guidance committee Dr. Gregory D. Fink and Dr. McGuffin L. Victoria. I appreciate their time and willingness to serve on my graduate committee for the last five years. Thank you for all the comments concerning my research, and also for your kindness. In addition, I would like to thank my former MS. degree advisor Dr. Sang-Hak Lee at Kyungpook National University in Korea for his support and encouragement. Special thank to Dr. Show for his advices on growing diamond-films. I owe many thanks to my colleagues in Dr. Swain lab (Dr. Jian Wang, Dr. Malgorzata Witek, Dr. Mateuz Hupert, Dr. Shannon Haymond, Dr. Josef Cvacka, Suzana Cvackova, Dr. Prema Sonthalia, Dr. Jason Stottor, Dr. Grace Muna, Dr. Veronika Mocko, Dr. Karolina Peckova, Dr. Martin Novotny, Dr. Bhavik Patel, Dr. Aihua Liu, Dr. lsao Sasaki, Dr. Jason Bennett, Dr. Ann Ficher, Audrey Martin, Dough Knigge, Yang Song, Shihua Wang, Elizabeth McGaw, Hua Dong, Yingrui Dai, Veron Matt, Ayten Ay, Michael Lowe, Jason Thornton, Luther Schaeffer) and Dr. Galligan’s lab (Dr. Xiaochun vi Bian, Dr. Xiaoling Dai, Dr. Alex Perez-Rivera, Dr. Jim Rain, Dr. Hui Xu, Hui Wang, Sandra Hlavacova, Vino Naidoo, Dima Alkawwas, Stacie Demel). Tank you all for your great support. Finally, but the most important in my life, I deeply thank my parents, sisters and brothers in laws for their endless love, sacrifices and support. I am grateful for their encouragement to do the best work possible through out my life. So many things happened to my life past five years while at MSU but meeting my wife, Jiwon Son was greatest thing that happened to me. I sincerely appreciate her sacrifices and encouragement to overcome all the problems. Thank you very much for being my soul mate, my confidant, my lover and my best friend. I dedicate this dissertation to my family. vii Table of Contents List of Figures x List of Tables xvr Chapter 1 1 Introduction 1 1.1 Overall Specific Aims ............................................................................................... 1 1.2 Hypertension ............................................................................................................. 3 1.3 Blood Vessel ............................................................................................................. 5 1.4 Sympathetic Nervous System (SN S) ........................................................................ 8 1.5 Sympathetic Neuroeffector Transmission in Arteries and Veins ............................ 12 1.6 Experimental Hypertension Animal Models ........................................................... 18 1.7 Hyperactivity of the SNS in Hypertension .............................................................. 20 1.8 Increased NE Release in Hypertension ................................................................... 21 1.9 Assessment of SN S Activity ................................................................................... 24 1.10 Microelectrode ...................................................................................................... 28 1.11 Carbon Fiber Microelectrode ................................................................................ 30 1.12 Properties of Diamond and Its Application ........................................................... 33 1.13 Boron-Doped Diamond Electrodes ....................................................................... 35 1.14 Outline ................................................................................................................... 36 Chapter 2 ...... 39 Experimental Section 39 2.1 Boron-Doped Diamond Film Growth ..................................................................... 39 2.2 Diamond Film Characterization .............................................................................. 41 2.3 Microelectrode Preparation ..................................................................................... 4-4 2.4 Electrochemical Measurement ................................................................................ 46 2.5 Animals ................................................................................................................... 52 2.6 In Vitro Electrochemical and Diameter Measurement System ............................... 53 2.7 Fluorescence Histochemistry .................................................................................. 57 2.8 Chemical and Drug Application .............................................................................. 57 2.9 Data Analysis .......................................................................................................... 58 Chapter 359 Comparison of Electrochemical Properties of Diamond and Carbon Fiber Microelectrodes During Exposure to Biological Environments...59 viii 3.1 Introduction ............................................................................................................. 59 3.2 Results and Discussion ............................................................................................ 61 3.3 Conclusion ............................................................................................................... 81 Chapter 4 ...................................................................................................... 84 Endogenous Norepinephrine and Its Effect on the Contractile Response of a Rat Mesenteric Artery Using Continuous Amperometry and Video Microscopy ................................................................................................... 84 4.1 Introduction ............................................................................................................. 84 4.2 Results and Discussion ............................................................................................ 87 4.3 Conclusion ............................................................................................................. 107 Chapter 5 .................................................................................................... 109 Differences in Sympathetic Neuroeffector Transmission to Rat Mesenteric Arteries and Veins as Probed by In Vitro Continuous Amperometry and Video Imaging .. ............. ............. ......... ...109 5.1 Introduction ........................................................................................................... 109 5.2 Results ................................................................................................................... 110 5.3 Discussion ............................................................................................................. 128 5.4 Conclusion ............................................................................................................. 135 Chapter 6 .................................................................................................... 136 Alterations in Sympathetic Neuroeffector Transmission to Mesenteric Arteries and Veins in DOCA-salt Hypertensive Rats ............................ 136 6.1 Introduction ........................................................................................................... 136 6.2 Results ................................................................................................................... 138 6.3 Discussion ............................................................................................................. 154 6.4 Conclusion ............................................................................................................. 160 Chapter 7 .................................................................................................... 161 Conclusions ................................................................................................ 161 References .................................................................................................. 166 ix List of Figures Figure 1.1. A part of cross section vessel of blood vessel ................................................. 5 Figure 1.2. Sympathetic divisions of the autonomic nervous system ................................. 9 Figure 1.3. Sympathetic nerves on a blood ....................................................................... 10 Figure 1.4. Sympathetic neuroeffector transmission to blood vessels. P: purinergic receptor, or: adrenergic receptor, ATP: adenosine 5’-triphosphate, and NE: norepinephrine ........................................................................................................................................... 12 Figure 1.5. Major steps in norepinephrine synthesis ......................................................... 15 Figure 2.1. SEM images of (A) an electrochemically sharpened Pt wire (top) and 3 Pt wire coated with a polycrystalline boron-doped diamond film (bottom). (B) An expanded view of the end of the diamond-coated Pt wire ................................................................. 41 Figure 2.2. Raman spectra for diamond thin film deposited on Pt wire with different methane-to-hydrogen ratios ............................................................................................... 43 Figure 2.3. (A) Diagram of the conically shaped diamond microelectrode insulated with polypropylene. SEM images of the polypropylene-insulated diamond microelectrode at (B) lower and (C) higher magnification. Top-view SEM images of the diamond film morphology without (D) and with (E) the polypropylene insulation layer ....................... 45 Figure 2.4. Background cyclic voltammetric i—E curves in (A) 1.0 M HC104 for a diamond (—), carbon fiber (— —) and Pt-exposed diamond microelectrode (----). Cyclic voltammetric i —E curves for a diamond and carbon fiber microelectrode in (B) 1.0 mM Fe(CN) 6'3“ in 1 M KCl. Scan rate=100 mV/s ................................................................ 50 Figure 2.5. Block diagram of the experimental set-up ...................................................... 54 Figure 3.1. Background cyclic voltammetric i—E curves for a (A) diamond and (B) carbon fiber microelectrode in 0.1 M phosphate buffer solutions of different pH ranging from 3 to 7.2. Scan rate = 100 mV/s. ............................................................................................ 62 Figure 3.2. Cyclic voltammetric i—E curves for a diamond and carbon fiber microelectrode in (A) 10 M NE and (B) 10 M EE, both in 0.1 M phosphate buffer, pH 7.2. Scan rate = 100 mV/s. ................................................................................................ 65 Figure 3.3. Cyclic voltammetric i-E curves for 1 mM Fe(CN)6’3"4 in 0.1 M KCl at (A) diamond and (B) carbon fiber microelectrode before and after laboratory atmosphere exposure. Scan rate = 100 mV/s. ....................................................................................... 68 Figure 3.4. Cyclic voltammetric i—E curves for 1 mM Fe(CN)(,’3/'4 in 0.1 M KC1 at a (A) diamond and (B) carbon fiber microelectrode before and after exposure to biological tissue. Scan rate = 100 mV/s. ............................................................................................ 73 Figure 3.5. Continuous amperometric i—t responses for a (A) diamond and (B) carbon fiber microelectrode during multiple injections of 0.1 uM NE in Krebs’ buffer, pH 7.4 before and afier exposure to tissue. The measurements were made in the flow bath. Injection volume = 3 mL. Flow rate = 1.6 mL/min. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCI. (C) Plot of the nominal current response as a function of time for the diamond and carbon microelectrodes. .................. 75 Figure 3.6. In Vitro continuous amperometric i—t responses for a diamond microelectrode during measurement of NE release from the mesenteric artery of a laboratory rat. The electrode response at (A) 0 V after electrical stimulation of the artery and (B) +0.8 V after electrical stimulation of the artery. The electrode response at +0.8 V after multiple electrical stimulation events with a time interval between each of (C) 20 s and (d) 10 min. The blood vessels were stimulated with the same pulse number and frequency (60 pulses at 20 Hz). Flow rate = 1.6 mL/min. ................................................................................ 778 Figure 3.7. In vitro continuous amperometric i—t curves for a (A) diamond and (B) carbon fiber microelectrode during NE release from a mesenteric artery as a function of time. The NE-release was evoked by electrical stimulation consisting of 60 pulses at 20 Hz. (C) Plots of the nominal current response as a function of time for the diamond and carbon fiber microelectrodes. Flow rate = 1.6 mL/min. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCl. ........................................................................... 80 Figure 3.8. Dose—response curves for NE at (A, C) diamond and (B, D) carbon fiber microelectrodes. Measurements with 0.01—1.0 uM NE were made. Flow rate = 1.6 mL/min. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCl. ........................................................................................................................................... 82 Figure 4.1. Video micrographs showing a mesenteric artery (A) before and (B) after a 20 Hz electrical stimulation (60 pulses with a 0.3 ms pulse width). The bipolar stimulator electrode and the diamond microelectrode, both positioned at the artery surface, are evident in (A). The NE oxidation reaction mechanism is shown in (C). Characteristic contractile (top) and NE oxidation current (bottom) response transients in response to a 20 Hz stimulation are presented in (D). Detection potential = 800 mV. Flow rate = 1.6 mL/min. Several numerical parameters are obtained from these curves including: the maximum oxidation current, 1...“; the current rise time, T,; the time required for current decay to 50% of the maximum, T50; the time required for full current decay, To; the percent constriction, Co/,; the time required to reach full constriction, Tc; the time required for full vessel relaxation, TR; and the latency period prior to the onset of constriction, TL. ........................................................................................................................................... 88 Figure 4.2. (A) An in vitro hydrodynamic voltammetric i-E curve recorded for NE released from the sympathetic nerves innervating a mesenteric artery. Contractile (top) xi and NE oxidation current (bottom) response transients for a mesenteric artery in response to a 20 Hz stimulation (60 pulses and a 0.3 ms pulse width) at detection potentials of (B) 0 and (C) 800 mV. The oxidation current was measured with a diamond microelectrode. Flow rate =1.6 mL/min of Krebs' buffer. ......................................................................... 92 Figure 4.3. (A) Typical contractile and NE oxidation current response transients for a mesenteric artery in response to a 20 Hz stimulation. Contractile and NE oxidation current response transients (B) in the presence of 'ITX (0.3 uM) and (C) after washing out the drug. The dotted circles indicate the time the electrical stimulation was applied. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs’ buffer. ....................................................................... 94 Figure 4.4. Effect of yohimbine (1.0 uM) on the contractile (top) and NE oxidation current (bottom) response transients recorded at a mesenteric artery. The transients were elicited by electrical stimulation at 20 and 3 Hz (60 pulses and a 0.3 ms pulse width) (A) without and (B) with added yohimbine. The dotted circles indicate the time the 3 Hz electrical stimulation was applied. Plots of the oxidation current for NE, with and without (i.e., control) added yohimbine, versus the stimulation frequency are shown in (C). Plots of the contractile response, with and without added yohimbine, versus the stimulation frequency are shown in (D). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs’ buffer 96 Figure 4.5. Effect of cocaine (10 uM) on the NE oxidation current (lefi) and contractile (right) response transients recorded at a mesenteric artery. The transients were elicited by electrical stimulation at (A) 20 and (B) 3 Hz (60 pulse and a 0.3 ms pulse width). Plots of the oxidation current for NE, with and without (i.e., control) added cocaine, versus the stimulation frequency are shown in (C). Plots of the contractile response, with and without added cocaine, versus the stimulation frequency are shown in (D). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs' buffer. ............................................................................... 100 Figure 4.6. Effect of pulse frequency (3, 5, 7, 10 and 20 Hz, all at 60 pulses with a 0.3 msec pulse width) on the contractile (top) and NE oxidation current (bottom) response transients recorded at a mesenteric artery (A). Plots of the NE oxidation current and contractile response versus the stimulation frequency are shown in (B). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs' buffer. ............................................................................... 103 Figure 4.7. Effect of pulse number (12, 30, 60, 90 and 120, all at 20 Hz) on the contractile (top) and NE oxidation current (bottom) response transients recorded at a mesenteric artery (A). Plots of the NE oxidation current and contractile response versus the stimulation frequency are shown in (B). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a xii diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs' buffer. .. ........................................................................................................................................ .105 Figure 5.1. Glyoxylic acid-induced fluorescence of catecholamines in peri-vascular sympathetic nerves. A and B show low magnification (100 X) images of the sympathetic nerve plexus while C and D show high magnification images (400 X). The nerve plexus in arteries has a mesh-like arrangement and nerve fibers are oriented in both the longitudinal and circular axis of the blood vessel (C). The plexus in veins has a selective circular arrangement with very few fibers oriented in the long-axis of the blood vessel (D). ................................................................................................................................. 111 Figure 5.2. Contractile (top) and NE oxidation current (bottom) response transients for (A) MA and (B) MV elicited by a 3 and 20 Hz stimulation. Plots of the (C) NE overflow current and (D) constriction for MA and MV as a function of the stimulation frequency. *Significantly different from MA and MV (P < 0.05). Data are mean i S.E.M. .......... .114 Figure 5.3. Calibration of a carbon fiber electrode for quantitative recording of NE overflow. (A) Oxidation currents produced by injection of standard NE solution. (B) Calibration plot showing how the NE oxidation current changes with concentration. Five different electrodes were used to record the data set. Data are mean 3: SEM. ............. 116 Figure 5.4. Contribution of al-adrenergic and P2 receptors to neurogenic constrictions of MA and MV. (A) Effects of PPADS (10 uM) on neurogenic constriction (top) and NE oxidation current caused by a 20 Hz stimulus train in a MA. PPADS blocked the constriction but not the NE oxidation current. (B) PPADS did not alter the neurogenic constriction or NE oxidation current in a MV. (C) Concentration-constriction curve for neurally-released NE in MA in the absence and presence of PPADS and prazosin (0.1 uM). Peak NE levels were measured during stimulus trains (1, 3, 7, 10 and 20 Hz) by converting oxidation currents to NE levels using the calibration curve shown in Figure SB. (D) Experiments similar to those shown in C except these studies were done in MV during stimulus trains (0.5, l, 3, 7, 10 and 20 Hz). ........................................................ 117 Figure 5.5. Contribution of a2-adrenergic receptor-mediated components to NE overflow and neurogenic constrictions of sham mesenteric arteries (MA) and mesenteric veins (MV). F requency-response for NE oxidation current before (control) and after application of yohimbine (1.0 uM) and UK 14,304 (1.0 uM) in MA (A) and MV (B). Frequency- response curves for contractile response before (control) and after application of yohimbine and UK 14,304 in MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control MA (MV) and in yohimbine MA (MV). #Indicates significantly different from NE oxidation current in control MA (MV) and in UK 14,304 MA (MV) (P<0.05). Data are mean :t S.E.M. ............................................................... 119 Figure 5.6. Contribution of the NE reuptake to NE clearance and neurogenic constriction of mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of cocaine (10 uM) and cocaine + yohimbine in MA (A) and MV (B). Frequency-response curves for contractile response xiii before (control) and after application of cocaine (10 M) and cocaine + yohimbine in MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control MA (MV) and in cocaine treated MA (MV) (P<0.05). “Indicates significantly different from NE oxidation current in control MA (MV) and in cocaine + yohimbine treated MA (MV) (P<0.05). Data are mean i S.E.M. .................................................... 124 Figure 6.1. Contractile (top) and NE oxidation current (bottom) response transients for (A) MA and (B) MV from sham and DOCA-salt rats in response to a 3 Hz stimulation. Comparison of the frequency response of (C) the NE oxidation current and (D) vasoconstriction in sham and DOCA-salt MA and MV. *Significant difference between sham and DOCA-salt MA (P < 0.05). &Significant difference between sham MA and sham MV (P < 0.05). Data are presented as mean :1: SEM. .......................................... 140 Figure 6.2. Contribution of aZ-adrenergic receptor-mediated components to NE overflow and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of yohimbine (1.0 uM) in sham and DOCA-salt MA (A) and MV (B). F requency-response curves for contractile response before (control) and after application of yohimbine in sham and DOCA-salt MA (C) and MV (D). *indicates significantly different from NE oxidation current in control sham MA(MV) (P<0.05). l”indicates significantly different from NE oxidation current in control DOCA-salt MA (MV) (P<0.05). Data are presented as mean i S.E.M. ................................................... 142 Figure 6.3. Contribution of oLZ-adrenergic receptor-mediated components to NE overflow and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of UK 14,304 (1.0 uM) in sham and DOCA-salt MA (A) and MV (B). Frequency-response curves for contractile response before (control) and afier application of UK 14,304 in sham and DOCA-salt MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control sham MA (MV) (P<0.05). #Indicates significantly different from NE oxidation current in control DOCA-salt MA (MV) (P<0.05). &Indicates significantly different from NE oxidation current in the drug treated sham MA (P<0.05). Data are mean 1 SEM. .................................................... 145 Figure 6.4. Contribution of reuptake transporter-mediated components to NE clearance and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of cocaine (10 uM) in sham and DOCA-salt MA (A) and MV (B). F requency-response curves for the contractile response before (control) and after application of cocaine in sham and DOCA-salt MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control sham MA (P<0.05). ”Indicates significantly different fi'om NE oxidation current in control DOCA-salt MA (P<0.05). &Indicates significantly different from NE oxidation current in the drug treated sham MA (P<0.05) Data are mean i S.E.M. ................................................................. 147 xiv Figure 6.5. Contribution of uptake transporter/aZ-adrenergic receptor-mediated components to NE clearance/release and neurogenic constrictions of sham and DOCA- salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and afier application of combined drug (cocaine (10 uM) + yohimbine (1.0 uM)) in sham and DOCA-salt MA (A) and MV (B). Frequency- response curves for contractile response before (control) and afier application of the combined drug in sham and DOCA-salt MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control sham MA (MV) (P<0.05). ”Indicates significantly different from NE oxidation current in control DOCA-salt MA(MV) (P<0.05). Data are mean i S.E.M. ................................................................................. 150 Figure 6.6. Contribution of al-adrenergic receptor and P2X—mediated components to NE release and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of prazosin (0.1 uM) or/and PPADS (10 uM) in sham and DOCA-salt arteries (A) and veins (B). Frequency-response curves for contractile response before (control) and after application of the drug in sham and DOCA-salt MA (C) and MV (D). Data are mean i S.E.M. ...................................................................... 152 XV List of Tables Table 3.1. Summary of cyclic voltammetric Em data for diamond and carbon microelectrodes. ................................................................................................................ 66 Table 4.1. Numerical parameters measured from the oxidation current and contractile response transients recorded in the presence and absence of yohimbine (1.0 uM). ......... 98 Table 4.2. Numerical parameters measured from the oxidation current and contractile response transients recorded in the presence and absence of cocaine (10 uM). ............. 101 Table 5.1. Numerical parameters obtained from the NE overflow oxidation current and contractile response transinents recorded in the presence and absence (control) of yohimbine (1.0 uM). ...................................................................................................... 121 Table 5.2. Numerical parameters obtained from the NE overflow oxidation current and contractile response recorded in the absence (control) and presence of UK 14,304 (1.0 uM) treatment. ............................................................................................................... 123 Table 5.3. Numerical parameters obtained from the NE overflow oxidation current and contractile response transients recorded in the presence and absence (control) of cocaine (10 uM). ......................................................................................................................... 126 Table 5.4. Numerical parameters obtained from the NE overflow oxidation current and contractile response transients recorded in the presence and absence (control) of cocaine + yohimbine (Co +Yo). .................................................................................................. 127 Table 6.1. Maximum constriction (Em) and half maximum stimulation frequency. (S 50) for MA and MV from sham and DOCA-salt rats. All data are expressed as the mean i SEM and “11” value refers to the number of animals from which the data were obtained. ......................................................................................................................................... 141 Table 6.2. Maximum constriction (Em) and half maximum stimulation frequency (S50) for MA and MV from sham and DOCA-salt rats in the absence (control) and presence of yohimbine (Yo, 1.0 pM) and UK 14,304 (UK, 1.0 uM). All data are expressed as the mean i SEM and “n” values refer to the number of animals from which the data were obtained. .......................................................................................................................... 144 Table 6.3. Maximum constriction (Em) and half maximum stimulation frequency (S50) for MA and MV from sham and DOCA-salt rats in the absence (control) and presence of cocaine (Co, 10 uM) and the combined drug (cocaine + yohimbine). All data are expressed as the mean i SEM and “n” values refer to the number of animals from which the data were obtained ..................................................................................................... 148 Table 6.4. Maximum constriction (Em) and half maximum stimulation frequency (S50) in MA and MV from sham and DOCA-salt rat in the absence (control) and presence of xvi prazosin (0.1 uM) and PPADS (10 uM) treatment. All data are expressed as the mean i SEM and “n” values refer to the number of animals from which the data were obtained. ......................................................................................................................................... 153 xvii Chapter 1 Introduction 1.1 Overall Specific Aims Hypertension is a major risk factor for stroke and heart and kidney failure. According to recent estimates, nearly one in three US. adults has hypertension, but because there are no symptoms, many people are unaware that they have high blood pressure. Numerous studies have reported that elevated levels of norepinephrine (N E), a vasoconstrictor neurotransmitter in the sympathetic nervous system, and altered postjunctional reactivity of the smooth muscle cells (contractile response) exist in human hypertension and animal hypertensive models. These alterations are at least partly associated with dysfunction of the sympathetic nervous system (SNS). The SNS plays a key role in regulating blood pressure by controlling vascular tone. Previous studies suggested that SNS dysfunction is due to functional changes of the presynaptic a2- adrenergic autoreceptors and NE uptake transporter (NET). However, these results are still controversial and the dysfimction differs from tissue-to-tissue and animal-to-animal. Moreover, little is known about the mechanism of local NE release from perivascular sympathetic nerves and the evoked vasoconstriction of arteries and veins in hypertension. Techniques that enable local and real time monitoring of endogenous NE release from sympathetic nerve endings and vasoconstriction simultaneously are needed to provide more insight into neural control processes in hypertension. The research described herein was conducted with three specific aims: Specific aim 1 : Design and implement instrumentation for directly monitoring NE overflow and vasoconstriction using continuous amperometry and a diamond microelectrode along with video imaging. This phase of the work involved fabricating and characterizing diamond microelectrodes, evaluating the basic electrochemical properties of the material and demonstrating the beneficial properties for in vitro electrochemical measurements. A test station with these two tools was constructed to examine sympathetic nerve activity at arteries and veins from the mesentery of laboratory test animals. Specific aim 2 : Identify fundamental differences between sympathetic neuroeffector transmission in mesenteric arteries (MA) and veins (MV) from healthy rats. Using these two tools, studies were performed to learn about the ftmctional differences between NE release, reuptake and the elicited contractile response of arteries and veins from healthy animals. Using the oxidation current or charge recorded for endogenous NE along with the various vasoactive agents, knowledge was gained about factors controlling NE release and clearance at sympathetic neuroeffector junctions. Specific aim 3 : Identify hypertension-associated alteration in sympathetic neuroeffector transmission in MA and MV in the DOCA-salt animal model for hypertension. Deoxycorticosterone acetate (DOCA)-salt hypertension is a well- characterized model of experimental hypertension, which is associated with altered sympathetic nerve activity. In Vitro continuous arnperometry and video microscopy, along with various vasoactive substances, were used to learn what changes in the neural control mechanism of MA and MV occur in salt-sensitive hypertension in terms of the mechanisms of release, reuptake and effector cell physiological response. 1.2 Hypertension Hypertension is an increasingly important disease state because high blood pressure is a common finding in and is one of the most important modifiable risk factors for cardiovascular disease [1, 2]. Hypertension is defined as a sustained systolic and/or diastolic blood pressure greater than 140/90 mmHg. 50 million or more Americans have high blood pressure and, worldwide, estimates suggest as many as 1 billion individuals suffer from the disease. Approximately 7.1 million deaths worldwide per year are attributable to hypertension. In 90-95 % of these cases, the causation is unclear and this is referred to as “essential hypertension”. A number of physiological mechanisms are involved in the maintenance of normal blood pressure, and their derangement may play a part in the development of essential hypertension [3]. It is probable that a great many interrelated factors, such as obesity, salt intake, insulin resistance, genetics, low birth weight, endothelial dysfunction, altered renin-angiotensin and sympathetic nervous system (SNS) activity, contribute to the elevated blood pressure in hypertensive patients. The relative influence of these factors may differ between individuals. The mechanisms underlying hypertension development are poorly understood, so rather than curing the disease, current therapeutic approaches simply serve to lower blood pressure and hopefiilly prevent other cardiovascular diseases secondary to hypertension. Understanding the causation factors could lead to better treatments and even a cure for essential hypertension. In the remaining cases, high blood pressure results from a identifiable underlying renal, adrenal disease or endocrine disorders as the cause for the raised blood pressure [4]. This kind of hypertension is called “secondary hypertension” and can often be cured by correcting its original cause. Regulation of blood pressure. The mean arterial blood pressure is determined by a balance between cardiac output (CO) and total peripheral vascular resistance (TPR). Peripheral resistance is mainly dependent on small arteries and arterioles, whereas cardiac output depends on the volume of blood being pumped by the heart (heart rate) and cardiac contractility [5]. The roles of CO and TPR are still controversial but most patients and animals with established essential hypertension have a normal CO and elevated TPR [6, 4]. However, TPR in the initial stages of hypertension is not raised or reduced so the blood pressure elevation is caused by increased CD, which is related to increased venous tone and caused, at least in part by, sympathetic overactivity [7, 4]. The subsequent rise in peripheral arteriolar resistance might, therefore, develop in a compensatory manner to prevent the raised pressure by the increased CO. The homeostasis of blood pressure involves a complex interplay between the nervous, endocrine, renal and cardiovascular systems. As examples, the central nervous system controls blood pressure by adjusting the autonomic outflow to the cardiovascular system and the kidney is important for maintaining soditun and water balance and controlling the intravascular blood volume [8]. Various hormonal and vasoactive factors, such as angiotensin II, vasopressin, nitric oxide and endothelin, also act on the cardiovascular system to regulate blood pressure [9, 10]. The above systems interact with one another, forming a delicate blood pressure controlling mechanism. 1.3 Blood Vessel The blood vessels are the plumbing of the circulatory system and function to transport blood throughout the body; a process is that is essential for life. Blood flows to or from cells and organs delivering oxygen and nutrients and removing carbon dioxide and waste products. Blood flow also maintains the optimum pH and the mobility of the proteins and cells of the immune system. The most important vessels, arteries and veins, are so termed because they carry blood away from or towards the heart, respectively. The walls of arteries and veins are composed of three layers (Figure 1.1): the innermost intima layer, the media layer and the outermost adventitia layer. Elasticai tema "v Endothelial ' ' Intima , Media 1 Adventitia Figure 1.1. A part of cross section of blood vessel. The innermost layer, which is in direct contact with the flow of blood, is the tunica intima, commonly called the intima. The intima is made up of mainly a single layer of endothelial cells. The endothelial cells are capable of releasing vasodilators, such as nitric oxide and adenosine, and vasoconstrictors, such as endothelin, which influence vascular tone. Outside this layer is the tunica media or media, which is made up of smooth muscle cells and elastic tissue [1 1]. Blood flow is regulated by contraction of the smooth muscle cells, which alters the diameter and resistance of the small blood vessels. The number of smooth muscle layers varies from 2 to 10 among the vessels and is largely in proportion to blood vessel wall thickness [12]. The outermost layer is known as the tunica adventitia or the adventitia, which is the most variable layer and contains the nerves that innervate the smooth muscular cell layer. Arterial system. There are several types of arteries according to their fimction and size. The aorta is the root and largest systemic artery that carries blood away from the heart. Branches of the aorta are called arteries, such as the carotid and mesenteric arteries. Pulmonary arteries carry deoxygenated blood that has just returned from the body to the lung, where carbon dioxide is exchanged for oxygen. Smaller than arteries are the arterioles followed by the smallest vessels, the capillaries. The capillaries connect the arterial to the venous circulation. Systemic arteries deliver blood to the arterioles and then to the capillaries, where nutrients and gases are exchanged. Capillaries consist of little more than a layer of endothelium and occasionally connective tissue. The small arteries (150-250 um), arterioles (<100 um), and capillaries account for 90% of the vascular resistance. The larger arteries have a low resistance to blood flow and function primarily as conduits. However, as arteries approach an organ they divide into many small arteries both just outside and within the organ, these increasing the vascular resistance. Venous system. There are also several types of veins according to their function and size. A venule is a small blood vessel that allows deoxygenated blood to return from the capillary beds to the larger blood vessels called veins. The largest vein, venae cavae returns the deoxygenated blood away from the body into the heart. The pulmonary veins carry oxygen-rich blood from the lungs to the left atrium of the heart. They are the only veins in the postal-fetal human body that carry oxygenated blood. The portal vein drains blood from the digestive system and its associated glands, such as mesenteric and splanchnic veins, to the liver. Although the venous system has a similar structure to that of the arterial system, there is less smooth muscle and connective tissue. This makes the walls of veins thinner than those of arteries, which is related to the fact that blood in the veins is under less pressure than in the arteries [11]. In addition, veins have a larger cross- sectional area (lumen diameter) than their arteriole counterparts. Therefore, veins are up to 60 times more distensible than arterioles due to these structural differences and have i the capacity to store 3-4 times more blood than arterioles. Thus, veins and venules are often described as the capacitance section of the vascular bed and usually contain approximately 2/3 of the body’s total blood volume [13]. Although multiple factors regulate vascular capacitance, the structure and smooth muscle constrictor activity (venomotor tone) of veins entirely determines the capacitance. Hypertension is also associated with altered structural, mechanical and pathophysiological properties of resistance arteries and capacitance veins [14, 15]. Structural changes in vascular smooth muscle cells in hypertension consist of two forms: remodeling and growth. Remodeling is a rearrangement of normal sized cells whereas growth is either an increased smooth muscle cell mass due to increased protein synthesis (hypertrophy) or cell replication (hyperplasia) [16]. In essential hypertension, the increased media/lumen ratio of small arteries due to remodeling could contribute to increased vascular resistance [17]. Also, research with hypertension animal models has revealed altered vascular structure by hyperplasia or hypertrophy. These structural changes, like thickening of the vascular wall, include narrowing of the lumen and external diameters, and increases in media width and media cross-sectional area [18, 19, 15]. However, whether the altered vascular structure is a cause or a consequence of the increased blood pressure is unclear. 1.4 Sympathetic Nervous System (SNS) The autonomic nervous system regulates bodily functions and the activity of specific organs. One of the its functions is to regulate the cardiovascular system by controlling vascular tone and cardiac strength through feedback with the central nervous system [20]. Thus, the autonomic nervous system has an important role in maintaining a normal blood pressure [4]. The autonomic nervous system originates from the brain stem and is divided into two large subsystems: sympathetic and parasympathetic. The sympathetic and parasympathetic systems have contrasting function in regulating the internal environment [21]. The sympathetic nervous system (SNS) has classically been divided into pre- and postganglionic fibers. Preganglionic neurons of the sympathetic division originate in the thoracic (T1 to T12) and upper lumbar (L1 to L3) region of the spinal cord (Figure 1.2). Postganglionic sympathetic neurons consist of the axons of the ganglionic neurons and innervate the heart and resistance vessels help to control cardiac output, arterial blood pressure and regional vascular conductance, thus ensuring the proper perfiision of vital organs [22]. Moreover, SNS is the primary mediator of acute changes in blood pressure and also contributes to long-terrn blood pressure regulation. Peripheral sympathetic nerves travel along arteries and veins and are present near the adventitial-medial border to make close contact with vascular smooth muscle cells (Figure 1.3). —— Sympathetic preganglionic fibers — — — , Sympathetic postganglionic fibers MW 2 Brain Stem Cervical Thoracic Lumbar Figure 1.2. Sympathetic divisions of the autonomic nervous system. This is the site of sympathetic neuroeffector transmission. The nerves communicate with the effector cells of the resistance and capacitance vessels and venous system through the release of norepinephrine (NE), a vasoconstrictor neurotransmitter, onto the surface of smooth muscle cells in vessel walls. This is the typical pattern for the sympathetic innervation of blood vessels [23]. Only the smooth muscle cells located at the adventitial/medial border are directly innervated [24, 25]. The appearance and arrangement of sympathetic nerves associated with the veins is similar to that of arteries although nerve density in veins is generally less than that observed in the adjacent arteries [26, 12]. Adventitia Media Intima Blood Vessel Figure 1.3. Sympathetic nerves on a blood vessel. Nerve-mediated constriction of arteries and veins increases peripheral resistance and cardiac output, respectively [27, 28]. Thus, the SNS plays an important role in blood pressure regulation by controlling the two functional parameters of blood pressure: TPR and CO. Although SNS has a central function in homeostasis, in general, and in circulatory adaption, in particular, it is paradoxical that so little is known about the possible contribution of disturbed sympathetic nervous firnction to the development of human diseases [29]. The diseases related to sympathetic nervous system dysfunction include essential hypertension, cardiac failure, coronary artery spasm, cirrhosis, mitral valve prolapse and Raynaud’s syndrome. SNS and mesenteric vasculature. Sympathetic nerves have the greatest quantitative importance in minute-to-minute regulation of vasoconstriction, especially in the splanchnic circulation. The splanchnic circulation receives up to 25 % of cardiac output at rest and is a major reservoir of blood in the body (the largest capacitance bed) [30, 12]. Sympathetic nerve stimulation can reduce intestinal blood volume by up to 60%, leading to a significant redistribution of blood from the splanchnic veins. This change in volume distribution can have a profound effect on overall cardiovascular fitnction including blood pressure [30, 31]. Therefore, the splanchnic vascular bed hemodynarnically represents one of the most important vascular regions [12]. The blood flow to the splanchnic organs, such as the gastrointestinal tract, liver, spleen, and pancreas, is derived from three main arterial trunks: the celiac, the superior mesenteric and the inferior mesenteric artery. The superior mesenteric artery supplies the small intestine and part of the large intestine [32]. The mesenteric circulation refers specifically to the vasculature of the intestines, whereas the splanchnic circulation provides blood flow to the entire abdominal portion of the digestive system. The superior mesenteric artery, the major inflow vessel of the mesenteric circulation, delivers about 12 % of the cardiac output and supplies the entire small intestine [33]. The mesenteric vascular bed originating from the superior mesenteric artery consists of arcades of blood vessels with a gradient of vessel caliber in which the largest blood vessels is most distal to (first order) and the smallest is most proximal to the intestinal wall (third or fourth order). 11 1.5 Sympathetic Neuroeffector Transmission in Arteries and Veins Chemical neurotransmission is initiated by an electrical signal, the action potential. Action potentials are sometimes called spikes or units. They are ca. 100 mV in amplitude and have a duration 1-2 ms [34]. When the action potential fires down to the sympathetic nerve terminal, it causes an influx of Ca2+ into the nerve terminal with subsequent fusion of vesicles with the plasma membrane, followed by exocytosis of NE and co-transmitters, adenosine 5'-triphosphate (ATP) and neuropeptide Y (NPY), into the junctional cleft. The transmitter substances, by convective or diffusive mass transport, reach receptor sites on the postjunctional membrane. Pre-junction (Sympathetic Nerve) Action Potential 3" Diffusion J unctional Cleft <100 nm) Figure 1.4. Sympathetic neuroeffector transmission to blood vessels. P: purinergic receptor, or: adrenergic receptor, ATP: adenosine 5'-triphosphate, and NE: norepinephrine. The transmitters then activate adrenergic and/or purinergic receptors in the effector cell to indirectly trigger specific second messenger pathways through adrenergic receptors and/or directly activate Ca2+ ion channels by purinergic receptors (Figure 1.4). This results in an increased intracellular Ca2+ concentration and vasoconstriction. This neurotransmission process will be discussed in more detail in the following sections. NE and co-transmitters. All blood vessels are innervated by sympathetic nerves that release the neurotransmitters NE, ATP and NPY. Neurotransmitters have been studied extensively since their discovery in 1921 because of their important role in the body. They are the key communication link between neurons. A lack or excess of even one of these neurotransmitters may result in serious disease. NE is one catecholamine, along with epinephrine (EE) and dopamine (DA), found in mammalian tissue. Although NE is present in the adrenal medulla, it functions mainly as a neurotransmitter in the SNS and generally acts locally on effector cells. Abnormal levels of NE are a cause for depression and attention deficit hyperactivity disorder (ADHD). NE is also a powerful medicine used in critically-ill patients as vasopressor. As mentioned, ATP and NPY are neurotransmitters that are co-released with NE from sympathetic nerves and mediate vasoconstriction [35, 36]. ATP serves as the major energy source within the cell for a number of biological processes, such as photosynthesis, muscle contraction and protein synthesis. It is produced by various cellular processes under aerobic conditions with the majority of the synthesis occurring in the mitochondria during oxidative phosphorylation, which is catalyzed by ATP synthase. The main fuels for ATP synthesis are glucose and triglycerides. Extracellularly, ATP has been found to act as a neurotransmitter in the arteries, intestines and bladder [37, 38]. NPY is also localized in sympathetic nerve 13 terminals [39]. NPY in porcine brain was discovered by K. Tatemoto in the early 19805 and is known to be one of the most abundant and widely distributed neuropeptides in the central and peripheral nervous systems [40, 41]. NPY modulates numerous physiological processes including regulation of cardiovascular and renal functions, intestinal motility, memory, anxiety and nociception. NPY is synthesized and co-stored with ATP and NE in peripheral neurons, particularly in the nerve endings surrounding blood vessels. Interestingly, while NPY is primarily co-stored in the large, dense-core vesicle, the smaller vesicles contain essentially only ATP and NE [42, 38]. This peptide is thought to predominantly mediate responses at higher frequencies of stimulation and, therefore, play a critical role during high levels of sympathetic nerve activity [43]. NPY is known to cause constriction of some vascular smooth muscle cells, such as those from the human coronary artery and cerebral circulation. However, some research has shown that NPY does not contribute to vascular constriction in the mesenteric vascular bed [35, 36]. NPY does not have a direct action on postjunctional cells but rather acts as a pre- and postjunctional modulator of both the release and the action of NE and ATP [44, 45]. Therefore, NPY is more likely to function as a neuromodulator in mesenteric vascular beds. Synthesis and metabolism of NE. NE is synthesized by a series of enzymatically catalyzed steps from the precursor amino acid, tyrosine (Figure 1.5). The first reaction is the oxidation of tyrosine into dihydroxyphenylalanine (DOPA) by the amino acid tyrosine hydroxylase (TH); the rate-limiting enzyme in the synthesis. This is followed by decarboxylation to form dopamine through the action of DOPA carboxylase. Finally, dopamine is converted to NE by side-chain hydroxylation through the action of 14 dopamine-B-hydroxylase (DBH) in the synaptic vesicles where NE is stored and released via exocytosis [12]. NE can be further methylated by phenylethanolamine N- methyltransferase to epinephrine. NHz HO m COOH HOWNHCE HO HO Tyrosine EE (epinephrine) Tyrosine hydroxylase Phenylethanolamiue NH; N-methyl transferase H HO DOPA (dihydroxyphenylalanine) NE (norepinephrine) DOPA carbo MA DA - h drox lase HO IDA (dopamine) Figure 1.5. Major steps in norepinephrine synthesis. Released NE in the junctional clefi is rapidly degraded to various metabolites via two main pathways, one involving deamination by monoarnine oxidase (MAC) and the other O-methylation by catechol-O-methyltransferase (COMT). The principle metabolites are norrnetanephrine (NMN) via the enzyme COMT, and 3,4-dihydroxymandelic acid (DMAE), 3-methoxy-4-hydroxymandelic acid (MHMA) and 3-methoxy-4- hydroxyphenylglycol (MHPG) via MAO. However, NE in tissue undergoes a 15 quantitatively different metabolic fate. The NE present in sympathetic nerves is metabolized intracellularly mainly by MAO. The physiologically inactive deaminated product is excreted from the neuron and is then O-methylated extracellularly [46, 47]. Fate of released NE. NE released from sympathetic nerves endings is rapidly cleared by several mechanisms: 1) a portion is removed by diffusion away from the release sites eventually passing into the circulatory system (overflow), 2) a portion is metabolically degraded by COMT or MAO and 3) the major portion is taken back up into the nerve terminal by the norepinephrine transporter (NET) through a process called “reuptake”. Once returned to the neuron, it is either repackaged into the synaptic vesicles for re- release or metabolized by MAO and excreted [48]. 0.2-adrenegic receptors and NE uptake transporters in prejunctional sites. Prejunctional a2-adrenergic receptors located at the sympathetic nerve terminals are called “autoreceptors” and are activated by the exocytotic release of NE. They function by a feedback mechanism to inhibit further release of NE [49-51]. All adrenergic receptors, such as a2 subtypes, are heterotrimeric G protein-coupled receptors. Activation of prejunctional a2-adrenergic receptors inhibit Ca2+ entry, which leads to inhibition of NE release [52]. Some studies have also shown that a2-adrenergic receptors modulate the release of ATP [53-56]. After its release into the junctional cleft, approximately 95% of the NE is transported back into the nerve terminal by the NET. NE reuptake, mediated by a neuronal pump, is frequently the most important of the clearance processes. The uptake probably occurs in the periods between transmitter release, and serves to rid the neuromuscular space of transmitter in order to quickly l6 terminate its vascular action [12]. The NET protein is thought to have 12 transmembrane- spanning domains with intracellularly oriented N- and C- termini [47]. There are two types of NET, one located in the junctional cleft and the other outside of it. Highly efficient NET, referred to as uptake 1, is located in the plasma membrane of sympathetic neurons and is responsible for clearing about 90% of the released NE. Approximately 5% of total released NE penetrates into extraneuronal NET, so-called uptake 2, and is rapidly inactivated by COMT to NMN. The uptake 2 transporter is much more diversely localized in vascular endothelium, smooth muscle, glial and other cells. The remainder of the released NE is either transported into the smooth muscle cell, metabolized by enzymes, or escaped the junction into the extracellular fluid by diffusion. The activity of released ATP is mainly regulated extracellularly by enzymatic degradation and not by reuptake. There are at least two kinds of nucleotidase mediating the degradation of ATP: membrane-bound ecto-nucleotidase particulate and releasable nucleotidase [51]. There is no conclusive evidence as to how the NPY process is regulated. Adrenergic and purinergic receptors in postjunctional sites. NE is a potent vasoconstrictor, acting through al- and a2-adrenergic receptors located in vascular smooth muscle cells. The adrenergic receptors are G protein-coupled metabotropic receptors and the binding affinity of NE for al-adrenergic receptors is higher than a2- adrenergic receptors [57]. NE binding activates the associated Gq protein, which is linked to phospholipase C (PLC). Subsequently increased PLC activity produces a second messenger, inositol 1,4,5-triphosphate (1P3) and diacylglycerol (DAG). 1P3 causes increased intracellular Ca2+ levels from intracellular Ca2+ stores by their binding to 1P3 l7 receptors on the stores, which promote smooth muscle contraction [58]. DAG activates membrane-associated protein kinase C (PKC) and regulates several intracellular substrates involved in Ca2+ handling, such as the Ca2+ channel [59]. NE mediates all components of sympathetic vasoconstriction in rat mesenteric vein (MV) associated with sympathetic nerves mainly by postjunctional a1 and/or a2-adrenergic receptors, whereas ATP plays a more important role than NE does in neurally-mediated vasoconstriction in rat mesenteric artery (MA) with their binding to postjunctional P2X purinergic receptors. The P2X receptors, one of the P2 receptor subtypes, are ligand-gated ionotropic protein exhibiting calcium permeability that contribute to vascular smooth muscle contractions [60]. Vasoconstriction, mediated via P2X-receptors, is mainly dependent upon influx of extracellular calcium ions through voltage-dependent calcium channels and the response to nerve-released ATP is faster while the response to NE is more gradual [3 7, 61]. 1.6 Experimental Hypertension Animal Models A number of animal models of hypertension have been developed to study the role of these various blood pressure regulating systems in the development of secondary and essential hypertension. Since each animal model has a different etiology, it is conceivable that the choice of a model significantly influences the outcome of the experiments performed. Both spontaneous hypertensive rat (SHR), generated from the Wistar-Kyoto (WKY) normotensive strain, and Dahl salt-sensitive rats, generated from normotensive Sprague-Dawley, are the most widely used genetic strain of hypertensive animals [62, 63]. The SHR model is relatively insensitive to changes in sodium intake in the developmental stages of hypertension, whereas the development of hypertension in Dahl 18 salt-sensitive depends on sodium intake. Therefore, Dahl salt-sensitive rats are used to study not only genetic but also environmental factors [63]. The two-kidney, one-clip (2K1C) and the one-kidney, one-clip (lKlC) hypertension models are also commonly used to study for renal-dependent human secondary hypertension. Renal models of hypertension use blockade of renal blood flow via stenosis of the supplying arteries with a silver clip. This causes a partial occlusion of blood flow and /or reduction of renal mass. In 2K1C, one renal artery supplying on one kidney is clipped but both kidneys are left intact, while in lKlC, one kidney is removed and the renal artery supplying another kidney is clipped. The deoxycorticosterone acetate (DOCA)-salt model, with suppressed renin activity due to water and sodium retention is also most widely used to examine neural and hormonal contributions to hypertension, which are relevant to the understanding of essential hypertension [64]. DOCA-salt hypertension. This model is a relatively cheap and easy to employ, and the control state is easy to define unlike many genetic states of hypertension [64]. The DOCA-salt hypertension model is produced by surgical uninephrectomization followed by implantation of the mineralocorticoid, called DOCA, and the administration of excess salt, NaCl, in the drinking fluid. Arterial blood pressure will significantly rise in a few weeks after such a treatment. If left untreated, the animal will eventually lose weight and develop end-organ damage [65, 64]. DOCA salt-treatment causes significant changes in the cardiovascular system that contribute to the overall pathology of this model. The administration of mineralocorticoids changes hormonal and neural pressor mechanisms [64]. The main role of mineralocorticoids is to affect ion transport, such as the Nair/H+ exchange and the Na+ pump in kidney epithelial cells, which conserves Na+ in the body. 19 Therefore, mineralocorticoids increase the rate of Na+ reabsorption by the kidneys. Na+ imbalance may be related to alternations in the central and peripheral control of autonomic nervous system function [64]. An increased reabsorption of Na+ will increase water retention, causing increased blood volume and finally increased blood pressure [66]. The sequence of events in the pathogenesis of DOCA-salt induced hypertension is not entirely clear but a number of consistent features have been observed allowing a generalized picture to be formed [64]. Several mechanisms have been suggested to contribute to the pathogenesis of DOCA-salt hypertension. Relevant prehypertensive changes would include elevated sympathetic activity, altered neural angiotensin II and vasopressin activity, and distorted baroreflex response [67, 64, 68]. In addition, numerous peripheral effects of DOCA-salt hypertension induce characteristic changes in vascular structure enhanced vascular reactivity, and altered ion transport that contribute to the overall pathophysiology of DOCA-salt state [64]. 1.7 Hyperactivity of the SNS in Hypertension Alterations in the activity and function of sympathetic nerves are associated with cardiovascular disease [69]. Although the precise causal mechanisms leading to sympathetic augmentation in hypertension are still poorly understood and controversial, growing evidence suggests that essential hypertension and hypertensive animal models are initiated and sustained by overactivity of the sympathetic nervous system [70]. Abnormality of peripheral sympathetic outflow and catecholamine metabolism have been described in the development of hypertension in humans and the animal hypertension models, SHR and DOCA-salt rats [71, 48, 72]. Overactivity of the sympathetic nervous 20 system is commonly present in younger patients with essential hypertension. Early hypertension is characterized by increased cardiac output as a result of hyperactivity of cardiac sympathetic nerves. Therefore, it is generally agreed that hyperactivity of the SNS is a critical initiating factor in hypertension development. Elevated cardiac and renal sympathetic activity were also demonstrated in patients with essential hypertension [73, 74]. In the longer-term, hyperactivity of the SNS may cause increased renal vasoconstriction and renal sodium retention, thickening of blood vessel walls, etc. These observations support an ongoing role of the SNS in chronic hypertension. Overactivity of the SNS is an attractive candidate mechanism to explain human and animal hypertension, however, it is unclear whether the SNS plays a substantial role in long-tenn blood pressure control in humans [6]. 1.8 Increased NE Release in Hypertension The development of hypertension in human and experimental hypertensive models has been associated with increased basal sympathetic tone, as was suggested by the observation of higher NE concentrations or its synthesis rats in the heart and other highly vascularized organs [75-77]. Increased NE release are seen in some disease states, including cardiac failure, cirrhosis, depressive illness, and essential hypertension [29]. The level of sympathetic drive differs between vascular beds and plasma NE is not consistently elevated in hypertension. Esler’s study showed that total NE spillover to plasma is increased in essential hypertension but is less than noted in patients with cardiac failure or cirrhosis. Approximately 50% of the increase in total NE spillover in 50 patients with untreated essential hypertension was explicable in terms of increased overflow of NE from the kidney and heart. However, there was no increase in NE 21 overflow from lungs, skeletal muscle or hepatomesenteric circulation in hypertensive patients [78]. This may reflect, in part, selective sympathetic activation to specific vascular beds, such as the kidney and heart, but a regional pattern of increased sympathetic nerve activation to the kidney and heart might have originated remains unclear. Increased NE release in hypertension suggests that such alteration can take place at the level of the sympathetic nerve terminal. It has been generally hypothesized that abnormalities of NE release in hypertension were due to impaired prejunctional a2- adrenergic receptor function and/or neuronal reuptake, and elevated nerve firing rates. However, few studies have shown that there was alteration in purinergic transmission associated with ATP in hypertension. The possible implication of neuropeptide Y, which was also found to exert presynaptic inhibitory influence on the liberation of NE by sympathetic nerves also needs to be explored in the hypertensive mechanism. Alteration in local adrenergic modulation mechanisms on sympathetic nerves in hypertension. Data from several studies indicate that the increased NE overflow in hypertensive animals is due, at least in part, to impaired prejunctional a2-adrenergic autoreceptor function [79-81]. As examples, NE release was increased in vascular sympathetic nerves of the isolated mesenteric vasculature of DOCA-salt rats and in the portal vein and caudal artery of SHRs due to impaired a2-adrenergic autoreceptors [82, 83, 68]. However, controversial results have been also reported indicating that there was no difference in inhibitory function of prejunctional a2-adrenergic receptors in the cadual artery, or the perfused mesenteric bed of young SHRs and in the isolated mesenteric artery of DOCA-salt rats [28, 82]. Although prejunctional otZ-adrenergic receptors associated with sympathetic nerves also regulate NE release in veins [84, 85], only a few 22 studies of theese function in hypertension and the results have been controversial [77, 68]. Another possibility is that other modulatory mechanisms mediated by a variety of presynaptic or local heteroreceptors can also be altered in hypertension but this remains to be investigated [28]. Alteration of NE uptake in hypertension. Neuronal uptake of NE is the activity mechanism responsible for removing NE from the synaptic cleft once released from the nerve terminal. Previous studies have shown that there is impairment in total and cardiac neuronal NE reuptake in patients with essential hypertension and in hypertension animal models [86, 87, 67, 48, 70]. However, the evidence is inconclusive and controversial. Some studies revealed that neuronal NE reuptake impairment may facilitate increased NE spillover from sympathetic nerves to plasma due to decreased clearance of NE from plasma for the body as a whole and in individual organs, such as the heart and brain [88, 89, 70]. Studies with a variety of hypertensive animal models, including SHR and DOCA-salt rats, have also shown that NE reuptake was reduced. As examples, reuptake was reduced in the skeletal muscle and kidney of chronic hypertensive SHR and the mesenteric artery of hypertensive dogs after the development of hypertension [90, 91, 72]. On the other hand, neuronal NE uptake was significantly increased in isolated subcutaneous resistance vessels of human essential hypertensive patients, in SHR mesenteric and tail arteries, and in DOCA-salt rat mesenteric veins [92, 93, 17, 31]. Moreover, there was no difference in uptakegactivity in DOCA-salt rat [94, 67]. Therefore, some pertinent questions remain to be answered such as whether or not impaired prejunctional a2-adrenergic receptors and/or alteration of NET in sympathetic 23 nerves are directly involved in the increased NE release, altered contractile responses of blood vessels and pathogenesis of essential hypertension. 1.9 Assessment of SNS Activity Clinical methods have been devised to assess SNS activity based on measurement of NE levels in tissue, spillover into plasma and urinary excretion. Also, analysis of the metabolite levels in patients with essential hypertension provides an assessment of the activity of sympathetic nerves innervating the organ [86, 95]. Therefore, NE overflow is an index of sympathetic activity [29]. However, great caution must be used when attempting to relate plasma or urine NE levels to increased nerve activity or to a specific disease state because the sympathetic outflow to all organs is not uniform. The level of circulating NE depends not only on the release of NE from the nerve terminals but also on plasma clearance of NE [96, 97]. Local and organ-specific increases or decreases in sympathetic activity can also occur with different reflexes and in different disease states. Moreover, only a small fraction of the NE released by sympathetic nerves throughout the body escapes into the blood stream and urine. Therefore, the release of NE from perivascular nerve has primarily been investigated using regional tissues, such as the heart and kidney, labeled with tritiated NE instead of measuring NE in the whole body [98, 99]. In a smaller number of studies, the release of endogenous NE overflow in tissues by electrical nerve stimulation has been also measured by hi gh-perfonnance liquid chromatography [100, 85]. However, due to the relatively low concentration of NE, it is usually necessary to measure NE release elicited by long trains of electrical stimuli; a condition that dose not closely mimic normal nerve firing [85]. Moreover, short stimulus 24 trains (5 3 8) may mimic in vivo sympathetic nerve activity, which occurs in short bursts (< 2 s) [101]. Although the increased NE release from tissue or spillover to plasma is also undoubtedly attributable, at least in part, to increased sympathetic nerve activity, NE release cannot be measured directly with these methods [70]. It turns out that little is known prejunctional or local modulatory mechanisms controlling NE release and action from altered sympathetic nerves in hypertension. The electrochemical recording approach provides simple and extremely sensitive approaches and makes it possible to monitor, near release rate, the real time rise and fall in the concentration of released endogenous NE in blood vessels on an impulse-by-impulse basis during low and short fi'equency trains of stimulation. Learning about the characteristics of NE release on an irnpulse-by- impulse basis is important for understanding how the sympathetic nervous system regulates blood pressure, and more importantly, what alterations in the regulatory mechanism are associated with hypertension and cardiovascular disease. Electroanalytical techniques. Electrochemical measurements with solid electrodes were introduced more than 50 years, but there was little interest in them by biologists until the 19703. A breakthrough occurred in the 19703 when Ralph Adams and his colleagues showed that it was possible to implant a small working electrode in a rat brain and detect electroactive biogenic amines and related substances in the extracellular fluid (ECF) in vivo using voltamrnetry [34]. Their work demonstrated that it was possible to study the release of neurotransmitters in intact animals, without the use of complicated radioactive labeling techniques [102]. These, electrochemical measurements with microelectrodes provide a sensitive, selective and rapid detection method for chemical events occurring in tissue. 25 Since that time, other electroanalytical techniques, including fast scan voltammetry, continuous amperometry and various pulsing methods, have been used with microelectrodes to provide unique information on changes in the local concentration of neurotransmitters and related substances. Easily oxidized or reduced species can be detected, both in vivo and in Vitro, in the mammalian central nerve system, in single-cell preparations, brain slices and other tissues [103-105]. Fast scan cyclic voltammetry (FSCV) and continuous amperometry are the tools commonly used in recent work, which has focused primarily on the biogenic amine neurotransmitters, dopamine, norepinephrine, serotonin and metabolites of these, and ascorbic acid [106-109]. Both electroanalytical techniques offer a degree of selectivity, temporal resolution and sensitivity in the study of neurochemical events [110]. In continuous amperometry, a microelectrode is held at a constant potential sufficient to oxidize or reduce in an analyte of interest at a mass transfer limited rate and the resulting electrochemical current is monitored with time. Therefore, amperometry has excellent temporal resolution and allows detailed real-time analysis of the kinetic processes involved in the release of vesicular content into the extracellular space. Using amperometry, the contents of very small vesicles and small concentrations of transmitter released from a single cell, such as a chromaffin cell, can be quantified [111, 112]. The charge (Q) in coulombs is related to the number of moles of analyte detected by Faraday’s law, Q = nFN, where in is the number of electrons involved in the electrochemical reaction (n = 2 for catecholamines) and F is Faraday’s constant (96,485 C/equivalent). Recent studies using amperometry have allowed for the detection and quantification of attomole and zeptomole amounts of neurotransmitter upon release from single cells [113]. Such sensitive measurements using 26 amperometry are enabled by the development of low noise instrumentation and enhanced filtering techniques [110]. Therefore, amperometry is an excellent technique for monitoring rapid transient changes associated with neurochemical signals of known origin. However, it provides no qualitative information due to the poor selectivity from interferences, other oxidizable neurotransmitters and their metabolites, ascorbic acid and uric acid, that are present in brain extracellular fluid at very high concentration [34, 105]. For FSCV, electroactive species can be detected at very high scan rates (> 100 V/s) with the scan lasting only a few milliseconds. This reduces the extent of electrocatalysis [114] and allows for more frequent measurements. Based on the potential of the oxidation and reduction peaks as well as the peak current ratio, FSCV provides qualitative as well as quantitative information about redox analyte. At high scan rates, a relatively large but constant charging current is observed. Smaller faradaic currents can only be detected by background subtraction. The resulting faradaic current provides a ‘fingerprint’ for the molecules detected at the electrode that represent changes in the local concentration of these molecules [110]. A major difference between amperometry and FSCV is the selectivity of FSCV possesses in vivo and in vitro for different electroactive molecules of interest. Most of the work to date using these methods has been to study neurogenic processes in the central nervous system. However, there have been some reports on the use of continuous amperometry to study NE overflow from peripheral sympathetic nerve terminals [115-118]. Electrical measurements with a microelectrode have been used since the 1990's to study neurogenic processes in the peripheral nervous system, such as the release of NE from the sympathetic nerve terminals of isolated organs [115, 119-121, 27 118]. Although NE, along with co-transmitters ATP and NPY, are released from sympathetic nerves innervating blood vessels, NE is the only one that can be detected directly electrochemically. NE is oxidatively detected with a microelectrode via a 2H’72e' oxidation reaction. Mermet, Gonon and Stj time were the first to study NE release from the sympathetic nerve terminals innervating the smooth muscle cells of the rat tail artery [115, 119, 122]. In fact, the rat tail artery has been the organ of focus in most investigations of the peripheral nervous system because it is densely innervated exclusively by sympathetic nerve fiber, that form a two-dimensional plexus at the external surface [115, 119, 122]. In a series of more recent papers, Brook and coworkers studied the characteristic features of NE release from postganglionic sympathetic nerves in rat mesenteric artery mesenteric artery [123, 121, 124]. The release of endogenous NE was measured by continuous amperometry using a carbon fiber microelectrode coated with the permselective ionomer, Nafionm. The group found that NE release is regulated by activation of prejunctional a2-adrenergic receptors and that clearance of the released NE in this tissue depends, in part, on neuronal reuptake. 1.10 Microelectrodes Microelectrodes can be fabricated with such small size (0.5 to 5 pm diam) that they can be used to probe chemical events inside a single biological cell. The extremely small size of the recording electrode can be placed in close proximity to the site of release or to a synapse without appreciable perturbation of the surrounding environment [125]. Its use enables the discrete sampling of small nuclei within the brain with a spatial resolution of 28 10-100 pm. Most electrodes used for in vivo or in vitro measurements have diameters of 10-30 pm and sample from a region of comparable size. Therefore, microelectrodes have a key role for detection of neurotransmitters. Advantages of microelectrodes. Their very small size provides the advantage of high temporal resolution and the ability to make spatially resolved chemical measurements in tissue [103]. Due to their small size (10-30 pm in diameter, 30-500 pm in length), the current at these electrodes is extremely low, so some sophisticated monitoring electronics are needed. Small electrodes are suitable for use in tissue with minimal peripheral damage to the surrounding environment. Moreover, since electrochemical currents at microelectrodes are 4 to 5 orders of magnitude smaller than those seen at electrodes of conventional size, they can be used to make measurements in solutions of high resistance (i.e., less iR loss). The low capacitance of the small area electrode allows the electrode potential to be changed rapidly. Therefore, microelectrodes are exclusively useful for fast measurements, like FSCV (300V/s) [126, 127]. This rapid measurement enables monitoring of millisecond and microsecond chemical events (e.g., exocytosis) and kinetic processes (e.g., reuptake). As a consequence of the reduced charging current and increased faradaic current (enhanced mass transport), microelectrodes exhibit excellent signal-to-noise (S/N) characteristics, which enhances sensitivity. Depending on the time scale of the experiment, voltammograms obtained at microelectrodes may differ from those obtained at electrodes of conventional size. Since the large diffusion layer relative to the microelectrode dimensions and the volume from molecules diffusing to suppdrt the current are relatively large by the radial diffusion, steady-state (time independent) response is observed. Contrarily, because the dimension of the diffusion layer is smaller 29 than that of the conventional size and the ratio of the number of molecules diffusion perpendicular (planar diffusion) to those diffusion in at the edges (radial diffusion) is much greater at conventional size, the peak shape response (time dependent) is observed under these conditions. The steady-state limiting current is directly proportional to the analyte concentration, which provides a direct means of quantitating electroactive substances [125]. 1.11 Carbon Fiber Microelectrode Carbon is the material of choice for fabricating microelectrodes for use in the brain. The response of different materials is differentiational selectivity, sensitivity, stability and reproducibility [125]. Noble metals, such as Pt and Au, and carbon fiber have been used as microelectrodes. However, Pt and Au are unstable and their surface is deactivated immediately upon exposure to brain tissue [128, 129]. Carbon fiber differs from metals in both its electronic properties and surface chemistry [130]. Another major difference between carbon and metal electrodes is the nature of the surface oxides. Metals form oxides and hydroxides easily and reversibly in aqueous solution, whereas carbon forms a richer variety of surface oxides, such as carboxyl and hydroxyl functional groups [131]. The surface oxides can have a variety of effects on the electron-transfer rates and adsorption of catecholamines and interferences. The presence of oxide functional groups raises the possibility of specific chemical interactions between the surface and a molecule in solution. Such interactions can catalyze redox reactions, and promote or inhibit adsorption. 30 Carbon surface oxides can also greatly increase the background voltammetric current, in part, due to interactions between charged surface groups and electrolyte ions and the psudocapacitance associated with redox-active functionalities (e. g., quinone/hydroquinone). Carbon fiber types differ significantly in the degree of microstructural orientation, crystallite size, and surface chemistry, and these parameters are important for electrochemical performance. Most carbon fibers are made with polyacrylonitrile (PAN) or petroleum pitch and formed during high temperature pyrolysis of the materials. PAN fibers have low tensile strength and more microstructural defects in the graphite sheets. Pitch type carbon fibers tend to have high tensile strength and modulus due to a greater level of microstructural ordering. Carbon fibers are traditionally used as strengthening materials and because of their good electrical conductivity, non-toxicity and small size, they have been applied for both the electrochemical detection of oxidizable compounds and extracellular single unit recording [132]. Moreover, because of their resistance to drifi when exposed to biological tissue and easy of surface modification, carbon fibers have become popular and are now widely employed for in vivo and in vitro monitoring of oxidizable compounds in the living tissue [133, 103, 134-136]. Limitations of carbon fibers for in vivo and in vitro measurements. One limitation of carbon fibers is their fragility and response deactivation over time during exposure to physiological environments due to fouling by biomolecule adsorption. The adsorption is linked to the presence of the extended 7! electron system and the carbon- oxygen functional groups on the surface. A 30 — 50% decrease in the short-tenn electrode response sensitivity is not uncommon following implantation into brain tissue. The 31 response generally levels off after ca. 2 hours [137, 131]. For in vivo measurements, electrode response stability is required for long periods of time, at least 8b in most experiments [126]. Therefore, the bare carbon fibers are unsuitable for chronic implantation in tissue [138]. A wide range of surface carbon-oxygen functional group exist on the electrode that can affect the electrode kinetics of various classes of redox systems in different ways. Quinone-like functional groups are known to be present and exhibit pH-dependent electroactivity. It is the oxidation/reduction of these functional groups that gives rise to background voltammetric features. The shifting of voltammetric background current with pH leads to complication in the background-subtracted voltammetric responses because these groups are oxidative at the same potentials as the catecholamine neurotransmitters. Simultaneous changes in the catecholamine concentration and pH frequently occur in biological tissue as a result of the coupling of respiration and energy production to pH [139, 140]. Several electrode pretreatrnents have been developed over the years to improve the response sensitivity, stability and selectivity especially for catecholamines. These including coating the surface with the parmselective polymer, NafionTM [134]. Coating with NafionTM improves selectivity and minimizes electrode fouling from biomolecule adsorption by repelling anions from SO3' groups but allowing cations through. However, polymer coatings increase the background current and decrease the electrode response time, which results in decreased sensitivity [141, 104, 134]. Another perfidious pitfall is that some drugs alter the oxidation reaction, and the response time to evoke release is slow (10-30 s) at treated electrodes. Therefore, the real kinetics of the evoked release cannot be directly recorded. Therefore, a more stable 32 microelectrode material would enable more widespread application of this informative technique. 1.12 Properties of Diamond and Its Application Diamond offers exceptional material properties including mechanical strength, thermal conductivity, optical transparency, and electrical conductivity when doped compared with any material. Another important property for electrochemical application is the microstructural stability and corrosion resistance. These properties are a consequence of the small interatomic distance between tetrahedrally-bonded carbon atoms in the diamond lattice and the hydrogen surface termination. Synthetic diamond can be produced by high-pressure, high—temperature techniques or by low-pressure chemical vapor deposition (CVD). CVD diamond became a reality in the early 19705 and affords the possibility of producing thin film diamond on a variety of substrates at relatively low cost. To date, hot filament and microwave-assisted CVD are the most popular deposition methods with a CH4/H2 source gas mixture routinely used for growth. The high concentration of H2 produces a high flux of atomic hydrogen, which is known to be critical for high quality diamond growth. It is possible to manipulate the physical and chemical properties of diamond by proper control over the source gas composition, system pressure and substrate temperature. The crystalline diamond produced is used for a wide range of industrial applications that require hardness and wear resistance. Other specialized applications also exist or are being developed including its use as active and passive materials, semiconductor. Diamond is naturally an excellent electrical insulator but it can be made electrically conducting by controlled doping with impurities, such as 33 boron. CVD allows for in situ doping with boron, phosphorus and sulfur, rendering the diamond film p-type or n-Iype semiconductors [142]. The boron impurity serves as an electron acceptor with an activation energy of ca. 0.37 eV [143]. Boron is the most prominent dopant because of its small covalent radius, which leach to easy incorporated into substitutional sites within the diamond without causing lattice distortion. In order to have sufficient electrical conductivity for electroanalytical measurements (_<. 0.1 0 cm), polycrystalline diamond films must be doped with boron at a concentration of l x 10'9 cm"3 or higher. This produces a carrier concentration in the high 1 x 1020 to low 1 x 1019 cm'3 and a carrier motility between 0.5 and 10 cm2 V'ls'l [144]. At present, no appropriate method exists for preparing highly conducting n-type diamond electrodes. Well-known donors, such as nitrogen and phosphorus, have too high an activation energy (1.7 and 0.6 eV, respectively) and thus, cannot impart reasonable room temperature conductivity. However, recently, sulfur was reported to be a donor impurity for diamond with an acceptable activation energy (0.36 - 0.38 eV) [145]. The advent of low-pressure diamond synthesis along with the facility of dopant incorporation, has prompted a growing number of investigations into its use as an electrode material. Diamond thin films with sufficient electrical conductivity undoubtedly are promising candidates and attractive electrode material. The use of diamond in electrochemistry is a relatively new field of research that has only begun to blossom in recent years since the early 19803 [146-150]. 34 1.13 Boron-Doped Diamond Electrodes A new electrode material, boron-doped diamond thin-film, with an sp3-bonded carbon microstructure and little surface oxygen (< 0.02 atomic %) due to a hydrogen surface termination, offers advantages for electroanalytical measurements in terms of linear dynamic range, limit of quantitation, response time, response precision, and response stability over other electrodes, especially commonly used sp2 carbon material with extended n-electron systems, such as glassy carbon, pyrolytic graphite or carbon paste [146, 151]. The relative absence of electroactive carbon-oxygen fimctionalities on the hydrogen-terminated diamond surface (hydrophobic, non-polar, and chemically stable) introduces superb electrochemical properties [152]. First, diamond possesses wide working potential window (3-4 V) and excellent response stability. These properties enable diamond’s use in measurements that otherwise would be difficult or impossible with spz-bonded carbon electrodes. Second, diamond electrodes give improved signal-to- background (S/B) and signal-to-noise (S/N) ratios in electroanalytical measures due to a low double layer capacitance (l to 8 uF/cmz) compared with sp2 carbon electrodes (30 to 40 uF/cmz), and the absence of background current due to redox-activity surface carbon- oxygen functional groups [153]. The slightly lower density of surface electronic states near the Fermi level caused by the semiconductor nature of boron-doped diamond contributes to the low capacitance of the material [153, 154]. A lower surface charge carrier density at a given potential leads to a reduced accumulation of counter-balancing ions and water dipoles on the solution side of the interface, thereby a lowering of the background current and capacitance. Third, diamond exhibits high resistance to surface fouling due to the weak adsorption of polar molecules and a pH-independent background 35 voltammetric response due to the hydrogen surface termination [153]. The majority of the research conducted so far has involved the use of a planar thin film (macroelectrodes). The fabrication, characterization, and application of diamond microelectrodes have not yet been extensively studied. There are only a few reports describing the preparation and basic electrochemical characterization of diamond microelectrodes [155-157]. Diamond microelectrodes have been found to provide superior detection figures of merit, as compared to carbon fibers in terms of linear dynamic range, limit of detection, response variability and stability [158]. 1.14 Outline This work described in this dissertation focused on the application of diamond microelectrodes for measurements in biological environments. Specifically, these new microelectrodes were used to record in vitro NE overflow from sympathetic nerves innervating mesenteric arteries (MA) and veins (MV) in normotensive (sham) and hypertensive (DOCA-salt) rats. We sought to accomplish these research goals: (i) fabricate and characterize diamond microelectrodes in comparison with conventional carbon fibers, (ii) use continuous amperometry and video microscopy in vitro to learn more about how the sympathetic nervous system (SNS) regulates the tone of MA and MV and (iii) use the same two tools to better understand hypertension-associated changes in NE overflow and adrenergic vasoconstriction mechanisms in MA and MV. An overview of each chapter is given below along with a synopsis of the findings. Chapter 1 introduces the background of and motivation for the research. Chapter 2 is referred to throughout as the experimental section, as it contains details of all 36 experimental procedures, tissue preparation and microelectrode fabrication. In Chapter 3, the characterization of microcrystalline boron-doped diamond microelectrodes is described in comparison with carbon fibers. The purpose of this work was to investigate the fundamental differences between carbon fiber and diamond and learn about the physicochemical properties and electrochemical responses of the microelectrodes during measurement in biological tissue. Chapter 4 describes the applications of a diamond microelectrode coupled with video imaging for simultaneous monitoring endogenous NE overflow at the surface of a MA and its effect on the contractile response in vitro. The use of various pharmacological agents to better understand the neurogenic response is also described. This study shows that there is a correlation between the NE overflow, as elicited by electrical stimulation, and the dynamics and extent of vascular constriction. Several observations indicate that the bare diamond microelectrode provides superior sensitivity, reproducibility and stability coupled with a bare carbon fiber microelectrode. Chapter 5 presents the fundamental differences in sympathetic neuroeffector transmission to rat MA and MV, which contribute to their different hemodynamic functions. These differences include the neurotransmitters released from sympathetic nerves, the time course of NE action, the arrangement of perivascular nerves, and the function of pre- and postjunctional receptors on MA and MV. Prej unctional a2- adrenergic autoreceptors and NE uptake play a greater role in regulating NE availability at the neuroeffector junction in MA than in MV, and NE overflow in rat MV exceed that in MA. The differences contribute to, at least in part, the greater sensitivity of MV to the constrictor effects of sympathetic nerve stimulation compared to MA. Chapter 6 presents alterations in sympathetic neuroeffector transmission to MA and MV in DOCA-salt 37 hypertensive rats. Increased NE availability in MA due to impaired a2-adrenoceptors contribute to increased peripheral resistance in DOCA-salt rats. There is also an increase in NE uptake in DOCA-salt hypertensive rat MA. The increased uptake offsets the vasoconstrictor effects of elevated NE overflow in these tissues. Finally, a summary of all the results, and the conclusions is given in Chapter 7. 38 Chapter 2 Experimental Section 2.1 Boron-Doped Diamond Film Growth Boron-doped diamond thin film was deposited on a sharpened Pt wire by microwave-assisted chemical vapor deposition (CVD) [155, 150, 159]. The Pt wire was electrochemically etched in l M KOH. An etching solution containing 7.0 g of CaClz-Hzo in a mixture of 20 mL of ultrapure water and 20 mL of acetone also functioned well [160, 159]. A 1.4 cm long piece of Pt wire (99.99%, Aldrich Chemical, 76 um diam) was used. Both ends of the wire were sharpened so that two electrodes could be made from one wire. The wire end was immersed to a depth of l to 2 mm and positioned in the center of four connected carbon rod counter electrodes during the etching. An AC polarization of 12 V (60 Hz) was applied between the wire and the counter electrodes using a variable autotransfonner (Staco Energy Products, Dayton, OH). The etching procedure lasted for approximately 5 3 until gas evolution visibly ceased at the tip. The etched wire was conically-shaped near the end and cylindrically-shaped above. A thin film of boron-doped diamond was deposited on the sharpened wire using a commercial CVD system (1.5 kW, 2.54 GHz, ASTeX, Woburn, MA)[158, 159]. Prior to deposition, the sharpened wire was ultrasonically cleaned in acetone (5-10 min) and ultrasonically seeded (30 min) in a diamond powder suspension (5 nm particles, ca. 20 mg in 100 mL of ethanol, Tomai Diamond Co., Tokyo, Japan). In order to prevent damage to the sharpened end, the wire was cleaned and seeded while vertically 39 suspended in the agitated solution. During the seeding process, the surface is scratched by the diamond particles with some getting embedded. Both the scratches and the embedded particles serve as the initial nucleation sites for film growth. A high instantaneous nucleation density is desired because this leads to the formation of a continuous film in the shortest time and at a low nominal thickness. The pretreated Pt wire (3-4 in parallel per deposition run) was mounted horizontally on top of a quartz plate (10X10X1 mm) in the reactor. The quartz plate was placed in the center of the reactor’s molybdenum substrate stage and served to thermally isolate the wire. The thin film of boron-doped diamond was deposited from a 0.5% CH4/H2 (v/v) source gas mixture with 10 ppm of diborane (0.1 % B2H6 diluted in H2) added for doping. All source gases were ultrahigh purity grade (99.999%). The system pressure was 45 Torr, the substrate temperature was ca. 700 °C, or less, (as estimated with an optical pyrometer), the microwave power was 400 W and the total gas flow was 200 standard cubic centimeters per min (sccm) during growth. The deposition time was 10 h. After deposition, the CH4 and B2H6 gas flows were shut-off and the films cooled in the presence of atomic hydrogen (H2 plasma) by slowly reducing the plasma power and system pressure over a 30 min period. This post-growth treatment was necessary for removing adventitious sp2 carbon impurity, minimizing dangling bonds and ensuring full hydrogen termination. Under these conditions, the nominal growth rate was estimated to be about 0.3 um/h based on the final film thickness of 3-5 pm. Once coated, the diameter of the cylindrical portion was approximately 80 um [155, 160, 159]. 40 2.2 Diamond Film Characterization The electrode surface morphology and microstructure were investigated by scanning electron microscopy (SEM) and visible-Rarnan spectroscopy. Scanning electron microscopy (SEM). SEM was performed with either a JEOL 6400V or a JEOL 6300F field-emission electron microscope. Figure 2.1A shows SEM images; a sharpened wire coated Pt wire (top) and a Pt wire coated with a boron-doped diamond film (bottom). The wire is covered with polycrystalline diamond as no major cracks or areas void of diamond were seen [155, 159]. Moreover, electrochemical tests described below in section 2.4 revealed no evidence for exposed Pt, consistent with complete film coverage. A higher magnification image of the coated wire is presented in Figure 2.18 The polycrystalline film consists of randomly oriented, well-faceted crystallites ranging in diameter from 0.5 to 3 um. There is a wide crystallite size distribution with smaller secondary crystallites existing on the larger primary crystallite facets. Figure 2.1. SEM images of (A) an electrochemically sharpened Pt wire (top) and a Pt wire coated with a polycrystalline boron-doped diamond film (bottom). (B) An expanded view of the end of the diamond-coated Pt wire. 41 This wire is not the only substrate that can be coated with diamond as smaller diameter Pt wires (25 and 10 um), tungsten (W) wire and quartz rod (ca. 270 um) have also been coated in our laboratory. The most conformal coating and reproducible electrode properties, however, were obtained with the 76 um diam Pt wire. Therefore, this substrate was used for all microelectrode preparation in this work. It should be noted that the diameter of this wire is large in comparison with microelectrode diameters commonly used (<30 um diam), but was nonetheless useful for our measurements. Raman Spectroscopy. Raman spectra were acquired at room temperature using a RAMAN 2000 spectrograph (Chromex, Inc., Albuquerque, NM) that consisted of a diode-pumped, frequency-doubled continuous wave (CW) Nd:YAG laser (500 mW at 532 nm, COHERENT), a Chromex 500 is spectrometer (f/4, 600 grooves/mm holographic grating), and a therrnoelectrically cooled 1024 x 256 element charge-coupled device (CCD) detector (ANDOR Tech., Ltd., South Windsor, CT). Spectra were collected with an incident power density of ca. 500 lecm2 (100 mW at the sample and 3 pm diameter spot size) and a 10 3 integration time. A white-light spectrum was collected under the same conditions and served as the background for spectral correction. The spectrometer was calibrated (wavelength position) with 4-acetamidophenol (CH3CONHC6H40H). Figure 2.2 shows a series of Raman spectra for diamond films deposited on the 76 um diam Pt wire using different CH4/H2 ratios. All spectra possess an intense one-phonon diamond line at 1332 cm". The fact that there is no up-shift or down- shift of the peak means that the diamond film is not under significant tensile or compressive stress [161]. There is also some scattering intensity near 1580 cm", 42 particularly for the 1% CH4/H2 film, which is attributable to amorphous spz-bonded carbon impurity. 1332 cm" 1530 cm" Intensity 7.. 0.5% 1600 ' 1500 2000 Raman Shift (cm") Figure 2.2. Raman spectra for diamond thin film deposited on Pt wire with different methane-to-hydrogen ratios. The scattering intensity in this region increases with the CH4/l-Iz ratio used in the source gas indicating that more sp2 carbon is formed at the higher source gas carbon levels [162, 163]. We suppose that much of the impurity responsible for the signal is located at the interface between the Pt and diamond film and not on the film surface. The 1332 cm'1 line intensity decreases and the linewidth (full width at half maximum, FWHM) increases with increasing CH4/H2 ratio. For example, the one-phonon mode linewidths are 7, 11 and 17 cm '1 for the films deposited with 0.5, 0.75, and 1.0% CH4/H2, respectively. By way of comparison, the line width for a piece of single crystal diamond using the same instrumental settings was ca. 2 cm'l . The linewidth, to a first approximation, is inversely related to the phonon lifetime. The larger linewidth for the films on Pt is caused by increased phonon scattering as a result of the grain boundaries and defects in the polycrystalline film [164, 163]. Based on the Raman spectral features, 43 the films deposited from 0.5% CH4/H2 are of good quality and were used throughout this work. 2.3 Microelectrode Preparation The carbon fiber was a heat-treated (3000 0C) pitch-base type (Textron Speciality) with a nominal diameter of 35 pm. The resulting microcylinder electrode was disk- shaped at the end. For the electroanalytical measurements, both the carbon fiber and diamond microelectrodes were attached to a Cu wire with conductive epoxy. Each electrode was then sealed in polypropylene [165, 166]. This was accomplished by inserting the microelectrode into a pipet tip and heating the tapered end using the heating coil of the micropipet puller. The assembly is shown in Figure 2.3A. Electrodes insulated with polypropylene are non-toxic to tissue and chemically resistant to the isopropyl alcohol (IPA) used for electrode cleaning. This insulating procedure was found to be great for sealing the rough diamond surface. SEM images of the polypropylene-coated insulated diamond microelectrode are presented in Figures 2.3B and C. The conical shape of the exposed microelectrode is seen in the middle image with an arrow pointing to the edge of the polypropylene insulation. The polymer layer appears thin and uniform over the wire. Figures 2.3D and E show top view images of the diamond film with and without the polypropylene coating. The microelectrode diameter at the narrowest point was 10-20 pm and at the widest point was about 80 pm. The length of the exposed electrode was 100- 200 pm. 44 A Conductive Epoxy Tip of Diamond 200 um Figure 2.3. (A) Diagram of the conically shaped diamond microelectrode insulated with polypropylene. SEM images of the polypropylene-insulated diamond microelectrode at (B) lower and (C) higher magnification. Top-view SEM images of the diamond film morphology without (D) and with (E) the polypropylene insulation layer. 45 Application of this polypropylene insulation method is quite reproducible in terms of coating the diamond microelectrode with a thin and continuous polymer propylene film, but precisely control of the exposed electrode length is difficult to achieve. 2.4 Electrochemical Measurement Cyclic voltammetric measurements were made with a CS-1200 computer- controlled potentiostat (Cypress Systems Inc., Lawrence, KS) in a three-electrode configuration using a single compartment glass cell [153]. The cell volume was approximately 5 mL. A carbon rod served as the counter electrode and a commercial Ag/AgCl electrode (3 M KCl, Cypress Systems Inc.) was used as the reference. All measurements were made in a Faraday cage at room temperature and all solutions were deoxygenated with N2 for at least 10 min prior to a measurement. Both microelectrode types were cleaned by rinsing and soaking with distilled isopropyl alcohol [167]. The in vitro continuous amperometric measurements were made using an Omni 90 analog potentiostat (Cypress Systems Inc.). An analog-to-digital converter (Labmaster 125) and a computer running Axotape software (version 2.0, Axon Instruments, Foster City, CA) were used to record the current-time profiles. Cyclic voltammetry. Cyclic voltammetry provides information not only on the thermodynamics of redox reactions but also on the kinetics of heterogeneous electron- transfer reactions and coupled chemical reactions. Routinely, voltammetric experiments are performed using a stationary working electrode in a quiet (unstirred) solution containing a dissolved redox-active molecule. A three electrode set-up is most often used consisting of a working, auxiliary and reference electrode. The working electrode is the 46 one at which the reaction of interest is studied. The properties and placement of the auxiliary and reference electrodes are critical and must be carefully considered for appropriate experiment design. The working electrode serves as a source or sink for electrons. Starting from an initial potential, Bi, 3 linear potential sweep (potential ramp) is applied to the working electrode. After reaching a switching potential, E1, the sweep is reversed and the potential returns linearly to its original value. The potential of the working electrode is controlled versus an appropriately placed reference electrode such as a silver/silver chloride electrode (Ag/AgCl). The value of this potential relative to the standard reduction potential, E°, for the redox molecule represents the driving force for Faradaic charge-transfer reactions. If the potential is negative of E° then the reduced form of the redox-active molecule is stable at the interface, and if the potential is positive of E° then the oxidized form is favored. The controlled potential can be considered the excitation signal in the measurement. The Faradaic current that flows at any time between the working and auxiliary electrodes is a direct measure of the rate of the electrochemical reaction taking place at the electrode surface. In cyclic voltammetry, mass transport of the redox-active molecule generally occurs solely by diffusion. At conventionally-sized electrodes, a peak current will be seen in the i-E curve for the forward cycle if the reaction rate is limited by diffusional mass transport. The peak current is given by the Randles-Sevcik equation 1', = (2.69 x 105) nmADl/ZC u‘” where ip is the peak current (A), n is the electron stoichiometry, A is the electrode area (cmz), D is the diffusion coefficient (cmz/s), C is the redox molecule concentration 47 (mol/cm3) and u is the potential sweep rate (V/s). Clearly, ip varies proportionally with A, C and pm. In contrast, a steady-state current is observed in the i-E curve for microelectrodes with a curve shape that depends on the geometry of the exposed microelectrode. In other words, the shape of the electrode determines the reactant flux (molecules/cmz-s), which, in fact, determines the current profile. The current at the carbon fiber is described by simply a cylindrical term. On the other hand, since the diamond microelectrode has both a cylindrical and conical shape exposed, the current is described by two terms according to the equations below [155, 159] iCylinder _ anADC SS r In t if“ = 4nFDcr(l + qH”) where n is the number of electrons, F is the Faraday constant, A is the electrode surface area, D is diffusion coefficient, c is bulk concentration of a compound, r is cylinder or cone radius, r = 4Dt/r2 , H is the aspect ratio h/ r , where h is height of the cone, q = 0.30661 and p = 1.14466. The aspect ratio was determined from SEM measurements. The exposed microelectrode area was calculated according to the equations. Electrochemical characterization of the diamond microelectrode. Cyclic voltammetry was used to evaluate the basic electrochemical properties of the diamond microelectrode. Comparison measurements were made with a conventional carbon fiber microelectrode. Figure 2.4A shows background cyclic voltammetric i—E curves in 1.0 M HClO4 for a typical diamond and carbon fiber microelectrode, and for a Pt wire incompletely coated 48 with a diamond overlayer. The area exposed to the solution was different for each electrode (and not measured), which makes direct comparison of the current magnitudes impossible. In general, though, diamond exhibits a lower background current and a wider working potential window than does the carbon fiber [155, 159]. For example, the background current density for diamond is typically a factor of 5—10 lower than that for a carbon fiber. This inherently leads to improved signal-to-background ratios in electroanalytical measurements. The background current for diamond results primarily from the charging of the electric double layer (i.e., potential dependent solvent dipole reorientation and ion movement at the interface). There is little contribution to this current from electroactive surface carbon-oxygen functionalities, as is the case for most sp2 carbon electrodes [104, 168, 139]. For the most part, the voltammogram for diamond is flat and featureless over the potential range, and is stable in shape with extended cycling. If defect (cracks and pinholes) exist in the diamond overlayer, then the solution will eventually reach the Pt substrate and electrochemical features characteristic of this metal will appear in the voltammogram. An example of this is the voltammetric i—E curve for the incompletely covered Pt wire. Pt is an active electrode for hydrogen evolution and this reaction is evidenced by the sharp increase in cathodic current at —250 mV. The fact that little current for hydrogen evolution flows until ca. —1500 mV for the diamond microelectrode indicates that the diamond film completely covers the Pt substrate. The kinetic overpotential for this reaction on diamond microelectrode is similar to what has been reported for high-quality and well characterized microcrystalline diamond thin-film electrode formed on Si [146, 153, 148]. 49 B —L C? . Current (nA) O 10- .14? D! ’“. ' r l f r ' IL? I - r w I ' -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 WNW-Adm) 3(1) . —Dianond 200. --Ca'bonFibar < 5 3100 2 '5 0 0. 4(1) 3.........-. -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 WNvaAdAfi) Figure 2.4. Background cyclic voltammetric i —E curves in (A) 1.0 M HClO4 for a diamond (—), carbon fiber (— —) and Pt-exposed diamond microelectrode (w). Cyclic voltammetric i —E curves for a diamond and carbon fiber microelectrode in (B) 1.0 mM Fe(CN) (3H in 1 M KCl. Scan rate=100 mV/s. 50 Another interesting and important feature of the diamond microelectrode response is the absence of peaks between the working potential limits. In the case of the carbon fiber microelectrode, an oxidation peak is present at about 400 mV and a corresponding reduction peak is seen at 250 mV. These peaks are associated with redox-active surface carbon—oxygen functionalities, specifically, the quinone/hydroquinone type [169, 166, 167]. Carbon electrodes, particularly 3p2 materials, possess a variety of carbon—oxygen functional groups at the exposed edge plane sites. Some of these functional groups are electroactive (e.g., quinone/hydroquinone) and some are not [170]. Other types are acidic and ionizable (e.g., R—COOH). In the case of the electroactive type, these functional groups undergo electron-transfer in the —0.2 to 0.6 V vs. Ag/AgCl potential window, at this pH, according to 2H+/2e’ redox reaction. The potential region in which this charge— transfer reaction takes place, as well as the extent of functional group deprotonization, depends on the solution pH, which can change during an in vitro or in vivo measurement [139, 140]. Both types of functional groups can be problematic in in vitro or in vivo electroanalytical measurements because (i) they affect the electrode surface charge, which influences the concentration of catechol and catecholamine molecules (cationic, neutral and anionic) at the surface, and (ii) they undergo electron-transfer in a potential region that overlaps that for the catecholamine neurotransmitters being detected. In vitro and in vivo voltammetric measurements often require background subtraction to enable the identification and quantification of endogenous catecholamine neurotransmitters. The presence of electroactive surface functional groups complicates this background correction [139]. The diamond microelectrode does not possess surface oxides when 51 hydrogen-terminated and thus is ideal for these measurements because the background voltammetric signal is featureless and independent of pH [155, 159]. Figure 2.4B shows cyclic voltammetric i—E curves for 1 mM Fe(CN)6‘3/‘4 in l M KCl at both a diamond and carbon fiber microelectrode. Fe(CN)(,'3/’4 is a good redox system to use for testing the electrochemical behavior of both diamond and 3p2 carbon electrodes [171-173]. In general, the heterogeneous electron-transfer rate constant for this redox system is highly sensitive to the surface cleanliness, exposed microstructure and the density of electronic states near the formal potential [174, 175, 148]. The voltammetric i—E curves for Fe(CN)6’3/‘4 are nearly identical in shape for both electrode types. The half-wave potentials (Em) are 267 and 269 mV, respectively, for the diamond and carbon fiber microelectrode. In general, the Em values for multiple diamond and carbon fiber microelectrodes tested were quite similar indicating that both electrodes are active for this redox system. This result indicates that (i) the response of diamond toward this redox system is similar to that for the carbon fiber, (ii) the diamond surface is effectively cleaned, when needed, by the alcohol soak and (iii) the diamond electrode possesses a high density of electronic states in this potential region enabling relatively rapid electron-transfer kinetics. 2.5 Animals The University Committee on Animal Use and Care at Michigan State University approved all procedures for handling and caring for the laboratory test animals. All rats were allowed 2-3 days of acclimatization prior to entry into any experimental protocol. Pelleted rat chow (Harlan/Teklad 8640 Rodent Diet; Harlan/Teklad) and water were 52 provided. Rats were housed in temperature and humidity controlled rooms using a 12-h on/off light cycle. Preparation of DOCA-salt and sham Rats. Male Sprague-Dawley rats (175-200 g, Charles River Inc., Portage, M1) were exteriorized and the left kidney was removed under anesthesia with sodium pentobarbital (50 mg/kg i.p.(intraperitoneal)) after ligation of the renal artery, vein and ureter with 4-0 silk sutures. DOCA implantation was made between the back shoulder blades of the animal using a 1 cm incision. DOCA implants (600 mg/kg) were prepared by mixing deoxycorticosterone acetate in silicone rubber resulting in a dose of 200 mg/kg. DOCA-implanted rats received standard pelleted rat chow and salt water (1% NaCl + 0.2% KCl). Sham, normotensive rats also underwent left kidney removal but no DOCA pellet implantation and received unsalted tap water. The surgery was performed on a heated pad and rats recovered in a heated box. Antibiotics (enrofloxacin, 5 mg/kg s.c.(subcutaneous)) and an analgesic (butorphanol tartrate, 2 mg/kg, so.) were administered immediately after the surgery. After recovery, the rats were housed under the above listed conditions for 4 weeks. Systolic blood pressure was measured using the tail-cuff method four weeks after surgery. Rats with mean systolic blood pressure of 2 150 mmHg were considered hypertensive [176, 31]. 2.6 In Vitro Electrochemical and Diameter Measurement System The combined use of continuous amperometry with diamond and carbon fiber microelectrodes and video imaging were used for the simultaneous in Vitro measurement of NE release from sympathetic nerves innervating a rat mesenteric artery (MA) and vein (MV) and the evoked contractile response. 53 These techniques, along with various drugs, were employed to show the relationship between the oxidation current associated with endogenous NE overflow and blood vessel constriction. A block diagram of the experimental set-up is shown in Figure 2.5. Mesentery (Small Intestine) . 3" i'i Blood Vessel ’° Diameter I 180-330 um Computer Q Buffer °nunl 35-37 °C Oxygenated Krebs’ Buffer Flow rate: 5~6 mL/min I Video Camera Figure 2.5. Block diagram of the experimental set-up Tissue preparation. The sham and DOCA-salt rats (300 — 430 g,) were euthanized with a lethal pentobarbital injection (50 mg, i.p.). The abdomen was surgically opened, and the small intestine carefully removed and placed in an oxygenated (95% 02, 5% C02) Krebs’ buffer solution (pH 7.4) of the following composition: 117 mM NaCl, 4.7 mM KCl, 2.5 mM CaClz, 1.2 mM MgC12, 25 mM NaHCO;, 1.2 mM NaH2P04, and 11 mM glucose. A section of the mesentery close to the ileal wall was carefully cut free from the intestine 54 and transferred to a small silicone elastomer-lined, Teflon flow bath (5 mL volume). Secondary or tertiary arteries and veins (180-330 um outside diameter) were isolated for in vitro study by carefully removing the surrounding adipose and connective tissue under a dissecting microscope using fine scissors and forceps. The Krebs’ buffer solution entered the bath at one end (center) and flowed out the opposite end (center). The bath, containing the fixed tissue preparation, was attached to the stage of an inverted microscope (Olympus CK-2) and superfused continuously with 37 °C Krebs’ buffer at a flow rate of 1.6 mL/min (Chapter 3 and 4) and 5-6 mL/min (Chapter 5 and 6). The solution flow was controlled by a peristaltic pump. Assessment of electrode performance. The diamond and carbon fiber microelectrodes were cleaned by soaking in distilled isopropyl alcohol (IPA) for at least 15 min prior to a measurement [167]. The microelectrode was affixed to a micromanipulator (MP-1, Narishige Instruments, Japan), which was used to position the electrode on the surface of the MA and MV. The microelectrode was positioned in the center of the bath containing the fixed tissue preparation, equidistant from the inlet and outlet solution ports. A Pt wire counter and a commercial “no leak” Ag/AgCl (3 M KCl, model EE009, Cypress Systems Inc.) reference electrode were also mounted in the bath to complete the electrochemical cell. All electrochemical measurements were made with an Omni 90 analog potentiostat (Cypress Systems Inc.), an analog-to-digital converter (Labmaster 125), and a computer running Axotape software (version 2.0, Axon Instruments, Foster City, CA). The software was used to record the continuous amperometric current-time profiles and blood vessel diameter changes. The analog output current from the potentiostat was low pass filtered (5 Hz or 200 ms RC time constant). All data were digitized using a sampling rate 55 of 100 Hz. The Krebs’ buffer flowed over the electrode and the tissue sample for 30 min prior to the start of a series of measurements. Focal stimulation of perivascular nerves. Short trains of electrical stimulation were used to trigger NE release. Perivascular nerves were stimulated using a bipolar focal stimulating electrode positioned on the surface of the blood vessel at a distance of ca. 200 pm from the tip of the carbon fiber. This positioning minimized the noise introduced into the current recordings. The focal stimulating electrode consisted of two AgCl-coated Ag wires inserted into double barrel capillary glass (tip diameter = 180 pm). The wires were connected to a stimulus isolation unit and a Grass Instruments stimulator (S88, Quincy, MA). Trains of 60 stimuli (0.3 ms duration, 30-70 V) at frequencies ranging between 0.5 and 30 Hz were used. In vitro video monitoring of vasoconstriction. The output of a black and white video camera (KP-111, Hitachi, Yokohama, Japan) attached to the microscope was fed to a PC Vision Plus frame-grabber board (Imaging Technology, Wobum, MA) mounted in a personal computer. Changes in blood vessel diameter of 0.1 pm could be resolved. Video images were analyzed using Diamtrak software (http://www.diamtrgk.com, Adelaide, Australia). The digitized signal was converted to an analog output (DAC-02 board, Keithley Metrabyte, Tauton, MA) for subsequent processing by an analog-to- digital converter (Labmaster 125) and analysis in a second computer running Axotape for a permanent recording of blood vessel diameter as function of time. Analog signals were sampled at 100 Hz and data were stored on the computer’s hard drive for subsequent analysis and display. 56 2.7 Fluorescence Histochemistry Tissue preparation. Second or third-order veins and arteries from the ileal mesentery, were isolated by carefully clearing away the surrounding adipose and connective tissues. Red blood cells were flushed out of the blood vessel lumen with phosphate buffer (0.1 M, pH 7.2) using a 30 ga. hypodermic needle and syringe, and tissues were cut to 1 cm lengths. For these experiments, we used more than 30 MA and MV segments from 13 different rats. Gyloxylic acid fluorescence histochemistry. Catecholamine fluorescence was revealed after incubating the prepared tissues in a 2% glyoxylic acid/0.2 M phosphate buffer (pH 7.0) solution for 5 minutes at room temperature. The blood vessels were stretched onto glass slides and then heated at 80 °C for 5 min. Preparations were mounted in mineral oil and observed using a fluorescence microscope (Nikon Eclipse TE 2000-U) equipped with a filter set, UV2E/C (excitation filter, 340-380 nm and emission filter, 435-485 nm). 2.8 Chemical and Drug Application All chemicals were reagent-grade quality, or better, and used without additional purification. The chemicals and drugs used for electrode performance testing were postassiurn ferrocyanide (II) trihydrate, epinephrine (EE), dopamine (DA), norepinephrine (N E), catechol (CA), t-butyl catechol (TBC), 4-methyl catechol (4MC) and 3,4-dihydroxyphenylacetic acid (DOPAC). Tetrodotoxin (TTX, 0.3 uM), pyridoxal- phosphate-6-azophenyl-2',4'-disulfonic acid (PPADS) (10 uM), prazosin (0.1 uM), yohimbine (1.0 uM), UK 14,304 (1.0 uM) and cocaine (10 uM) were used for in Vitro study. The drugs were added to the superfusing Krebs’ buffer solution. Drugs were 57 applied for 20 min prior to assessing their effects on nerve-mediated responses and were applied after a series of control measurements were made. All chemicals and drugs were obtained from Sigma Chemical Company (Saint Louis, MO). Ultrapure water (distilled, deionized, and passed over activated carbon, 17-18 MQ, Barnstead E-PureSystem) was used for all solution preparation, and glassware and electrode cleaning. 2.9 Data Analysis Data were obtained from >150 sham rats and >150 DOCA-salt rats for the research. Mean systolic blood pressure from sham and DOCA-salt rats was 125 mm Hg and 194 mm Hg, respectively. The mean weight of sham rats was 416 g, and the mean weight of DOCA-salt rats was 336 g. Each series of measurements required at least 3 rats. pCLAMP 8.1 software (Axon Instruments, Foster City, CA) was used to analyze all data. The rise time and decay of NE signals (S/N > 3) or blood vessel constriction were fitted using a standard exponential model in Clarnpfit 8.1. Data are presented as mean i S.E.M; where SEM is standard error of the mean. Means were compared by using Student’s t- test and P < 0.05 was regarded as significant. MA and MV constrictions are expressed as a percentage of the initial resting diameter (in pm) of the blood vessel. For nerve stimulation frequency-response curves, the half maximal effective stimulation frequency (Sso) and maximum constriction (Emu) were calculated from curves obtained in individual tissues using the following function: Y=EmaxX/(S50+X) where X is the stimulation frequency tested and Y is the peak response amplitude. 58 Chapter 3 Comparison of Electrochemical Properties of Diamond and Carbon Fiber Microelectrodes During Exposure to Biological Environments 3.1 Introduction An electrochemical electrode used in biological media should possess several properties: good sensitivity for the target analyte, rapid response time, a stable background response that is unaffected by changes in the surrounding solution environment (e.g., pH) and resistance to biomolecule adsorption (deactivation and fouling). Carbon fibers are useful for monitoring small quantities of various electroactive catecholamines because of their small size (<30 um) and temporal response [103, 177, 178, 139, 140]. However, a drawback with their use for measurements in vitro and in vivo is the lost response sensitivity and poor stability that typically occurs during exposure to physiological environments. Specifically, background current features (e.g., redox active functional groups) and drift, sensitivity to local pH changes and response attenuation due to surface fouling caused by biomolecule adsorption are complications that arise. The cause for these complications is linked to the presence of carbon—oxygen functional groups on the electrode surface [104, 139]. Conductive diamond is a new electrode material with an sp3-bonded carbon microstructure with no extended n-electron system, and little surface oxygen due to a hydrogen termination. As will be shown in this Chapter, the diamond microelectrode has properties that make it useful for measurements in complex biological environments. Ideally, one would like to have a microelectrode that is void of surface carbon—oxygen functional groups and that possesses a surface on which 59 weak molecular adsorption occurs. Planar diamond thin-film electrodes have attracted considerable interest in recent years due to some superb electrochemical properties [179, 153, 180, 148]. However, so far, there have been only a few reports describing the electrochemical properties and application of diamond microelectrodes [155, 156]. Such electrodes are typically prepared by coating a thin layer of conducting diamond on a small diameter metal wire. The diamond microelectrode is attractive for electroanalytical measurements in complex media, like biological environments, because of its (i) hard and lubricious nature that enables easy penetration into tissue with minimal peripheral damage, (ii) low and stable background current over a wide potential range, (iii) good chemical and microstructural stability, (iv) low surface oxygen content (when H- terrninated) - a property that leads to minimal change in the background current with variation in solution pH, (v) chemical inertness and (iv) a non-polar, hydrophobic surface that renders it resistant to molecular adsorption (i.e., deactivation and fouling). Another purported property of diamond is biocompatibility. Diamond electrodes are presumed to be highly resistant to deactivation and fouling when exposed to complex biological environments; however, to the best of our knowledge, there has been no systematic study of the response sensitivity, precision and stability in such environments. How stable and resistant to deactivation this new microelectrode is during exposure to biological environments and how useful it is for in vitro electroanalytical measurements are two questions this work sought to answer. An objective for this work was to evaluate the diamond microelectrode response for endogenous NE in vitro and to compare the response with that of a conventional carbon fiber microelectrode [181, 34, 104]. Preparations of rat mesenteric arteries 60 maintained in vitro were used as a test system to make this comparison. It has previously been shown that electrical stimulation of these tissues causes the perivascular nerves to release NE [119, 121]. We report presently on the response sensitivity, precision and stability of a diamond microelectrode for (i) NE and Fe(CN)6’3/‘4 before and after exposure to the laboratory atmosphere and to biological tissue, and (ii) NE during in vitro release measurements at the surface of a test animal’s mesenteric artery. 3.2 Results and Discussion Effect of pH on the background voltammetric response. Figures 3.1A and B show background voltammetric i—E curves for a diamond and carbon fiber microelectrode in phosphate buffer solutions ranging in pH from 3 to 7. The i—E curves for the diamond microelectrode between —300 and 500 mV are all featureless and the current magnitude is the same at all pH values. The current beyond 600 mV is presumably due to the onset of oxygen evolution, which increases as expected with increasing solution pH. The calculated geometric areas of the diamond and carbon fiber microelectrode were 6.2 x 10'5 and 1.5 x 10’5 cm2, respectively. Even though the diamond microelectrode area is approximately four times larger, it exhibits a lower background current. The low background current is a characteristic feature of diamond and leads to improved signal- to-background ratios in electroanalytical measurements [179, 153, 180, 148]. The electrode capacitance was not measured directly but the values calculated from the cyclic voltammetric anodic current at 0 V were 11 and 200 uF/cmz, respectively, for the diamond and carbon fiber. 61 Q: 04 am -03 pI-Iso gag: #450 Eat mm 5°". - o0m “’ .02. 3‘ an ' I ' r v r v u v r 400 -200 0 200 400 am 800 mart/unmet) B p 9 - Current (nA) is . 04- 400-200 0 "zin'400'a'n'800 warmqugo) Figure 3.1. Background cyclic voltammetric i—E curves for a (A) diamond and (B) carbon fiber microelectrode in 0.1 M phosphate buffer solutions of different pH ranging from 3 to 7.2. Scan rate = 100 mV/s. 62 The assumption in this calculation is that the background current is all capacitive in nature. For the diamond electrode, this is likely the case, but for the carbon fiber, the background current contains a significant Faradaic contribution from the electroactive surface oxide functionalities. Therefore, the electrode capacitance for the carbon fiber calculated in this manner is much larger than the true value (expected to be approximately 30 uF/cmz). The voltammetric i—E curves for the carbon fiber have redox waves in the 200 — 600 mV range that shift with pH. Specifically, the peaks shift by —59 mV/pH, consistent with an equal number of electrons and protons being transferred in the redox reaction. These peaks are attributed to redox-active carbon—oxygen functionalities on the electrode surface that exist at edge plane and defect sites. It is generally accepted that the redox-active groups are of the quinone/hydroquinone type [170, 130]. Carbon electrode surfaces, particularly 3p2 materials possess a variety of oxygen functional groups, some electroactive and some not. A non-electroactive carbon—oxygen firnctional group that is sensitive to pH is the carboxylic acid type. Negative excess surface charge will exist on the electrode when these groups are deprotonated at pH values above the pKa, which is near 4.5. Increasing the excess surface charge will alter the ionic charge stored in the electric double layer and is reflected by an increasing background voltammetric current. These redox peaks and changes in the background D , rrent with pH are problematic in in vitro and in vivo electroanalytical measurements, particularly for catecholamine analysis. This is because they occur in a potential range that overlaps the oxidation and reduction peaks for the catecholamines. The fact that background current varies with pH complicates background subtraction/correction [104, 139, 140]. The pH changes occur in biological tissue as a result of coupling respiration 63 and energy production to pH [104, 139, 140]. The diamond microelectrode does not possess surface oxides when hydrogen-terminated, therefore, there are no redox waves and the background is relatively insensitive to changes in solution pH at constant ionic strength. Catecholamine voltammetric response. Figures 3.2A and B show cyclic voltammetric i—E curves for 10 uM norepinephrine (NE) and epinephrine (BE) in 0.1 M phosphate buffer (pH 7.2) at the two microelectrode types. The calculated geometric areas of the diamond and carbon fiber microelectrode were 2.4 x 10'3 and 1.3 x 10‘3 cm2, respectively. Both curves have some peak-shape character at this scan rate, which reveals that a steady-state flux of analyte to the electrode surface is not fully achieved on the time scale of this measurement. Rather, a thin depletion layer develops leading to a time- dependent flux. We estimate, based on the 80 um diamond microelectrode diameter, that scan rates less than 50 mV/s would be needed to observe a steady state oxidation current. There may also be some molecular adsorption particularly on the carbon fiber, which would contribute to the peak-shape character. We did not conduct any electroanalytical measurements to probe for molecular adsorption though. The maximum oxidation peak currents for both NE and EE at diamond are 13 nA and at the carbon fiber are 8 nA. The factor of 1.6 difference in current is comparable to the factor of 1.8 difference in the calculated area for the two microelectrodes. The oxidation half-wave potentials (El/2) for NE and EE at the diamond microelectrode are approximately 200 mV more positive than the values for the carbon microelectrode. 64 $-23-“ Current (nA) - 5 -9.‘." -s 400200 0 '200'400'600'80011m0 marmvuwgcr) $3238 Current (nA) 9-9‘- 2m0m2014006008t'l) meantime/rumor) Figure 3.2. Cyclic voltammetric i—E curves for a diamond and carbon fiber microelectrode in (A) 10 M NE and (B) 10 uM EE, both in 0.1 M phosphate buffer, pH 7.2. Scan rate = 100 mV/s. 65 This is a common observation for catecholamine electrochemistry at diamond with the more positive Em being attributed to more sluggish reaction kinetics, not ohmic resistance effects [146, 153, 180, 154]. The reason for the more sluggish kinetics at diamond is still under study but one postulate is that molecular adsorption is not involved in the reaction mechanism on the hydrogen-terminated, sp3-bonded carbon surface. Of particular relevance here is the fact that McCreery and DuVall have shown that catecholamines adsorb strongly on GC and have correlated the adsorption with more rapid electrode reaction kinetics [182]. Importantly from an electroanalytical point of view, the more positive detection potential for diamond does not appear to be a major drawback as the background current remains low and the physicochemical properties remain unchanged. A summary of the cyclic voltammetric Em values for several catechols and catecholamines at the two microelectrode types (n 2 4) is presented in Table 3.1. The analyte concentrations were all 0.1 mM. Consistently for all the analytes, the Em for diamond is 150—250 mV more positive than for the carbon fiber. The response precision is quite good for both electrode types with a relative standard deviation (RSD) of l—4%. The results indicate that the diamond microelectrode can be formed with reproducible electrochemical properties. Diamond electrode response stability during exposure to the laboratory atmosphere. A series of measurements was made to assess the stability of the diamond microelectrode response during exposure to the laboratory atmosphere. This is a good test of stability because carbon electrodes are notoriously prone to deactivation during exposure to air [170]. 66 Table 3.1. Summary of cyclic voltammetric El,2 data for diamond and carbon microelectrodes. Analyte Electrode Diamond E1 /2 (mV) Carbon Fiber Em (mV) Dopamine 421 d: 17 188 i 3 Norepinephrine 456 i 12 236 :l: 10 Epinephrine 472 i 18 266 :t 14 Catechol 543 i 8 291 :1: 12 t-Butyl catechol 497 :l: 7 218 :t 5 4-Methyl catechol 495 d: 6 241 :E 8 DOPAC 614 :t 19 352 i 11 The solutions were 0.1 mM norepinephrine (NE), epinephrine (EE), 3,4-dihydroxy- phenylacetic acid (DOPAC), dopamine (DA), catechol (CA), t-butyl catechol (TBC), and 4-methyl catechol (4MC) all in 0.1 M phosphate buffer (pH 7.2). Scan rate = 20 mV/s. Furthermore, the electrode response sensitivity and reproducibility can vary considerably depending on the electrode preparation conditions, the storage conditions and the past history of use. Figure 3.3A and B show cyclic voltammetric i—E curves for a diamond and carbon fiber microelectrode in 1 mM Fe(CN)(,"3/'4 and 0.1 M KCl before and after a multiple-day exposure to the laboratory atmosphere. Fe(CN)t,-’3/’4 is a good redox system for this kind of test because the electrode reaction kinetics are quite sensitive to the surface cleanliness of carbon electrodes [174]. The kinetics are also sensitive to the surface microstructure of 3p2 carbon electrodes [174, 17 5] and to the surface oxygen content on diamond [172]. The electrode reaction kinetics are influenced by the surface oxygen content on sp2 carbon electrodes, but mainly when a thick oxide layer exists [175]. Both electrodes were initially cleaned by rinsing and soaking in distilled isopropanol for 20 min [167]. 67 Current (nA) Current(nA) 9.9.8833ggg 400-2110 0 2000400600 80001000 warms. range) Figure 33. Cyclic voltammetric i-E curves for 1 mM Fe(CN)6’3/'4 in 0.1 M KC1 at (A) diamond and (B) carbon fiber microelectrode before and after laboratory atmosphere exposure. Scan rate = 100 mV/s. 68 It is important to note that while this treatment was applied to diamond for consistency, it was seldom needed as the “as grown”, untreated electrode typically exhibited an active response for this surface-sensitive redox system [153, 154]. The calculated geometric areas for the diamond and carbon fiber microelectrodes were 1.8 x 10'4 and 9.7 x 10'4 cm2, respectively. Well-defined oxidation peaks are seen for both electrodes with the initial Em of 260 and 282 mV, respectively, for the diamond and carbon fiber microelectrode. After a two-week exposure to the laboratory atmosphere, the diamond microelectrode response was largely unaltered. No pretreatment (e.g., cleaning with alcohol) was applied to either electrode after the air exposure. A pseudo-sigrnoidal oxidation current is still observed with the maximum current remaining about the same and Em shifting positive by only 7 mV. This indicates that the relatively non-polar, hydrogen-terminated diamond surface is resistant to deactivation in the laboratory atmosphere (e.g., surface oxidation and adsorption of airborne contaminants). A pseudo-sigrnoidal oxidation current is also observed for the carbon fiber microelectrode, but the Em is shifted positive by 138 mV after a one-week exposure. The positive shift was progressive with time and reflects more sluggish electrode reaction kinetics, presumably because of deactivation by contaminant adsorption [170]. The maximum oxidation current of 52 nA for diamond and 240 nA for the carbon fiber differ by a factor of approximately 5, and this is consistent with the factor of 5 difference in geometric area. Clearly, the diamond microelectrode is more resistant to deactivation in the air than is the carbon fiber. The extent to which the physical, chemical, and electronic properties affect the electrochemical response of 3p2 (e.g., carbon fiber) and 3p3 (e.g., diamond) carbon 69 electrodes depends on the mechanism for the particular redox test system [175, 146, 182, 153]. Some redox systems are more outer-sphere than others in terms of their electron- +3l+2 , methyl viologen, and IrClrg‘ZJ'3 fall into transfer mechanism. For example, Ru(NH3)6 this category for both GC and diamond electrodes [175, 146, 182, 153]. For such redox systems, the apparent heterogeneous electron-transfer rate constant, k°app, is relatively unaffected by the physical and chemical properties of the electrode surface. The primary factor influencing k°app is the electronic properties of the electrode — the density of electronic states near the formal potential of the redox system. On the other hand, some _ .4 . . . 3’ , catecholamines, ascorbic acrd, and 02, are more redox systems, such as Fe(CN)6 inner sphere in terms of their electron-transfer mechanism [175, 146, 182, 153]. For such systems, k°app is highly sensitive to the surface cleanliness, microstructure and chemistry, as well as the density of electronic states near the formal potential. As mentioned above, F e(CN)6"3/ '4 is particularly sensitive to the surface cleanliness and exposed microstructure (i.e., fraction of clean edge plane) of 3p2 carbon electrodes [175, 146, 182, 153], and to the presence of surface oxygen functionalities on diamond electrodes [172]. Therefore, F e(CN)6‘3’ 7‘ is a very good probe of a carbon electrode’s response sensitivity and stability. Diamond electrode response stability during exposure to biological tissue. Another series of measurements was performed to learn how exposure to biological tissue affects the diamond microelectrode response. A concern with any electroanalytical measurement in biological environments is response attenuation caused by the irreversible adsorption of biomolecules (e.g., proteins, lipids). Fouling of oxygenated sp2 carbon electrodes can occur in biological media because of strong molecular adsorption on the polar, 70 hydrophilic surface. One of the ways fouling of 3p2 carbon fiber microelectrodes can be minimized is by coating with the ionomer, NafionTM [183]. While this treatment improves the electrode sensitivity for cationic neurotransmitters and renders the electrode more resistant to fouling, it also causes an increase in the electrode response time to concentration changes brought about by a '3/"4 was again used as the redox test neuronal event (in vivo or in vitro) [131]. Fe(CN)6 system to probe the bare diamond electrode response sensitivity and stability. The test involved first recording a cyclic voltammetric i—E curve for 1 mM Fe(CN)6‘3/“4 in 0.1 M KCl. The microelectrode was then removed from the standard electrochemical cell and transferred to the flow bath containing a mesenteric artery preparation in which most of the surrounding adipose tissue had not been removed. The microelectrode was inserted into the tissue for 10 min before being removed and transferred back to the electrochemical cell containing the Fe(CN)6‘3"4 solution. The cyclic voltammetric response was then re-recorded. A bare carbon fiber microelectrode was also tested for comparison. The geometric areas for the diamond and carbon fiber microelectrode were 2.2 x 10"4 cm2 and 1.5 x 10’3 cm2, respectively. No soak in isopropyl alcohol was used to clean the electrode surfaces after the tissue exposure. Observation of the implanted microelectrodes with an optical microscope revealed that the hardness and lubricious nature of diamond allowed it to smoothly penetration the tissue without causing significant peripheral damage. On the other hand, the flexibility of the carbon fiber microelectrode made insertion difficult without bending or breaking. In these measurements, the carbon fiber microelectrode did not actually penetrate the tissue but merely was in contact with it on the outside surface. 71 Figures 3.4A and B show cyclic voltammetric i—E curves for 1 mM Fe(CN)6_3/T1 in 0.1 M KCl at both microelectrode types, before and after the tissue exposure. The curves for both electrodes before exposure possess some peak-shape character with E”; values of 262 and 333 mV, respectively, for the diamond and carbon fiber. The maximum oxidation current was 60 nA for diamond and 360 nA for the carbon fiber. The factor of 6 difference in current is consistent with the factor of 7 difference in the calculated geometric area. After exposure to the tissue sample, the response for the diamond microelectrode was slightly affected as there was only a 21 mV positive shift in Em with the oxidation current maximum remaining about the same. In contrast, a 205 mV positive shift in Em was seen for the carbon fiber along with a slight decrease in the maximum current. It appears that biomolecule adsorption occurs to a greater extent on the polar, oxygenated carbon fiber surface even though the electrode did not actually penetrate the tissue. Molecular adsorption deactivates the electrode for this redox system by partially blocking some specific surface sites involved in the reaction. Furthermore, it was '3” could not be regained observed that the original carbon fiber response for Fe(CN)6 after soaking in clean isopropanol for 20 min. This indicates that the biomolecule adsorption is very strong on the it-bonded, oxygen-containing carbon fiber surface. In contrast, the original diamond microelectrode response, which was only weakly affected by the tissue exposure anyway, was regained after a 20 min isopropanol soak. This is consistent with weak biomolecule adsorption, as expected. The diamond and carbon fiber microelectrode response sensitivity and stability after tissue exposure was also probed using NE. 72 5-8-8 Current (nA) 9- B - 1'3- 4001200'0'200'400'600'800 Whit/BMW) Current (nA) ¢. 8 - § § § I E I l 500 ' 0 ' 500 ' 10'00 murmnmmgo) Figure 3.4. Cyclic voltammetric i—E curves for 1 mM Fe(CN)6'3/'4 in 0.1 M KCl at a (A) diamond and (B) carbon fiber microelectrode before and after exposure to biological tissue. Scan rate = 100 mV/s. 73 The diamond microelectrode is being used for the in vitro measurement of NE released in the vasculature of normal and hypertensive test animals. Therefore, it was important to learn how tissue exposure affects the short and long-term electrode response for this neurotransmitter. Measurements of NE were made by continuous amperometry using the previously described flow bath and involved first placing the microelectrode in the bath above the tissue. An NE solution was then flowed through the bath with the electrode poised at a potential positive enough to oxidize the catecholamine at a mass-transfer limited rate. The optimum detection potentials were determined from hydrodynamic voltammetric measurements to be 0.8 and 0.4 V, respectively, for the diamond and carbon fiber microelectrode. The electrode was inserted into the tissue for at least 10 min. The electrode was then removed from the tissue and a NE-containing solution injected into the flow bath. Figures 3.5A and B show the results of such a test. The calculated geometric areas were 6.4x 10’4 cm2 and 4.3x 10‘4 cm2 for the diamond and carbon fiber microelectrode, respectively. Again, the biological sample contained the blood vessels with the surrounding adipose tissue intact. The responses shown are for a 3 ml injection of 0.1 uM NE at a flow rate of 1.6 ml/min. Both bare electrodes showed some deactivation during the tissue exposure with the diamond microelectrode being less affected. For diamond, the nominal current response before exposure was 65.0 i 0.8 pA and after 8 h decreased to 36.3 i 0.7 pA (n 2 5). In contrast for the carbon fiber, the nominal current response was 41.4 i 1.6 pA before exposure and after 7 h decreased to 14.3 :t 0.5 pA (n _>. 5). This corresponds to a response attenuation of 44% for diamond and 66% for the carbon fiber microelectrode. 74 A At beginning [Diamond I 120 pA 8 h later 1'91? B At beginning [Carbon Fiber I 1:10 PA 7 h later c 1°". '\ imam» 804 i\{> a 60: \¢\§ t» l + \t :3 40 \+ TM (1‘) Figure 3.5. Continuous amperometric i—t responses for a (A) diamond and (B) carbon fiber microelectrode during multiple injections of 0.1 uM NE in Krebs’ buffer, pH 7.4 before and after exposure to tissue. The measurements were made in the flow bath. Injection volume = 3 ml. Flow rate = 1.6 mL/min. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCl. (C) Plot of the nominal current response as a function of time for the diamond and carbon microelectrodes. 75 The time dependence of the response attenuation is shown in Figure 3.5C. Greater and more rapid response attenuation is seen for the carbon fiber. Even though less extensive, the diamond microelectrode is deactivated some during exposure to the biological tissue, particularly when the adipose tissue is present. A major difference between the two electrode types, however, is the fact that the diamond response could be almost fully regained after a short soak in distilled isopropyl alcohol, while the carbon fiber response could not. This observation again indicates that weak biomolecule adsorption occurs on the diamond surface. Application of the diamond microelectrode for the in vitro detection of NE release from a mesenteric artery. It is possible to directly measure endogenous NE released from perivascular nerve fibers in real time when a microelectrode is positioned near a blood vessel. Such NE overflow from rat tail and small mesenteric arteries has been measured using carbon fiber microelectrodes [119, 121, 118]. In this work, a diamond microelectrode was used for the first time to measure NE release from a mesenteric artery after electrical stimulation. Much of the surrounding adipose tissue, in this case, was carefully removed from the blood vessel prior to electrode insertion. This permitted placement of the electrodes directly on the blood vessel surface in close proximity to the perivascular nerve fibers. The tissue was bathed with warm Krebs’ buffer for 1 h prior to the continuous amperometric measurements. Multiple diamond microelectrodes were used in these measurements with geometric areas ranging from 5.3 to 9.0 x 10‘4 cm2. Figure 3.6A shows a current—time profile before and after an electrical stimulation with the electrode potential poised at 0 mV. Clearly, no NE oxidation current is detected as the electrode potential is not positive enough to oxidize the released neurotransmitter. More 76 importantly, the electrical stimulation does not introduce noise into the electrode signal. Figure 363 shows a current—time profile before and after an electrical stimulation event, which reveals that the NE-release can be detected at 0.8 V; a potential that is sufficiently positive to oxidize NE at a mass-transfer limited rate. A series of NE-release events evoked by a train of 60 pulses at 20 Hz with a interval of 20 s between each train is shown in Figure 3.6C. Another series of release events with an interval of 10 min between each pulse train is presented in Figure 3.6D. Using a series of agonists and antagonists, we have determined that this current is associated primarily with the oxidation of NE released from sympathetic neurons. There could also be a contribution from the oxidation of the three major NE metabolites (norrnetanephrine, 3,4-dihydroxy-phenylethyleneglycol (DOPEG) and vanillyl mandelic acid (VMA)) as we have observed that all three undergo oxidation at the diamond microelectrode in the same potential range as NE. Clearly, the diamond microelectrode exhibits excellent reproducibility for the in vitro detection of NE. The nominal current response was 9.8 i 0.3 and 8.2 i 0.6 pA (n = 4) in Figures 3.6C and D, respectively. The response precision was good with RSD values ranging from 3 to 7%. Figure 3.7A and B show examples of the bare diamond and carbon fiber microelectrode response over several hours of in vitro NE release measurements. Current time profiles for a diamond (0.8 V, Fig. 3.7A) and a carbon fiber (0.4 V, Fig. 7B) microelectrode are presented. The calculated electrode areas were 9.0 x 10‘4 cm2 for diamond and 6.1 x 10‘4 cm2 for the carbon fiber. 77 A B 0mV mOmV Bewiedstimiaim 10 A [UH—Ina- 10 PAjw l—-l 103 103 5min 103 Figure 3.6. In vitro continuous amperometric i—t responses for a diamond microelectrode during measurement of NE-release from the mesenteric artery of a laboratory rat. The electrode response at (A) 0 V after electrical stimulation of the artery and (B) +0.8 V after electrical stimulation of the artery. The electrode response at +0.8 V after multiple electrical stimulation events with a time interval between each of (C) 20 s and (d) 10 min. The blood vessels were stimulated with the same pulse number and frequency (60 pulses at 20 Hz). Flow rate = 1.6 mL/min. 78 For diamond, the nominal current response at the beginning of the measurements was 16.9 i 0.4 pA and after 3 h the response was nearly unchanged at 16.1 i 0.5 pA (n = 3). This improved response stability, as compared to that shown in Figure 3.5, is due to the removal of the adipose and connective tissues surrounding the artery. In contrast, the nominal current response for the carbon fiber was 8.0 :I: 0.3 pA at the beginning of the measurements and after 1 h decreased to 7.1 :l: 0.3 pA (n = 3). The peak-to-peak noise for the diamond microelectrode remained constant over time (approximately 2.4—2.6 pA, Fig. 3.7(A)) while it varied somewhat for the carbon fiber ranging between 2.9 and 3.4 pA. The response attenuation for the diamond microelectrode was 5% over 3 h and for the carbon fiber was is 11% after just] h. The factor of 2.1 difference in nominal current for the two microelectrode types is slightly larger than the factor of 1.5 difference in calculated geometric area, therefore, we are probably underestimating the actual area. An additional reason for the difference in current could be the density of neurons, something that is variable along a blood vessel. The response stability of the two microelectrode types over a 4 h period of in vitro use is presented in Figure 3.7C. The NE oxidation current for the bare diamond decreased by only 8% over the period while the NE current for the bare carbon fiber decreased by about 30%. Clearly, the diamond microelectrode response is stable over several hours without the need for a protective ionomer coating (e.g., NafionTM). Diamond microelectrode response sensitivity for NE in vitro Calibration of a microelectrode response during an in vitro electrochemical measurement is a critical issue in neurochemical studies [131, 184]. 79 C 100» .\e*—O\ 90 \ <> _ 8° §\§ a 7°. \t .3 m o\° 50‘ —o—|jamnd « —O—Ca‘bon 40 30' . . . . . TII'I‘B (h) Figure 3.7. In vitro continuous amperometric i—t curves for a (A) diamond and (B) carbon fiber microelectrode during NE-release from a mesenteric artery as a function of time. The NE-release was evoked by electrical stimulation consisting of 60 pulses at 20 Hz. (C) Plots of the nominal current response as a function of time for the diamond and carbon fiber microelectrodes. Flow rate = 1.6 mL/min. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCl. 80 Calibration curves for NE were constructed and used to quantitate the amount of NE overflow detected a mesenteric artery adventitia. Calibration curves were generated for an electrode before and after a series of NE-release measurements [184]. Figures 3.8A and B show continuous amperometric i-t responses for a diamond and carbon fiber microelectrode, respectively, for different injected NE concentrations (dose—response curves). The NE was introduced through a non-metallic syringe needle (id. 250 um, MicroFilTM) that was fixed in position adjacent to the microelectrode at a distance of approximately 10 pm. The injected volume was approximately 2 mL and flow rate of Krebs’ buffer in the bath was 1.6 mL/min. Figures 3.8C and D demonstrate that both electrode types exhibit an oxidation current that increases proportionally with the NE concentration injected. The steady-state current for both microelectrodes varied linearly with the NE concentration between 0.01 and 1.0 uM (r2 > 0.99, n = 3, not tested above 1.0 uM). The difference in sensitivity for the two electrodes (slopes differ by 1.5) is attributed to differences in the electrode geometric area. Importantly, the response sensitivity was largely unchanged by exposing the diamond microelectrode to the tissue. The minimum detected concentration was 10 nM (SfN > 3) for the diamond microelectrode. 3.3 Conclusion It has been shown for the first time that the diamond microelectrode yields a sensitive, reproducible and stable response for the catecholamine neurotransmitter, norepinephrine, during in vitro electrochemical measurements. 81 1400 00 pA] -1200? f/ $1000d 20 PA1 :g eooi l g 500: / r T o 400' Area:5.4x10'4cm2 _.,—..—-—--—-'—’—" 200‘ Slope = 1.18 (pA/nM) |.___. 0‘ f2=0.999 1003 0.0 0.2 0.4 0.6 0.8 1.0 Concentratiompfl) 1000 800- , ‘2 ‘ //V :600- , // 5 ‘ . 5400- 0 ‘ Area=6.0x10'4om2 2001 Slope = 0.78 (pA/nM) ‘ r2: . Oi 099’ 0T0 ' 0T2 ' 0T4 ' ofe ' 0:8 ' 1T0 Concentration (all) Figure 3.8. Dose—response curves for NE at (A, C) diamond and (B, D) carbon fiber microelectrodes. Measurements with 0.01—1.0 uM NE were made. Flow rate — 1.6 mL/rnin. Detection potential = +0.8 V (diamond) and +0.4 V (carbon fiber) vs. Ag/AgCl. 82 This new microelectrode is useful for electrochemical measurements in biological environments, providing comparable response sensitivity and precision, and superior response stability as compared to a carbon fiber microelectrode. The superior diamond response stability was achieved without a protective ionomeric coating and results because of the hydrophobic, sp3-bonded carbon surface on which weak adsorption of polar biomolecules and contaminants occurs. Furthermore, it was demonstrated that the diamond microelectrode response for Fe(CN)?”—4 is unaltered during exposure to the laboratory atmosphere. It is postulated that the same surface properties which lead to the superb diamond response precision and stability (i.e., the absence of it bonding and surface oxides) are also the cause for the more sluggish electrode reaction kinetics for catechols and catecholamines. Additionally, it was shown that the diamond microelectrode, when H-terrninated, is devoid of detectable electroactive surface carbon— oxygen functionalities and exhibits a low and stable background current that is independent of the solution pH. 83 Chapter 4 Monitoring Endogenous Norepinephrine Release and Its Effect on the Contractile Response of a Rat Mesenteric Artery Using Continuous Amperometry and Video Microscopy 4.1 Introduction Electroanalytical methods have proven useful over the years for the in vitro and in vivo monitoring of electroactive catecholamine neurotransmitters, like dopamine (DA) and norepinephrine (N E), in the central and peripheral nervous systems. The microelectrode is usually 5-10 pm in diameter and must be prepared appropriately for use in biological environments in order to maximize the response sensitivity and stability [185, 104, 112]. For example, carbon electrode preparation can involve electrochemical pretreatment for activation. However, while useful for improving sensitivity, such pretreatment can introduce surface toughening, surface oxide formation and microstructural damage [34, 141]. The magnitude of these physical and chemical changes and the extent to which they affect the carbon electrode response for the neurotransmitter depend on the pretreatment conditions and the initial electrode microstructure. Pretreatment that causes such physiochemical change has the benefit of increasing the carbon electrode response sensitivity for the catecholamine but the drawbacks of increasing the background current and decreasing the electrode response time. Also, protective polymer coatings, like Nafionm, can be applied to improve the response sensitivity and selectivity for the catecholamine and to minimize fouling in the biological environment by protecting the electrode from biomolecule adsorption [141, 104]. 84 Application of the polymer coating has the drawback of increasing the electrode response time. Electroanalytical methods with carbon fiber microelectrodes have been used extensively for catecholamine monitoring in the brain and the central nervous system since the pioneering work of Gonon and Adams [131, 115, 177, 119-121, 118]. By comparison, there have been fewer reports of in vitro or in vivo electrochemical monitoring of neurogenic processes in the peripheral nervous system, such as the release of NE from the sympathetic nerve terminals of isolated organs [115, 119-121, 118]. As stated in Chapter 1, the sympathetic nervous system regulates blood pressure by controlling the tone of muscular resistance arteries and capacitance veins. At sympathetic neuroeffector junctions with smooth muscle cells, the electrical response is mediated by neurotransmitter release. A triad of neurotransmitters can be released including NE, ATP (adenosine 5’-triphosphate) and neuropeptide Y. Once released, the neurotransmitters diffuse across the synaptic cleft, bind briefly to receptor proteins in the effector cell membrane and, if present in an adequate amount, elicit a specific physiological response (e.g., contraction). However, in many respects, the pathophysiological mechanisms of neural control in arteries and veins are incompletely understood [50, 54, 118]. Long-term, we seek a better understanding of the functional differentiation of noradrenergic transmission in arteries and veins and how these control mechanisms are altered in cardiovascular disease states (e. g., hypertension). Of the neurotransmitters released from sympathetic nerves, NE is the only compound that is easily electrooxidizable, thus, it can be monitoring electrochemically. To date, published reports have demonstrated that electrically or chemically elicited NE 85 release can be electrochemically monitored in vitro in vasculature preparations [123, 121, 124, 118]. For example, continuous amperometry with a carbon fiber microelectrode has been employed to record NE release, as an overflow from sympathetic nerve endings at rat tail and mesenteric arteries [123, 121, 124, 118]. In the majority of these studies, rat tail artery was used because of its dense sympathetic innervation that forms a two- dimensional plexus at the external surface [120, 122, 121, 186, 124]. The carbon fiber microelectrode used in these studies was prepared for use by either anodic polarization [120, 122, 121, 186] or coating with NafionTM [124, 187, 118]. Using the oxidation current or charge recorded for endogenous NE along with the effect of various vasoactive agents, researchers have learned about factors controlling release, receptor binding and clearance at sympathetic neuroeffector junctions. As stated above, we seek to better understand the functional differentiation of noradrenergic transmission in arteries and veins and how these control mechanisms are altered in hypertension. In Chapter 3, the fabrication and electrochemical characterization of the diamond microelectrode were described, along with initial results from its application in electroanalytical measurements in biological tissue [155, 159, 157]. This new carbon electrode provides a greater level of response (i.e., oxidation) sensitivity, reproduciblility and stability than does a bare carbon fiber for monitoring exogenous and endogeous NE in tissue [159, 157]. This new electrode has three characteristics beneficial for in vitro electrochemical measurement: (i) conventional pretreatment is not required to prepare the electrode for use, (ii) the electrode surface is relatively oxygen-free so that the background voltammetric current is low and independent of the solution pH, and there are no features present associated with redox-active carbon-oxygen functional groups, 86 and (iii) a high level of response sensitivity and stability are achieved without the use of a protective polymer film, like NafionTM. In this Chapter, we report on the combined use of in vitro continuous amperometry with a diamond microelectrode and video microscopy to simultaneously measure endogenous NE released from sympathetic nerves innervating a rat mesenteric artery and the resulting contractile response. The NE released was measured as an overflow at blood vessel surface. These two techniques, along with various vasoactive agents, were employed to study the functional relationship between the oxidation current associated with endogenous NE released at neuroeffector junctions near the electrode and the extent of arterial constriction from sympathetic nerve fibers. 4.2 Results and Discussion Simultaneous recording of NE release and vascular constriction evoked by electrical stimulation. Figure 4.1A is a video micrograph showing the placement of the diamond microelectrode (right) and the bipolar stimulator electrode (left) at the surface of a mesenteric artery. The separation between the two was maintained at about 200 pm. The artery is visible in the image and has a rest diameter of 228 um. Figure 4.1B is a video micrograph of the same artery after electrical stimulation, which shows the contracted vessel with a diameter decrease of about 28%. The NE that is released can be oxidatively detected according to the redox reaction shown in Figure 4.1C. The contractile (top) and NE oxidation current (bottom) response transients to a 20 Hz stimulus are presented in Figure 4.1D. Maximal responses were found for this frequency, as discussed later in this Chapter. 87 H OH HO NH? O\ NH» —> + 2HJr + 26' HO 0 NE NE-o-quinone Tl 1‘ TC Era TR ‘1 Constriction . Tso A IIIIIX Current W T50 Tr TD 7' Figure 4.1. Video micrographs showing a mesenteric artery (A) before and (B) after a 20 Hz electrical stimulation (60 pulses with a 0.3 ms pulse width). The bipolar stimulator electrode and the diamond microelectrode, both positioned at the artery surface, are evident in (A). The NE oxidation reaction mechanism is shown in (C). Characteristic contractile (top) and NE oxidation current (bottom) response transients in response to a 20 Hz stimulation are presented in (D). Detection potential = 800 mV. Flow rate = 1.6 mL/min. Several numerical parameters are obtained from these curves including: the maximum oxidation current, 1",“; the current rise time, T,; the time required for current decay to 50% of the maximum, T50; the time required for full current decay, TD; the percent constriction, Co/,; the time required to reach full constriction, Tc; the time required for full vessel relaxation, TR; and the latency period prior to the onset of constriction, TL. 88 The oxidation current recorded with the diamond microelectrode poised at a detection potential of 800 mV rapidly increased after application of the stimulus. In our methodology, the oxidation current is a measure of the concentration of endogenous NE released at multiple varicosities in the vicinity of the microelectrode. The measurement does not provide a direct measure of NE from an individual release site, rather it provides a temporal profile of the nerve stimulation-induced rise in NE levels at the surface of the artery. The magnitude of the measured current (i.e., NE flux) depends on a balance between the amount released from the nearby varicosities and diffusion from the junctional cleft to the microelectrode, and clearance by neuronal reuptake and extracellular enzymatic degradation. The time from the current onset to the point the maximum is reached is described by the rise time, T,. The latency in the onset of the current after application of the stimulus was less than 30 ms. After reaching a maximum, the current decreased slowly over time back to the prestimulation level due to a gradually reduced flux of NE to the electrode. The decrease in current is characterized by a delay time, TD, which in this case was about 30 s. The decreased current reflects a reduction in the NE concentration at the artery surface caused by a combination of diffusion away and junctional clearance mechanisms. The contractile response follows similar temporal dynamics consistent with there being a connection between the oxidation current magnitude and the extent of vasoconstriction. After release, NE diffuses across the junction and binds to al-adrenoreceptors on vascular smooth muscle cells eliciting the contractile response. As it is cleared from the junction, less and less NE is available to activate the 011 receptors, the smooth muscle cells relax and the artery dilates. As with all neuroeffector junctions, the spatio-temporal 89 dynamics of neurotransmitter action at sympathetic varicosities are influenced by diffusion, uptake, enzymatic degradation and receptor desensitization. Noteworthy is the fact that the onset of the oxidation current appears ca. 300 ms earlier than the onset of constriction after stimulation at 20 Hz.This observation is similar to that of Gonon and coworkers who reported that the onset of the oxidation current occurs 10x faster than the onset of arterial constriction [115]. A cause for this may be the longer time required for NE to diffuse into the smooth muscle cell layers and reach a concentration necessary to elicit the contractile response. As mentioned earlier, the constriction did not influence the magnitude of the oxidation current. This was confirmed by simultaneously recording the oxidation current and the evoked contractile response with and without added pyridoxal-phosphate-6- azophenyl-2’,4’-disulfonic acid (PPADS, 10 uM). PPADS is a P2X receptor antagonist that blocks the constrictor effects of ATP released from sympathetic nerves. ATP and NE are both released from sympathetic nerves at mesenteric arteries but ATP is the dominant vasoconstrictor, particularly at low stimulation frequencies. Blocking the P2X receptor largely abolished constriction under the electrical stimulation conditions used in this work. The oxidation current measured for NE release in the presence of PPADS was 12.1 i 1.7 pA and was unchanged at 12.2 i 1.0 pA in the absence of the drug. Potential dependence of the NE oxidation current. Confirmation that the electrically evoked oxidation current reflects endogenous NE is provided by the following data. Figure 4.2A shows an in vitro hydrodynamic voltammetric i-E curve for the oxidation of NE released from sympathetic nerves innervating a mesenteric artery. The oxidation current was measured as a function of the electrode potential using a 20 Hz electrical 90 stimulation. The oxidation current begins to increase at ca. 300 mV and reaches a limiting value at 800 mV. The half-wave potential, Em, is ca. 500 mV and the limiting current, itim, is 12.5 pA. Based on this curve and others like it, a detection potential of 800 mV was selected for the in vitro continuous amperometric recordings described herein. Evidence that the oxidation current results from endogenous NE is the fact that this E1 ,2 is identical to that for a standard solution of NE. Contractile (top) and oxidation current (bottom) response transients for a mesenteric artery electrically stimulated at 20 Hz are shown for a detection potential of (B) 0 and (C) 800 mV. A reproducible contractile response was observed for both stimuli but oxidation current was only detected when the electrode potential was poised at 800 mV; a potential at which NE is oxidized at a mass transfer limited rate. It should be noted that E1 /2 for NE at diamond is some 200-400 mV more positive than the value for a carbon fiber microelectrode [155, 159]. This positive shift reflects more sluggish electrode reaction kinetics; a characteristic of diamond for this class of redox molecules. Importantly, this shift is not due to ohmic resistance (iR loss). It is supposed that the more sluggish kinetics arise because of the absence of strong molecular adsorption on the H-terrninated sp3-bonded surface [182, 155, 159]. McCreery and coworkers have studied the relationship between the adsorption of catechols and the observed electron-transfer kinetics on glassy carbon, and concluded that an adsorbed layer acts as an electrocatalyst for solution phase redox components [182]. This is explained in Chapter 3. A mechanism of “self-catalysis” was proposed. Adsorption apparently lowers the activation barrier for the reaction leading to an increase in the overall redox reaction rate. 91 > .3 - «h A '9 .3 -93- Current (pA) '9 - «P - 9’ 9° B C 0 mV 800 mV l98um W. 25ml] f“ 10pA W 10mm H 53 lng stimulation Figure 4.2. (A) An in vitro hydrodynamic voltammetric i-E curve recorded for NE released from the sympathetic nerves innervating a mesenteric artery. Contractile (top) and NE oxidation current (bottom) response transients for a mesenteric artery in response to a 20 Hz stimulation (60 pulses and a 0.3 ms pulse width) at detection potentials of (B) 0 and (C) 800 mV. The oxidation current was measured with a diamond microelectrode. Flow rate =1.6 mL/min of Krebs' buffer. 92 Effect of TTX on NE release and vascular constriction. Additional evidence that the oxidation current arises from endogenous NE and that the constriction is neurogenically mediated came from studies using various vasoactive substances. The sodium channel antagonist, tetrodotoxin (TTX), was first tested at stimulation frequencies from 1 to 20 Hz. This potent neurotoxin blocks action potentials in nerves by binding to voltage-gated sodium channels in the cell membrane, thereby inhibiting neurotransmitter release and vascular constriction. It is routinely used to verify that changes in vascular tone are neurogenically mediated [115, 186]. Figure 4.3A-C shows contractile response (top) and oxidation current (bottom) transients recorded in response to a 20 Hz stimulation before, after bathing the tissue with TTX, and after washing out the drug. Prior to adding the TTX, reproducible contractile responses and oxidation currents were seen (Figure 4.3A). The nominal maximum current and percent constriction were 17.3 :t 0.2 pA and 29.7 i 0.4 % (n = 3), respectively. Clearly, the response precision for both was good. The latency, TL, in the onset of the oxidation current and the constriction after application of the stimulus was 40 and 500 ms, respectively (about a factor of 10 difference). After the addition of TTX (0.3 uM) for 5 min, no oxidation current or vasoconstriction was seen (Figure 4.3B). There was some constriction evident after the 3rd and 4th stimulation events because the process of washing out the TTX from the tissue had begun. After completely washing out the TTX, though, the elicited oxidation current and contractile response were fully recovered (Figure 4.3C). The results indicate that the contractile response of the mesenteric artery is neurogenically mediated and that NE release is a sodium-dependent process. 93 25 um]: Figure 4.3. (A) Typical contractile and NE oxidation current response transients for a mesenteric artery in response to a 20 Hz stimulation. Contractile and NE oxidation current response transients (B) in the presence of TTX (0.3 11M) and (C) after washing out the drug. The dotted circles indicate the time the electrical stimulation was applied. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 lemin of Krebs’ buffer. 94 Effect of yohimbine on NE release and vasoconstriction. Nerves that communicate using NE have prejunctional a2-adrenergic autoreceptors that regulate release of the neurotransmitter when activated during electrical stimulation [100, 85, 68]. The receptor antagonist, yohimbine, was tested to learn its effect on the oxidation current and the dynamics of vascular tone [115, 119, 186]. It was hypothesized that blocking this autoreceptor would cause an increase in NE overflow at the neuroeffector junction in response to an electrical stimulus. As a consequence, a larger oxidation current was expected. Furthermore, it was hypothesized that the increased junctional NE would increase the magnitude of the contractile response, particularly at low stimulation frequencies. It was previously shown in tissue over-flow measurements that yohimbine (1.0 uM) increases the amount of NE released from the sympathetic nerves innervating a rat mesenteric artery [100, 85, 68]. Other researchers, like Brook and coworkers, have used (la-adrenergic receptor antagonists, like yohimbine and idazoxan, to study noradrenergic transmission in arteries and electrochemically measured increased NE oxidation currents [115, 119, 121, 186, 124]. Figure 4.4 shows contractile response (top) and oxidation current (bottom) transients elicited by 3 and 20 Hz stimulations with (A) being the control data set and (B) being the data set recorded with yohimbine. In the presence or absence of yohimbine, both the oxidation current and contractile response magnitudes were greatest for the 20 Hz stimulation. In fact, the oxidation current in response to the 3 Hz stimulation, particularly in the absence of the drug, was difficult to resolve from the background. However, in both cases, a broad contractile response was seen. As expected, both the oxidation current and the contractile response magnitudes increased in the presence of yohimbine. 95 254 —o—W 40 _.—W . -o—Yotirrflm(1uM) 30] —o—Youttum(1ttm) s . / § 3 l /i t/ :15‘ / 5 m. 3 030/ 0/ ° 10 0+ 5. ./’ 32 / .o/ ‘9 0 5 10 15 20 0 5 10 15 20 Well) Fmrcya-Iz) Figure 4.4. Effect of yohimbine (1.0 M) on the contractile (top) and NE oxidation current (bottom) response transients recorded at a mesenteric artery. The transients were elicited by electrical stimulation at 20 and 3 Hz (60 pulses and a 0.3 ms pulse width) (A) without and (B) with added yohimbine. The dotted circles indicate the time the 3 Hz electrical stimulation was applied. Plots of the oxidation current for NE, with and without (i.e., control) added yohimbine, versus the stimulation frequency are shown in (C). Plots of the contractile response, with and without added yohimbine, versus the stimulation frequency are shown in (D). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs' buffer. 96 This trend is more clearly revealed in the plots presented in Figures 4.4C and D. An increase in the stimulation frequency up to about 10 Hz resulted in an increase in both the NE oxidation current and the percent constriction. More oxidation current was recorded at higher stimulation frequencies because each stimulus during a pulse train causes NE release. Given the fact there is less time for clearance between each pulse, a sustained increase in the oxidation current results because the signals evoked by each successive stimulus sum with the junctional NE remaining from the previous ones. The contractile response curve in the presence of yohimbine is left-shifted from the curve without the drug at frequencies up to about 10 Hz, which reflects the presence of greater junctional levels of NE. Above 10 Hz, the contractile response became constant even though more junctional NE was present. Under these conditions, there is simply more than enough NE present at the surface of the effector cells to activate all of the Oil-adrenergic receptors, hence the unchanging contractile response. Table 4.1 presents a summary of the oxidation current and contractile response data for the 20 Hz stimulation, with and without yohimbine. As noted before, both the oxidation current and contractile responses were quite reproducible. In the presence of yohimbine, the nominal oxidation current increased by 74% from 14.8 to 25.8 pA but there was not much difference in the rise time, T,, or the half-life, T50, of the oxidation current. The decay time, TD, was lengthened as more time was required to clear the high junctional concentration. Within statistical variance, the contractile response, Co/,, was about the same with and without the drug, as were the latency period, TL, and the recovery time, TR. What did increase with yohimbine was the constriction time, Tc, from a nominal value of 5.32 to 6.48 s, and the half-life, T50, from 6.94 to 11.11 s. 97 Table 4.1. Numerical parameters measured from the oxidation current and contractile response transients recorded in the presence and absence of yohimbine (1.0 pM). 20 Hz Control (n = 3) Yohimbine (n = 3) NE release (1mm pA) 14.8 i 0.2 25.8 i 03*— Rise Time (T,, s) 2.78 i 0.05 2.86 i 0.03 Half-life (T50, s) 5.43 i 0.16 4.46 i 0.28 Decay Time (TD, 3) 21.28 ;|: 0.95 26.99 i 1.72* Constriction (Co/,) 33.58 :t 1.27 29.92 i 1.53 Latency (TL, ms) 420 j: 12 490 i 51 Constriction Time (Tc, 3) 5.32 i 0.24 6.48 i 0.31 Half-life (T50, s) 6.94 i 0.43 11.11 i 0.81 * Relaxation Time (TR, 3) 30.0 :I: 3.9 31.07 i 2.98 The oxidation current and contractile response were elicited by electrical stimulation at 20 Hz. Data shown are means i SEM. The values were compared by paired t-tests. *P < 0.05. Even though TR was unchanged, the dynamics of full vessel relaxation were altered by yohimbine. The larger Tc and T50 values were caused by the higher junctional NE concentration and the longer time required for clearance. These results further indicate that the oxidation current measured resulted from endogenous NE. Effect of cocaine on NE clearance and vasoconstriction. In another series of measurements, the NE reuptake inhibitor, cocaine, was added and its effect on the evoked oxidation current and contractile response was investigated [115, 119, 120]. Reuptake of NE back into the nerve terminal by the norepinephrine transporter protein (NET) is inhibited by cocaine. This reduces the clearance rate of junctional NE. It was 98 hypothesized that its presence would cause an increase in the NE oxidation current and a greater contractile response. Figures 4.5A and B show oxidation current and the contractile response transients for 3 and 20 Hz stimulations, with and without added cocaine. For both stimulation frequencies, the oxidation current increased in the presence of cocaine with a larger increase seen for the lower frequency. The contractile response at 20 Hz was unaffected by the presence of cocaine while the response at 3 Hz was significantly increased. It is supposed that this results because the junctional concentration of NE at the higher frequency is sufficient to saturate the nearby smooth muscle cell receptor sites. Interestingly, the increased contractile response at 3 Hz has a biphasic character to it. Sufficient data are not in-hand at present to confirm the origin of this transient response. Figures 4.5C and D show plots of the oxidation current and the contractile responses as a function of the stimulation frequency. At all frequencies, the oxidation current was larger in the presence of cocaine. At frequencies less than 10 Hz, a point at which nearby receptor site saturation occurs, increased constriction was seen in the presence of the drug as evidenced by the leftward shift in the response curve. Table 4.2 presents a summary of the oxidation current and contractile response transient data for the 3 and 20 Hz stimulations, with and without added cocaine. In the presence of cocaine, the nominal oxidation current, imax, increased by 23% from 15.1 to 18.6 pA at 20 Hz and by 222% fi'om 5.06 to 16.3 pA at 3 Hz. Minor changes in T50 were seen for the 20 Hz stimulation; however, TD increased significantly in the presence of cocaine from 16.3 to 27.2 3 due to the lower rate of clearance. 99 Cocaine 25- 8 Current (pA) a a: DUI «monitor -o- Cocaine (10 an) ¢\¢ % Constriction Figure 4.5. Effect of cocaine (10 uM) on the NE oxidation current (left) and contractile (right) response transients recorded at a mesenteric artery. The transients were elicited by electrical stimulation at (A) 20 and (B) 3 Hz (60 pulse and a 0.3 ms pulse width). Plots of the oxidation current for NE, with and without (i.e., control) added cocaine, versus the stimulation frequency are shown in (C). Plots of the contractile response, with and without added cocaine, versus the stimulation frequency are shown in (D). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow //<>/ / /r O ,0/9 ../." 0 '5 1'0 ' 1'5 I 20 Mll'k) rate = 1.6 mL/min of Krebs' buffer. 100 —o-oonuor —o— Coedne(10 uM) o/ O/ 04/ / i 0 1'0 ' 1'5 ' WM) 5 2'0 For the 3 Hz stimulation, significant increases in both T50 and TD were seen; the latter by 400 % from 4.1 to 20.6 s. The contractile response increased by 115% from 11.6 to 24.9% in the presence of the drug. Table 4.2. Ntunerical parameters measured from the oxidation current and contractile response transients recorded in the presence and absence of cocaine (10 uM). 20 Hz Control (n = 3) Cocaine (n = 3) NE release (ImaX, pA) 15.1 :t 0.2 18.6 i 0.2* Half-life (T50, s) 5.08 i 0.10 6.60 i- 0.44 Decay Time (TD, 3) 16.3 i 0.3 27.2 i 2.2* Constriction (Co/,) 33.9 :t 0.5 34.2 i- 0.5 % 3 Hz Control (n = 3) Cocaine (n = 3) NE release (Imam, pA) 5.06 i: 0.2 16.3 i 0.4* Half-life (T50, s) 1.09 i 0.16 5.10 i 0.37* DecayTime (TD, 3) 4.1 -l_- 0.8 20.6 i 1.0* Constriction (C%) 11.6 i 0.4 24.9 i 0.6 * The oxidation current and contractile response were elicited by electrical stimulation at 20 and 3 Hz. Data shown are means i SEM. The values were compared by paired t- tests. *P < 0.05. These trends are all consistent with a reduced rate of junctional clearance. These results further confirm the oxidation current is caused by endogenous NE and that the contractile response is mediated to a significant extent by NE. NE overflow and vascular constriction as a function of the stimulation conditions. Since diamond is a new type of electrode for neurochemical studies, the electrode response for NE was investigated as a function of the stimulation frequency and pulse number. In blood vessels with only a few smooth muscle cell layers, all the cells are all in close proximity to the nerve terminals. This condition produces a strong neurogenic 101 response during electrical stimulation [188, 189]. The frequency dependence of the oxidation current and the contractile responses are presented in Figure 4.6. Increasing the stimulation frequency from 3 to 20 Hz (60 pulse sequence with a 0.3 ms pulse width) caused a proportional increase in both the NE oxidation current and contractile response. Figure 4.6A displays contractile (top) and oxidation current (bottom) response transients in response to several low frequency stimulations. Clearly, the oxidation current increases with increasing frequency, reflecting a greater junctional NE concentration. The increased levels of junctional NE elicit increased vascular constriction. Figure 4.6B presents plots of the NE oxidation current and the percent constriction as a function of the stimulation frequency up to 60 Hz. Both the NE oxidation current and the percent constriction track one another and pass through a maximum at 20 Hz. It was based on these results that measurements with the drugs (vide supra) were performed using a 20 Hz stimulation. However, the oxidation current and constriction both decline at higher stimulation frequencies. There are several dynamic events in nerve signaling processes including (i) storage and release of the neurotransmitter from the neuron, (ii) interaction of the neurotransmitter with the receptor sites on the vascular smooth muscle cells and subsequent production of the post-junctional response, (iii) clearance of the neurotransmitter from the junction by diffusion, reuptake and metabolism and (iv) packaging NE in intemeuronal storage vesicles for release. One possible explanation for the decrease in oxidation current and contractile response with increasing stimulation frequency is rapid depletion of a readily releasable pool of NE-containing synaptic vesicles. 102 A 3 5 7 10 20 Hz {W5 [5 pA 18 4o 16: -o—IEOdddlonQnart _ A , —.—20 Hz). While this is one plausible explanation, it is important to recognize that several other factors can influence the frequency-response characteristics of signaling at a neuroeffector junction, such as the distance between the nerve ending and the smooth muscle cells and the density of innervation (e.g., neuroeffector junctions per length of vessel). Interestingly, not only did the oxidation current and contractile response increase with increasing frequency up to about 20 Hz, but the temporal shape of both responses was altered as well. This was most notable for the low frequency stimulations (1, 3 and 7 Hz) for which multiple maxima are seen in the transient responses. An example of this is presented for the 3 Hz stimulation (dotted circle) in Figure 4.6A. Again, we do not have sufficient data at present to fully explain these trends. Figures 4.7A and B show the effect of the stimulation pulse number, at 20 Hz, on the oxidation current and contractile responses. Figure 4.7A shows contractile response (top) and oxidation current (bottom) transients as a function of the pulse number from 12 to 120 using a 4 min interval between each sequence. Measurable effects (SN 2 3) were observed only after at least 3 pulses at 20 Hz. In other words, the NE released by individual pulses was too low to be detected in our present arrangement. 104 12 30 60 I20 pm 90 120 2 min 25 50 1 —areouanauanut _ A20. -O-W _40= < , ¢/<} _ .9. 5 / ‘6 #515- O/¢ 50% o .// t ‘ Q 0/’\ /9 ’ 5 310- );0/ i 202 . , . a- 5- ~10 010T 50 1 100 ' 1' ' 208 Edsanwflmr2% Figure 4.7. Effect of pulse number (12, 30, 60, 90 and 120, all at 20 Hz) on the contractile (top) and NE oxidation current (bottom) response transients recorded at a mesenteric artery (A). Plots of the NE oxidation current and contractile response versus the stimulation frequency are shown in (B). The data are presented as mean values with the bars reflecting the standard error of the mean. The oxidation current was measured with a diamond microelectrode poised at 800 mV. Flow rate = 1.6 mL/min of Krebs’ buffer. 105 Both the percent constriction and oxidation current increase with pulse number, but only up to a point. Figure 4.78 shows the corresponding plots of these data. The oxidation current increases with pulse number up to about 60. Beyond this, the increase in current with increase in pulse number becomes less. On the other hand, the maximum contractile response is reached after 12 pulses. Unlike the current, the constriction does reach a constant value at higher pulse numbers. It is supposed that this is caused by saturation of all the available smooth muscle cell receptor sites being activated by NE. Quantitative Determination of NE Release. Even though these measurements reflect the concentration of NE at the artery surface and not the concentration released at an individual neuroeffector junction, it is still important to know the concentration corresponding to the measured current. Calibration of a microelectrode during in vitro electroanalytical measurements is a critical issue in neurochemical studies [104, 190, 112]. Calibration curves for NE were constructed and used to determine the concentration detected at the artery surface in response to the electrical stimulation. At the beginning and the end of a series of in vitro measurements (ca. 4 h), the electrode, while remaining in the tissue, was exposed to different injected concentrations of NE. The concentrations were introduced through a nonmetallic (a combination of plastic and fused silica) syringe needle (id. 250 um, MicroFilTM) positioned adjacent to the microelectrode at a distance of about 10 um. Buffer was flowing continuously through the organ bath at 1.6 mL/min during the measurement. This was the same flow rate used in the in vitro tissue measurements described vide supra. A delay time of 1 min was used between injections. The steady-state oxidation current was found to increase linearly with the injected NE 106 concentration between 10 and 100 nM (12> 0.999, n = 3). The electrode response to higher concentrations of NE was not examined. The minimum concenuation detectable at a S/N 2 3 was 10 nM. Importantly, only a slight decrease of 6% in the electrode response sensitivity was observed after a 4 h period of continuous use in tissue. Using this response curve, it was determined that the nominal oxidation current measured in response to a 20 Hz stimulation corresponded to an NE concentration at the artery surface of between 10 and 30 nM. The concentration detected depended on the particular artery. When the nerves were stimulated continuously to achieve a steady-state level of junctional NE, the concentration determined was approximately 70 nM. This value is in good agreement with detected concentrations reported by Gonon et al. for the densely innervated rat tail artery using an anodized carbon fiber [115, 119, 186]. Differences are expected from artery to artery, and even with position along an artery, because the local NE concentration at the electrode surface depends on the distance between microelectrode and varicosities, density of varicosities in the vicinity of the microelectrode, and the concentration of NE released at each varicosity after the stimulation [77, 177]. 4.3 Conclusion In vitro continuous amperometry with a diamond microelectrode and video imaging were used to simultaneously record NE released from sympathetic nerves at the surface of a mesenteric artery and the evoked contractile response. Electrical stimulation of the innervating sympathetic nerves elicited a transient oxidation current, as measured with the microelectrode, that was attributed to endogenous NE. Several observations 107 support the conclusion that NE was the species being oxidatively detected and that the contractile response was neurogenically mediated. First, the hydrodynamic voltammetric E1 /2 for the endogenous NE was identical to that for a standard solution of the neurotransmitter. Second, the evoked oxidation current exhibited the expected characteristics of neurogenically-controlled release, that is, it was abolished by TTX. Third, the oxidation current increased as did the contractile response, particularly at low stimulation frequency, in the presence of the 012-adrenoreceptor antagonist, yohimbine. Fourth, the oxidation current increased, the clearance rate decreased and the contractile response increased at low frequency in the presence of the NE reuptake inhibitor, cocaine. The methodology reported on herein represents an advancement for studies of synaptic physiology in monoaminergic systems. The uncoated diamond microelectrode provided a sensitive, reproducible and stable oxidation current response that varied with the stimulation frequency and pulse number in a predictable manner, similar to that for a carbon fiber microelectrode. 108 Chapter 5 Differences in Sympathetic Neuroeffector Transmission to Rat Mesenteric Arteries and Veins as Probed by In Vitro Continuous Amperometry and Video Imaging 5.1 Introduction Veins in the splanchnic circulation, make important contributions to the regulation of overall hemodynamics including blood pressure [191, 192]. Sympathetic nerve activity plays a prominent role in control of the capacitance function of veins and the resistance function of arteries. Veins differ significantly from arteries in the properties of their contractile and electrical responses to sympathetic nerve stimulation or to drugs that mimic the activity of sympathetic nerves [22, 193, 85]. One reason for these differences is that arteries and veins may be innervated by different sympathetic neurons and these neurons may communicate using different vasoconstrictor neurotransmitters. Adenosine 5’-triphosphate (ATP) and norepinephrine (N E) are the vasoconstrictor transmitters released by sympathetic nerves supplying small (5 200 um) mesenteric arteries while NE is the vasoconstrictor released by peri-venous nerves [193, 194, 50, 53]. In addition, the dynamics of sympathetic neuroeffector transmission to arteries and veins may also differ because the dynamics of vasomotor responses in arteries and veins are different. Arterial diameter controls the delivery of oxygenated blood to tissues and, therefore, arterial sympathetic neuroeffector transmission should permit moment-to-moment control of blood pressure and flow to tissues [195]. Alternatively, veins are a reservoir of blood and shifting blood from the veins to arteries in response to orthostatic adjustments would 109 require rapid and sustained activation of venous sympathetic neuroeffector transmission [196]. Advantage was taken of the excellent time and spatial resolution provided by continuous amperometry to monitor NE oxidation currents subsequent to nerve stimulation in isolated rat mesenteric artery (MA) and vein (MV) preparations in vitro. A carbon fiber was used in this work rather than a diamond microelectrode because it was found that several of the drugs employed to study the pathophysiology were oxidized at the positive detection potential needed to detect NE at the diamond. Detailed studies of the dynamics of NE release in veins have not been reported in the literature because, at least in part, it is a low pressure system that does not determine total peripheral resistance and, thus, has been viewed as largely unimportant in acute or chronic changes in blood pressure. Testing of the hypothesis that there are fundamental differences in sympathetic neural control in mesenteric arteries (MA) and veins (MV) from a normotensive rat is reported on herein. These studies also made use of video imaging, that permitted simultaneous measurement of blood vessel diameter changes in response to nerve stimulation. This allowed for a comparison of the relationship between NE release and vasomotor dynamics in MA and MV. 5.2 Results Distribution of sympathetic nerve fibers associated with MA and MV. Glyoxylic acid (GA) converts NE into an intensely fluorescent 2-carboxymethyl-dihydroisoquinoline derivative [197]. Therefore, the distribution of GA-induced fluorescence of neuronal 110 stores of NE was used to assess the disposition of periarterial and perivenous sympathetic nerves. Figures 5.1A and B show GA-induced fluorescence that revealed a network of varicose nerve fibers around MA and MV. The periarterial nerve plexus consisted of bundles of several axons arranged in a net-like manner. The perivenous plexus was less dense than in arteries and consisted of individual varicose axons largely oriented circumferentially around the vessel. hnages for both are shown in Figures 5.1C and D. The difference in the arrangement of periarterial and perivenous sympathetic nerves was quantitated by counting the number of nerve fibers crossing horizontal and vertical grids superimposed on an image at a 400 X magnification. The number of vertical and horizontal crossing nerve fibers was similar in MA (vertical = 9.7 i 1.2, horizontal = 9.6 i 1.2, n = 4 rats), whereas there were fewer vertical than horizontal nerve fiber crossings in MV (vertical = 3.6 i 0.4, horizontal = 6.8 i 0.9, n = 4; P < 0.05). In addition, the number of vertical crossings in MV was significantly less than the number in MA (P < 0.05). In Vitro NE overflow and contractile responses in MA and MV. The previous study revealed that an electrode potential of 400 mV is sufficient to oxidize NE at a carbon fiber at a mass-transfer limited rate. Furthermore, the oxidation current measured at this potential reflects the concentration of endogenous NE released from sympathetic varicosities near the electrode tip [159]. A detection potential of 400 mV was used for all measurements reported herein. Preparations in which the electrically evoked constriction or oxidation current were not blocked by tetrodotoxin (TTX) were discarded. 111 Figure 5.1. Glyoxylic acid-induced fluorescence of catecholamines in peri-vascular sympathetic nerves. A and B show low magnification (100 X) images of the sympathetic nerve plexus while C and D show high magnification images (400 X). The nerve plexus in arteries has a mesh-like arrangement and nerve fibers are oriented in both the longitudinal and circular axis of the blood vessel (C). The plexus in veins has a selective circular arrangement with very few fibers oriented in the long-axis of the blood vessel (D). 112 The oxidation current and vascular constriction require activation of TTX-sensitive sodium channels and, therefore, these responses are neurogenically mediated. The anatomical data presented above indicated that there is higher nerve density in MA compared to MV. In all work reported on herein, several locations on a blood vessel were probed for NE overflow in order to find a location of maximum current. Sites yielding the largest oxidation current were then chosen for the comparative studies of NE overflow from perivenous and periarterial nerves. It was not possible to reliably detect NE overflow at arteries or veins in response to a single electrical stimuli. Rather, short trains (60 pulses) of stimuli were needed to produce a frequency dependent (0.2 — 30 Hz) and reliably detectable NE oxidation current. The same trains of stimuli also produced frequency-dependent constrictions of MA and MV. NE overflow, as evidenced by the oxidation current, and the evoked contractile response were detected at a lower threshold frequency in MV (0.5 and 0.2 Hz, respectively) than in MA (1 and 0.5 Hz, respectively). Figures 5.2A and B display representative recordings of constriction (top) and NE oxidation current (bottom) response transients from MA and MV at stimulation frequencies of 3 and 20 Hz. It can be seen that at 20 Hz, a constriction of ca. 25% occurs in both vessels. The vessel constricts rapidly in response to the stimulation followed by a slow relaxation to the rest diameter. However, at 3 Hz, the range in arterial constriction was hardly detectable while a broad contractile response was seen for the vein at 3 Hz. For both MA and MV, an intense oxidation current is seen with maximum currents of 12.7 and 23.8 pA, respectively at 20 Hz. The oxidation current elicited by the 3 Hz stimulation is broad but detectable. The current magnitude was larger for MV than for MA, 5.1 and 9.8 pA, respectively. 113 Artery Vein 3 Hz 20 Hz 3 Hz 20 112 W zottmi ] f 20...] l f I—I I—-—l 105 10: NHL—— MIR... W A ns A ns A ns A ns C D 25 30 -m("=24) —o— MV(n=19) I A 20. DW(D=17) -.-M(fl=21) / <5. * .5 201 r: 15. ft; a / 3 10- (f3 10 §/§ \° / 5. ° é _ 0 4 £11! . -- 20 0.1 1 10 Frequeonl-Iz) Figure 5.2. Contractile (top) and NE oxidation current (bottom) response transients for (A) MA and (B) MV elicited by a 3 and 20 Hz stimulation. Plots of the (C) NE overflow current and (D) constriction for MA and MV as a function of the stimulation frequency. *Significantly different from MA and MV (P < 0.05). Data are mean i S.E.M. 114 Figures 5.2C and D contain summary plots for the oxidation current and contractile response magnitudes for MA and MV. These data demonstrate a frequency-dependent (0.2-30 Hz) increase in the NE overflow and constriction for both MA and MV. These data also reveal that NE oxidation currents were greater for MV than for MA and that the frequency-response curve for nerve stimulation-induced constriction in MV was shifted to the left of the curve for MA (Figure 5.2D). The half maximum stimulation frequency (S 50) of the constriction was 1.9 i 0.2 Hz in MV, this value that was significantly lower (P < 0.05) than that for MA (4.8 i 0.3 Hz) (Figure 5.2D). The maximum constriction (Ema) was similar in MV (28 i 2%) and MA (31 i 3%). Quantitative detection of NE overflow. The carbon fiber microelectrodes used for recording NE overflow were calibrated by generating a response curve using a series of standard NE solutions. The NE solution was introduced through a nonmetallic (combination of plastic and fused silica) needle (id. 250 um, MicroFilTM) that was fixed in position adjacent to the microelectrode at a distance of ca. 10 um [159]. Linear response curves were seen from 10 to 300 nM (r2>0.995), as shown in Figures 5.3A and B. In all measurements, the NE overflow current dose not reflect release from an individual varicosity but rather the NE concentration present at the vessel adventitia that is released from multiple varicosities near the electrode. In order to directly compare the sensitivity of MA and MV to the constrictor effects of endogenous NE, the noradrenergic contribution to neurogenically-mediated constriction was investigated in the both blood vessel types. This is discussed further in the next section. 115 Norepinephrine (nM) 70 1 008 oo 0 O) O A .h o A Current (pA) N O 4 j O I ' I V I ' U ' 20 40 60 80 '100'120 Concentration (nM) C Figure 5.3. Calibration of a carbon fiber electrode for quantitative recording of NE overflow. (A) Oxidation currents produced by injection of standard NE solution. (B) Calibration plot showing how the NE oxidation current changes with concentration. Five different electrodes were used to record the data set. Data are mean i S.E.M. 116 Influence of postjunctional receptor antagonists on NE release and vasoconstriction. Along with NE, ATP acting at P2X; receptors contributes to the neurogenic constrictions of arteries in the rat mesentery [31]. In order to study the noradrenergic constriction, the P2X receptor antagonist, PPADS (10 uM) was used to block the constrictor effect of ATP. Figure 5.4A shows that PPADS did not alter NE oxidation current but it did substantially reduce the constriction of MA. In contrast, PPADS did not affect the oxidation current or the constriction of MV as seen in Figure 5.4B. This result indicates that ATP, at least at this stimulation frequency, mediates arterial constriction while NE mediates venous constriction. The all-adrenergic receptor antagonist, prazosin (0.1 uM), was also applied to the tissue individually and in combination with PPADS. For MA, the presence of prazosin partially reduced the contractile response mediated by NE while the presence of both antagonists almost totally abolished constriction in Figure 5.4C. In contrast, adding PPADS had little effect on either the constriction or NE overflow from MV in Figure 5.4D. The addition of prazosin nearly completely abolished the constriction. Comparison of the endogenous NE- concentration-constriction curves in the presence of PPADS for MA and MV indicates that MA are relatively insensitive to endogenously released NE with a maximum constriction of ca. 7% at 20 nM NE (Figure 5.4C). MV are much more sensitive to endogenous NE as 20 nM caused a 3-fold greater constriction (ca. 23%) (Figure 5.4D). Influence of a prejunctional 112-adrenergic receptor agonist and antagonist on NE release and vasoconstriction. Prejunctional a2-adrenergic receptors located at the sympathetic nerve terminals, also called an “autoreceptor”, are activated by NE and function to inhibit further NE release [49-51]. 117 PPADS \ El5pA C -o-cmru(n=11) -o-Oalrd(n=6) 40—‘-P'mi"("=91 40' —A—Pmosln(n=4) "—PPADS("=4I c —-—PPA03(n=3) : w—v—PPAEB+Pramein(n=7) O 3.)- .2 '3 .§ '5 Hi7» ' " 520 '9/ _‘ 320 / r: O 8 /I—‘-‘l O o\°10' + °\10< .‘ .__._ 5.1-Hie 0‘ 0‘ tI‘fll l 4I I I 05101520253035 05101520253035 Qierflmltbmpinefltimkdlll) f‘ ‘ i.“ -' -" In Figure 5.4. Contribution of all—adrenergic and P2 receptors to neurogenic constrictions of MA and MV. (A) Effects of PPADS (10 uM) on neurogenic constriction (top) and NE oxidation current caused by a 20 Hz stimulus train in a MA. PPADS blocked the constriction but not the NE oxidation current. (B) PPADS did not alter the neurogenic constriction or NE oxidation current in a MV. (C) Concentration-constriction curve for neurally-released NE in MA in the absence and presence of PPADS and prazosin (0.1 uM). Peak NE levels were measured during stimulus trains (1, 3, 7, 10 and 20 Hz) by converting oxidation currents to NE levels using the calibration curve shown in Figure SB. (D) Experiments similar to those shown in C except these studies were done in MV during stimulus trains (0.5, 1, 3, 7, 10 and 20 Hz). 118 The influence of the a2-adrenergic receptor antagonist, yohimbine (1.0 uM), and the agonist, UK 14,304 (1.0 uM), on the NE oxidation current and contractile response of MA and MV was studied. Figures 5.5 A-D show the NE oxidation current and the contractile responses detected at the surface of MA and MV in the absence (control) and presence of yohimbine and UK 14,304. Each datum represents the signal seen for a particular stimulation frequency up to 20 Hz. In the presence of yohimbine, the a2-adrenergic autoreceptor was inactivated and NE overflow for MA increased at all frequencies (Figure 5.5A) while an increase in the contractile response was seen only at low frequencies up to 10 Hz (Figure 5.5C). In contrast, for MV in the presence of yohimbine, the NE overflow current increased with frequency up to 7 Hz, but generally had less effect on NE overflow at high frequencies for MV than for MA (Figure 5.5B). The contractile response was similar or lower than the control, particularly at 10 and 20 Hz (Figure 5.5 D). For example, in the presence of yohimbine, the nominal oxidation current for MA increased by 50 % fiom 8.0 to 12.0 after a stimulation, 7 Hz, whereas the current for MV increased by 25 % from 11.8 to 15.8 pA. Interesting is the fact that the contractile responses for MV decreased at high frequencies in magnitude with yohimbine at 10 and 20 Hz (Figure 5.5D). Table 5.1 presents a statistical analysis of the NE oxidation current and contractile response data with and without (control) added yohimbine. In the presence of yohimbine, rate of rise of the oxidation current (slope) increased significantly at both 3 and 20 Hz for MA but there were no significant differences in the half-life decay time (T50) of the oxidation current. 119 A . . B 30 . . 30 ,[:JCorIml(n=8) JDWM'W $1 /Ym(n=8) 257-Ymm01.” :5; 20.-u<14.304(n=0 .. €20: -“"4'°°“"'5’ I: l r/ :7 ‘ f g 15, a 515‘. Z 3 10. g '5 10‘ Z 0 ‘ a /# U . g 54 f 5‘ é 0‘ g 0‘ % 0 C WI”) 35 35 D ‘_A_w(n.6) ai—A—m(nas 30. —v—Yolinflna(n=61 ‘—v—Yotirrflno(n=5) r: fi—A—mmtnfi) : 25%-WWW”) .2 . ° ‘ a 20 '8 2° '5 1 'E ‘ 8 10+ 8 1° .\ 5] .\° 5 l 0. 0‘ k/ l . t v .i v r 1'170 mm Figure 5.5. Contribution of 112-adrenergic receptor-mediated components to NE overflow and neurogenic constrictions of sham mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of yohimbine (1.0 nM) and UK 14,304 (1.0 nM) in MA (A) and MV (B). F requency-response curves for contractile response before (control) and after application of yohimbine and UK 14,304 in MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control MA (MV) and in yohimbine MA (MV). “Indicates significantly different from NE oxidation current in control MA (MV) and in UK 14,304 MA (MV) (P<0.05). Data are mean i S.E.M. 120 Table 5.1. Numerical parameters obtained from the NE overflow oxidation current and contractile response transients recorded in the presence and absence (control) of yohimbine (1.0 uM). Artery (n = 4) Vein (n = 4) Hz Control Yohimbine Control Yohimbine slope 3 0.05 i 0.02 0.09 :I: 0.01* 0.07 i- 0.01 0.15 i 0.02 * (pA/s) 20 0.92i0.03 1.33:t0.17* 1.10i0.09 1.12:0.12 T50 (S) 3 2.80 i 0.55 4.52 i' 1.76 2.45 i 0.10 4.68 i 0.59 * 20 3.35 i 0.74 5.04 i 1.17 3.08 i 0.49 3.64 :l‘. 0.82 TL (S) 3 2.12 i 0.49 2.01 i' 0.54 1.78 i 0.34 2.69 i 0.56 20 0.49 i 0.06 0.53 :t 0.05 1.12 :t 0.12 1.45 i 0.24 RC 3 5.09 i 1.01 4.64 :l: 0.38 4.29 :I: 0.54 9.97 i 1.97* (W’s) 20 3.68 a 0.48 3.93 i 0.70 2.60 a 0.19 2.85 1 0.30 TR (3) 3 7.52 i 1.67 8.86 i 3.29 28.5 i 8.5 38.7 i 7.7 20 10.7 i- l.0 12.1 i: 3.5 23.2 i 5.5 19.6 i 2.9 The oxidation current and contractile responses were elicited by electrical stimulation at 3 and 20 Hz. *Indicates significantly different from NE oxidation current and contractile responses in control MA (MV) (P<0.05). Data are mean i S.E.M. Slope, rate of rise of oxidation current; T50, half-life current decay time; TL, latency to onset of constriction; RC, rate of constriction; TR, the time required for full vessel relaxation. 121 Interestingly, the slope of the current and T50 and the rate of constriction (RC) were increased significantly only at 3 Hz for MV. Other responses, such as the latency period to onset of constriction (TL) and the time required for full vessel relaxation (TR) were changed little by the drug. While blocking of a2-adrenergic autoreceptors with antagonist would be expected to increase the junctional NE levels, activating the 112-adrenergic autoreceptors with an agonist would be expected to decrease the amount of transmitter released. This was the case as the data reveal that both NE overflow and contractile responses in MA were mostly blocked in the presence of UK 14,304 (1.0 uM) (Figure 5.5A and C). However, UK 14,304 reduced NE overflow and the contractile response less for MV than for MA (Figures 5.5B and D). As an example, in the presence of the agonist, the nominal oxidation current and contractile responses for MA decreased by 68% from 11.9 to 3.8 pA and by 71% from 27.8 to 8.1% at 20 Hz, respectively, whereas the responses in MV decreased by 39 % from 15.0 to 9.1 pA and by 55% from 28.3 to 12.8% at 20 Hz, respectively. Table 5.2 shows TL and RC were significantly increased in both MA and MV but other values were decreased due to decreased NE oxidation current in the presence of UK 14,304. Interestingly, UK 14,304 preconstricted MV, 15.1% i 5.7 (n = 6) during addition of the drug. Influence of cocaine and yohimbine + cocaine on NE overflow and vasoconstriction. NE reuptake by the norepinephrine transporter protein (NET) can be blocked by cocaine. Such inhibition will reduce the clearance rate of junctional NE and increase in NE overflow and contractile responses. 122 Table 5.2. Numerical parameters obtained from the NE overflow oxidation current and contractile response recorded in the absence (control) and presence of UK 14,304 (1.0 uM) treatment. Artery (n = 3) Vein (n = 3) Hz Control UK 14,304 Control UK 14,304 Slope 3 0.04 a 0.02 -* 0.12 a 0.02 0.02 a 001* (pm) 20 0.97 i 0.22 0.11 a 005* 1.41 i 0.26 0.36 i 009* T50 (s) 3 2.9 a 0.6 -* 3.0 a 0.6 1.0 a 05* 20 4.0 a 1.5 0.4 at. 02* 1.8 i 0.2 1.0 i 01* 1.C (s) 3 3.8 a 0.9 -* 1.9 :l: 0.4 5.6 i 04* 20 0.41 i006 1.34:013* 0.93 :01] 1.83 5015* Re 3 4.9 a 0.7 -* 3.3 :t 0.5 13.3 a 2.5* (“m/S) 20 3.8 i 0.5 3.4 .‘L' 0.2 2.1 a 0.2 2.8 :l: 02* TR (3) 3 7.2 a 1.4 -* 20.7 i 2.6 4.4 i 07* 20 10.6 a 1.5 4.4 a 02* 45.6 a 14.9 15.2 i 5.4* The oxidation current and contractile responses were elicited by electrical stimulation at 3 and 20 Hz. *Indicates significantly different from NE oxidation current and contractile responses in control MA (MV) and in the drug treated MA (MV) (P<0.05). Data are mean i S.E.M. Slope, rate of rise of oxidation current; T50, half-life current decay time; TL, latency to onset of constriction; RC, rate of constriction; TR, the time required for full vessel relaxation. 123 Figure 5.6 shows NE oxidation currents and the contractile responses for MA and MV as a function of the stimulation frequency up to 20 Hz, before (control) and after the addition of cocaine (10 uM) and yohimbine (1.0 uM) + cocaine (10 uM). In the presence of cocaine, the NE oxidation current increased for MA (Figure 5.6A) at all frequencies while the contractile response increased at low frequencies below about 7 Hz (Figure 5.6C). In contrast, the NE oxidation current and contractile responses for MV were relatively unchanged except at 3 Hz (Figures 5.68 and D). As an example, in the presence of cocaine, the nominal NE oxidation current for MA increased by 106 % from 6.4 to 13.2 pA at 3 Hz, whereas for MV the increase was 46 % from 7.9 to 11.5 pA. Table 5.3 presents a statistical analysis of the NE oxidation current and contractile response data before (control) and after addition of cocaine. In the presence of cocaine, slope, T50, and other values significantly increased for MA while NE oxidation current and contractile responses in MV are little changed. In order to study an interaction between reuptake and the prejunctional 02- adrenergic autoreceptor, cocaine (10 11M) and yohimbine (1.0 uM) were used in MA and MV. Figures 5.6A and B show that combined drug treatment increased markedly the oxidation current for MA compared with cocaine alone, whereas both drugs increased the oxidation current only at low frequencies for MV. The contractile response was increased at low frequencies (< 7 Hz) but not at high frequencies for MA (Figure 5.6C), whereas decreased constriction was observed at all frequencies for MV (Figure 5.6D). Table 5.4 shows that the combined drugs significantly increased all values related to oxidation current in MA at both 3 and 20 Hz but only increased at 3 Hz in MV. 124 A B 60- 1: Comol (n =10) 604:1 Como! (n = 10) l-C(XHI‘B(I'l-'=8) 1-W(n:9) Asol-CocfineWoflnflnamsa A 50‘-Cocalne+Yohlnflm(n=5) :40 g 40. E ‘ *5 ‘ ~13 3°. g 39 fl 8 20* 0 20. 10 1o- 0. WWI) C D 40 -A—Corlrol(n=6) 40' _A_ C ("‘9’ “"‘ WNW“) ‘-V—Oocdm(n=9) 3),—F Oocdm+Ydirtim(n=6) md—A—W+Ym(ngs C T T g 0 .2 ‘6 :6 3520- § 20. 2 g 8 /. / o 0 / o . f 2 , j 1 10 1 10 WW FIRING-k) Figure 5.6. Contribution of the NE reuptake to NE clearance and neurogenic constriction of mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of cocaine (10 uM) and cocaine + yohimbine in MA (A) and MV (B). Frequency-response curves for contractile response before (control) and after application of cocaine (10 uM) and cocaine + yohimbine in MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control MA (MV) and in cocaine treated MA (MV) (P<0.05). ”Indicates significantly different from NE oxidation current in control MA (MV) and in cocaine + yohimbine treated MA (MV) (P<0.05). Data are mean i S.E.M. 125 Table 5.3. Numerical parameters obtained from the NE overflow oxidation current and contractile response transients recorded in the presence and absence (control) of cocaine (10 11M). Artery (n = 4) Vein (n = 4) Hz Control Cocaine Control Cocaine Slope 3 0.06 i 0.02 0.18 i 0.05”“ 0.20 i 0.11 0.24 i 0.09 (pA/s) 20 0.67 i 0.08 1.28 i 0.28“ 1.10 i 0.18 1.14 i 0.32 T50 (s) 3 3.2 i 0.2 6.9 i 1.2* 6.2 i 1.5 6.7 i 1.5 20 2.6 i 0.3 5.74 :t 2.0 3.2 i 1.0 2.5 i 0.5 LC (5) 3 3.1 i 0.7 1.7 i 0.4* 2.1 i 0.4 2.6 :1: 0.4 20 0.51 :t 0.03 0.58 i 0.06 1.01 i 0.08 1.01 i 0.06 RC 3 4.4 i 0.8 4.4 i 0.5 12.9 i 5.3 18.2 i 3.9 (“m/S) 20 4.0 i 0.2 3.6 i 0.2 2.8 i: 0.2 2.9 i 0.2 TR (5) 3 8.4 i 0.6 7.9 i- 2.5 106 i 40 116 i 38 20 20.3 i 1.6 20.7 i 3.2 88.5 i 29.8 79.9 i 36.4 The oxidation current and contractile response were elicited by electrical stimulation at 3 and 20 Hz. *Indicates significantly different from NE oxidation current in control MA (MV) (P<0.05). Data are mean i SEM. Slope, rate of rise of oxidation current; T50, half- life current decay time; TL, latency to onset of constriction; RC, rate of constriction; TR, the time required for full vessel relaxation. 126 Table 5.4. Numerical parameters obtained from the NE overflow oxidation current and contractile response transients recorded in the presence and absence (control) of cocaine + yohimbine (Co + Yo). Artery (n = 4) Vein (n = 4) Hz Control Co + Yo Control Co + Yo Slope 3 0.04 i 0.01 0.19 i 005* 0.09 :0 .01 0.14 i 0.01 * (pm) 20 0.88 i 0.08 1.89 i 032* 0.95 i 0.07 0.82 3: 0.16 T50 (s) 3 2.6 i 0.20 9.8 i 09* 2.8 i 0.2 6.3 3. 0.8 * 20 2.9 i 0.5 8.2 i 1.5* 3.1 i 0.5 4.7 i 0.9 LC (3) 3 3.2 i 0.4 2.3 i 05* 2.1 i 0.3 1.8 :1: 0.2 20 0.52 i 0.1 0.6 4: 0.06 0.99 i: 0.12 0.93 i 0.1 Re 3 4.6 i 0.6 4.5 :1: 0.6 4.5 i 0.6 11.7 i 2.9 * (“m/S) 20 3.2 i 0.3 3.5 i 0.3 2.3 i 0.2 2.8 i 0.2 TR (5) 3 6.5 i 0.8 9.1 i 2.8 38.9 i 8.1 39.5 i 7.9 20 10.2 i 0.7 9.6 i 2.5 42.6 3: 8.4 47.6 i 9.6 The oxidation current and contractile response were elicited by electrical stimulation at 3 and 20 Hz. *Indicates significantly different from NE oxidation current in control MA (MV) (P<0.05). Slope, rate of rise of oxidation current; T50, half-life current decay time; TL, latency to onset of constriction; RC, rate of constriction; TR, the time required for full vessel relaxation. 127 RC was only significantly increased at 3 Hz in MV in the presence of both drugs but other contractile values were little changed in both MA and MV. 5.3 Discussion The fluorescence images showed that MA and MV have networks of sympathetic nerve fibers running over their adventitia] surfaces but there are fewer nerve fibers in MV [26, 189]. This does not necessarily indicate that the density of innervation of MV is less than MA. MV are thinner than MA so the number of axons per smooth muscle cell could be similar in MA and MV. This suggestion is supported by the fact that ultrastructural studies have shown that in guinea pigs the density of neuromuscular junctions was similar for MA and MV [198]. The nerve plexus for MA consists of bundles of axons arranged in a mesh-like network with nerve fibers equally likely to run parallel or perpendicular to the longitudinal axis of the vessel. For MV, the plexus consisted of single axons with more of a circumferential arrangement. This arrangement made optimal detection of NE overflow quite dependent on the position of the recording microelectrode on the MV surface, more so than for MA. Small microelectrode movements (5 50 um) resulted a marked change in NE oxidation current for MV. Therefore, NE overflow measurements were made with MV only after the electrode had been optimally positioned to record the maximum current (i.e., positioned in the vicinity of multiple varicosities). Perivenous sympathetic nerves are more sensitive to nerve stimulation than as periarterial nerves. Frequency-response curves for the NE oxidation current measured at the surface of MV were left-shified from those for MA. Similar trends have been 128 observed for NE spillover from whole MA and MV segments by high-performance liquid chromatography (HPLC) [100, 85]. Although the trends in the data from these previously reported studies are similar to those reported herein, there is major difference in pathophysiological interpretation. In the present work, short stimulus trains were used to evoke measurable concentrations of NE at the recording microelectrode very close to the release sites. Spillover measurements, on the other hand, require longer stimulus trains to evoke detectable NE release and reflect NE concentration from large segments of vasculature. Short stimulus trains (5 3 s) mimic in vivo sympathetic nerve activity, which occurs in short bursts (< 2 s) [101]. MV are more sensitive to the constrictor effects of sympathetic nerve stimulation than MA [22, 193, 31]. As MV are also more sensitive to the constrictor effects of exogenous NE, post-junctional factors clearly contribute to this functional difference between MA and MV. Additionally, there may be functional differences in periarterial and perivenous sympathetic nerves. F requency-response curves for the NE overflow oxidation currents for MV are shifted to the lefi of those for MA. These data are similar to those obtained in previous work which revealed that NE overflow in canine MV exceeded that in MA [54]. Increased NE overflow in MV relative to MA could be due to several factors. Firstly, the mechanisms of NE release from perivenous and periarterial sympathetic nerves could differ such that individual action potentials release more NE in MV. There are no data available to address this issue; however, different populations of sympathetic nerves are known to supply arteries and veins [199]. There may be differential expression of calcium channels, autoreceptors or vesicular proteins in arterial and venous sympathetic nerves, which contributes to the functional differentiation. 129 Continuous amperometric recording and traditional spillover measurements are sensitive to the fraction of released NE that diffuses away fiom the neuroeffector junction. With amperometric techniques, the distance from the release site to the recording microelectrode is minimized as are the number of variables that could alter NE measurement. However, differences in NE oxidation currents in MA and MV might be due to differences in the efficiency with which neuroeffector junctions in MA and MV clear NE. If little NE escapes the neuroeffector junction because of efficient recovery by the norepinephrine transporter (NET) in MA for example, then less NE will be measured outside the MA junction. In this study, it was hypothesized that prejunctional modulation in rat MV and MA regulates NE release differently. Multiple autoreceptors and heteroreceptors localized at the nerve terminal modulate sympathetic nerve activity. These modulatory receptors, such as 112-adrenergic autoreceptors, can either enhance or attenuate transmitter release during sympathetic nerve stimulation. Secondly, it was hypothesized that the functional properties of NET in rat MA and MV also differentially regulate the rate of NE clearance from the neuroeffector junction. Prejunctional 0.2-adrenergic receptors regulate NE release differently in MA and MV. Prej unctional 112-adrenergic autoreceptors located at the sympathetic nerve terminals are activated by exocytotic release of NE and inhibit further release [49]. Therefore, the autoreceptors mediate NE release through a feedback mechanism at the axon terminal. The results demonstrate that blocking the neuronal aZ-adrenergic receptors with yohimbine increased NE overflow oxidation current for MA at low and high frequencies. However, the oxidation current for MV was increased only at low 130 frequencies. Conversely, UK 14,304 significantly decreased the NE oxidation current for both MA and MV, but more significantly for MA. These results suggest that prejunctional (112-adrenergic autoreceptors play more prominent role in regulating NE release in MA than in MV. There may be quantitative (density) differences in 012- adrenoceptor-mediated neuromodulation in the MA and MV. The aZ-adrenoceptor sites may also be spatially separated from the sites of NE release in MV, which may participate to differing NE release in the presence of aZ-adrenoceptors antagonist or agonist in MA and MV. Differential activity of norepinephrine transporter (NET) in MV and MA. Blocking the neuronal 02-adrenergic autoreceptors caused an increase in the amount of NE released, which leads to a higher junctional concentration. Inhibition of reuptake also increased the junctional NE concentration because of reduced clearance. Therefore, the junctional concentration of NE and its postjunctional effect is influenced not only by the rate of NE release but also by reuptake, metabolism and diffusion [48]. Termination of transmitter actions at the neuroeffector junction is dependent on both removal and metabolic processes. The mechanism of NE clearance may vary between MA and MV. After its released into the junctional region, approximately 95% of the released NE is transported back into the nerve terminal via the NET. The neuronal NET (uptake 1) accounts for the clearance of almost 90% of the total NE released. There is also an extraneuronal transporter [47]. The extraneuronal transporter (uptake 2) is much more diversely localized in the vascular endothelium and smooth muscle cells and has lower affinity for NE than does neuronal NET. It is responsible for clearing only about 5% of the total NE released [47]. The remainder of the NE is metabolized extraneuronally by 131 enzymes or simply diffuses away into the extracellular medium. It is then latter NE that is oxidatively detected by the microelectrode. Inhibition of the NET was accomplished with cocaine. In the present of cocaine, the endogenous NE concentration was significantly increased for MA but not for MV. Interestingly, cocaine increased the half-life decay time (T50) at 3 Hz for MA but not for MV, whereas yohimbine increased T50 at 3 Hz for MV but not MA. This result suggests that NET play a more important role in the clearance of released NE at MA than at MV. However, an uptake 2 blocker, corticosterone (10 uM), did not significantly change the amplitude or time course of the electrochemical signals evoked by electrical stimulation, indicating that uptake 2 does not contribute significantly to the clearance of NE at either rat MA and MV (not shown). The corticosterone result is similar to that reported for the rat mesenteric and tail arteries [200, 1211 The finding that the endogenous NE concentration was not altered by blockade of neuronal uptake in MV can be explained if clearance of NE from its site of release is primarily by diffusion and not by reuptake. One possible reason is for the reduced reuptake is that the NET sites are spatially separated from the sites of NE release in the rat MV, whereas the NET is more closely located to the sites of NE release in MA. Another possibility is that strong feedback mechanism between uptake and the autoreceptors may function more effectively in MV than MA. When higher extracellular NE concentrations occur after cocaine treatment, this effect would be suppressed by the activation of inhibitory autoreceptors in MV. Thus, the increased extracellular NE levels resulting from cocaine uptake blockade might activate orZ-autoreceptors and shut down subsequent NE release in MV. Alternatively, increased NE release by yohimbine 132 activates uptake transporter more in MV. Therefore, in the presence of either cocaine or yohimbine in MV, NE overflow was less increased compared with MA. Autocompensation system between prejunctional 0.2-receptors and NE uptake transporters in MA but not in MV. The higher concentrations of NE in the region of the nerve-ending limit release of the neurotransmitter by feedback inhibition via prejunctional 012-adrenergic autoreceptors, thereby potentiation by uptake inhibition could be masked to sympathetic stimulation. The results showed that NE overflow was significantly increased in the presence of cocaine + yohimbine for MA but not for MV, as compared with either cocaine or yohimbine alone. Also, the more significant decline in the NE oxidation current for MA after repeated stimulation in the present of cocaine or cocaine + yohimbine may result from exhaustion of prejunctional NE stores. The rate of synthesis of NE may not compensate for the NE lost during a firing event [136]. The present finding indicates that NE reuptake and autoinhibition play an important role in maintaining available stores of NE in MA. Different neurogenic contractile responses in MA and MV. Differences in the transmitters released from periarterial and perivenous sympathetic nerves contribute to functional differences in MA and MV. The al-adrenergic receptor antagonist, prazosin, had little effect on neurogenic constriction of MA but it blocked venous responses. Alternatively, the P2X receptor antagonist, PPADS, blocked constriction in MA but not MV. Therefore, neurogenic constrictions in MV are mediated moreso by NE acting at al-adrenergic receptors, while MA tone is mediated by ATP acting at P2X receptors 133 [100, 201, 36, 31]. These results confirm that in small MA, ATP is the dominant neurotransmitter, whereas in MV, NE is the transmitter mediating constriction. It was also showed that MA is relatively insensitive to the constrictor effects of endogenous NE and this highlights the importance of ATP as a transmitter in MA. The NE concentrations measured were undoubtedly lower than the junctional concentrations [121]. Recovery of NE via NET will limit the amount of NE diffusing to the microelectrode [47]. It can be seen from these data that change in vascular tone did not influence the NE oxidation current recorded with the microelectrodes. Postjunctional orZ-adrenergic receptors in MV. Although the NE overflow oxidation current increased in the presence of yohimbine or cocaine, the contractile response reached a maximal value and was the same magnitude with and without the drugs at high frequencies (> 7 Hz). In some cases, constriction decreased with the combined drugs. This may be because the postjunctional receptor sites on the vascular smooth muscle cells were saturated at the concentrations of NE released from nearby varicosities and maintained in the junction. Another possibility is that the postjunctional Oil-adrenergic and P2X purinergic receptors for MA were downregulated or desensitized by increased NE and ATP concentrations in junctional clefi in the present of the drug. Interestingly, MV was preconstricted in the presence of the orZ-agonist, UK 14,304 (1.0 uM), ca.15 % of their original diameter. Also, MV contractile response in the presence of the orZ-antagonist, yohimbine (1.0 uM), remained constant or decreased despite the increased NE overflow at low frequencies of 1 and 3 Hz, and even decreased at high frequencies 10 and 20 Hz. The results suggest that neurogenic constrictions in MV may also be mediated partly by 134 orZ-adrenergic receptors not only by or] -adrenergic receptors. However, other possibilities are that high concentrations of aZ-adrenergic receptor antagonist and agonist influenced on (rd-adrenergic receptors as partial al-adrenergic receptor antagonist and agonist in MV. 5.4 Conclusion In this Chapter, the contribution of NET and the prejunctional aZ-adrenergic autoreceptor to the regulation NE overflow from sympathetic nerves in MA and MV was evaluated. NE overflow was evaluated using continuous amperometry with carbon fiber microelectrode by measuring the oxidation current. The results confirmed that the mechanism of NE release and the corresponding vascular constriction in MV are significantly different from that in MA. The different sympathetic neuroeffector mechanisms in MV could be due to different functions of prejunctional NET and autoreceptors, and different types of postjunctional receptors. NET and aZ-adrenergic autoreceptors play a more prominent role in controlling NE clearance and release at the neuroeffector junction in MA than in MV. It would contribute that NE overflow in rat MV exceeded that in MA. Therefore, the greater NE overflow in veins compared to arteries is not likely due to the difference in innervation density of arteries and veins. Rather, it is due to the difference in functional properties of sympathetic nerves innervating veins and arteries. This provides further evidence supporting the notion that sympathetic nerves supplying veins and arteries are different [22, 193]. These differences in neuroeffector transmission may contribute to the different hemodynamic functions of arteries and veins. 135 Chapter 6 Alterations in Sympathetic Neuroeffector Transmission to Mesenteric Arteries and Veins in DOCA-salt Hypertensive Rats 6.1 Introduction The sympathetic nervous system (SNS) influences importantly on the cardiovascular hemodynamics through its control over the liberation of NE into the circulation. Therefore, alternations in the peripheral sympathetic tone and reactivity might facilitate the development of various pathologies of the cardiovascular system including hypertension [69]. Since increased SNS activity increases arterial and venous constriction and blood pressure, hyperactivity of the SNS is a critical factor for hypertension. There is growing evidence that human essential hypertension and many animal models for the disease are associated with overactivity of the SNS. However, the precise causal mechanisms leading to sympathetic augmentation in hypertension are still poorly understood because maintenance of normal blood pressure involves a complex interplay among the nervous, endocrine, renal and cardiovascular systems. The deoxycorticosterone acetate (DOCA)-salt hypertensive rat model is used widely to study salt-sensitive hypertension, which involves neuronal and hormonal mechanisms [64, 31]. The DOCA-salt rat model has been used to study the involvement of both the central and peripheral components of the SNS in the development and/or maintenance of hypertension [202, 203, 28, 80]. This animal model allows one to evaluate the functional consequences of altered sympathetic regulation of the cardiovascular system and providing insight into the defects that underlie the sympathetic 136 impairment [67]. Mesenteric arteries (MA) and veins (MV) are extensively used in physiological and pharmacological experiments to understand the role of the SNS in modulation of vascular smooth muscle tone [12]. Many studies have reported that increased activity of the SNS increases plasma catecholamine level in MA and/or MV [202, 28, 204-206]. The elevated NE concentrations might reflect elevated sympathetic tone. Therefore, measurement of NE levels is an indirect index of sympathoadrenal activity under certain standardized conditions [207, 208, 67]. The increased sympathetic nerve activity and NE release observed in DOCA-salt hypertension suggest that there may be alterations in the local mechanisms that modulate sympathetic neurotransmission in arteries and veins. Altered NE reuptake, impaired prejunctional aZ-adrenergic autoreceptors and/or increased sympathetic nerve firing rates have been proposed as causes for elevate NE release or could be responsible for the increased plasma catecholamine levels in hypertension animal models [86, 87, 67, 79, 70, 31, 68]. Increased NE release and or reactivity of postjunctional receptors may also result in increased or decreased contractile responses [209, 31]. However, the results are still controversial. Moreover, there have been few detailed studies of the mechanisms underlying increases in sympathetic input to veins in hypertension [210, 176, 31, 68]. As emphasized in this dissertation changes in neuroeffector transmission to veins in hypertension is an important factor for blood pressure regulation [176]. Most previous studies have indirectly investigated endogenous NE release from perivascular sympathetic nerve endings using tissues labeled with tritiated NE and high- performance liquid chromatography [98-100, 85, 68]. Therefore, little is known about 137 altered characteristics of NE release from perivascular sympathetic nerves in hypertension on an impulse-by-impulse basis [121]. It was hypothesized that there are dysfunctional 112-adrenergic autoreceptors and or NET in DOCA-salt hypertensive rat MA and MV is a cause for the increased junctional NE concentration and altered contractile responses compared to normotensive sham rat MA and MV. Therefore, altered neuroeffector transmission is a cause for the increased blood pressure in DOCA-salt hypertensive rats. To test the hypothesis, neurogenic NE overflow and vasoconstriction were measured directly in the absence and presence of reuptake blocker, prejunctional (112-adrenergic receptor antagonist and agonist, postjunctional 011 -adrenergic receptor antagonist and P2X purinergic receptor antagonist. 6.2 Results In Vitro measurement of neurogenic NE overflow and vasoconstriction in normotensive and DOCA-salt MA and MV. The administration of DOCA and saline water to rats for a period of 4 weeks results in increased blood pressure and decreased body weight. Mean systolic blood pressure for sham and DOCA-salt rats was 125 i 2 mmHg (11 = 40) and 194 i 3 mmHg (n = 32), respectively (P < 0.05). The mean weight of sham rats was 416 j: 5 g (n = 34) which the mean weight of DOCA-salt rats was 336 t 7g (n = 35) (P < 0.05). NE overflow and vasoconstriction caused by electrical stimulation of the nerve endings was investigated using MA and MV from sham and DOCA-salt hypertensive rats. Short trains (60 pulses) of stimuli produced frequency dependent (0.2— 20 Hz) and reliably detectable NE oxidation current from periarterial and perivenous nerves. NE overflow and contractile response were detected at lower threshold frequency 138 for MV (0.5 and 0.2 Hz, respectively) than for MA (1 and 0.5 Hz, respectively). Figures 6.1A and B display representative recordings of the contractile (top) and NE oxidation current (bottom) response transients for a 3 Hz stimulation of sham and DOCA-salt MA and MV. As mentioned in Chapter 5, since there are more nerve fibers associated with MA compared to MV, measurement of NE overflow at MV was critically dependent on the electrode position. Several positions along a blood vessel were tested in order to find a maximum current. It was at this location that all of the subsequent testing was done. Figures 6.1C and D are plots of the oxidation current and percent (%) constriction as a function of stimulation frequency for both vessel types. These data demonstrate a frequency-dependent (0.2 -20 Hz) increase in the NE oxidation current (NE overflow) and constriction for both sham and DOCA-salt MA and MV. The data reveal that NE overflow from sham MV was greater than from sham MA but not significantly different compared with DOCA-salt MV (Figure 6.1C). However, NE overflow in DOCA-salt MA was significantly higher compared to that in sham MA. The frequency-response curves for the contractile response of sham and DOCA-salt MV were both shified to the left of the curves for sham and DOCA—salt MA (Figure 6.1D). The mean of the half maximum stimulation frequency (S50) were similar for sham and DOCA-salt MV but the maximum constriction (Em) obtained from DOCA-salt MV was significantly lower than that for sham MV (Table 6.1). No significant differences were seen for sham and DOCA-salt MA in terms of the Em or S50 values. 139 A B Artery Veln 273 1““ ww 309 pm 273 pm I201mm 1% I5 p A I5 pA SS SS Sham DOCA-calf Sham DOCA-salt U 5 ‘ -D-SlnnMA(n=21) A c 25. -A-SlunMV(n=16) A if? g- ,9 ‘ .. . :I 33 20 c 1.1 1 E g 15 3 . U U 10 °\" 5- 0 O. Figure 6.1. Contractile (top) and NE oxidation current (bottom) response transients for (A) MA and (B) MV from sham and DOCA-salt rats in response to a 3 Hz stimulation. Comparison of the frequency response of (C) the oxidation current and (D) vasoconstriction in sham and DOCA- salt MA and MV. *Significant difference between sham and DOCA- salt MA (P < 0. 05). &Significant difference between sham MA and sham MV (P < 0. 05). Data are presented as mean + S. E. M. 140 Table 6.1. Maximum constriction (Em) and half maximum stimulation frequency (S50) for MA and MV from sham and DOCA-salt rats. All data are expressed as the mean i S.E.M and “11” value refers to the number of animals from which the data were obtained. Artery Vein Sham (n=15) DOCA-salt (n=l6) Sham (n=13) DOCA-salt (n=14) S50 (Hz) 5.8 a: 0.3 5.6 :1: 0.7 1.6 i 0.3 " 1.3 3. 0.4 ” 15......r (%) 31.3 i 3.4 29.8 :1: 2.9 28.0 i 2.2 21.6 i 1.9* ”Indicates significantly different from the S50 in mesenteric arteries (MA) (P<0.05). *lndicates significantly different from the Em in sham mesenteric veins (MV) (P<0.05). Effect of prejunctional 0.2-adrenergic receptors on NE release and constrictions in DOCA MV and MA. To determine whether the 112-adrenergic autoreceptor that regulates NE release is altered in DOCA-salt MA and MV, the effects of the 02- adrenergic receptor antagonist, yohimbine (1.0 uM) and the agonist, UK 14,304 (1.0 uM), on NB overflow and the elicited contractile response were investigated. Figure 6.2A Shows that yohimbine significantly increased NE overflow at all frequencies for sham MA. However, yohimbine was less effective increasing NE overflow in DOCA-salt MA. In the presence of yohimbine, the nominal oxidation current increased by 49.4 % from 14.0 to 21.0 pA at 20 Hz (11 = 5) and by 120.6 % from 5.8 to 12.7 pA at 3 Hz (11 = 5) for sham MA, whereas the nominal oxidation current increased by 10.4 % from 16.8 to 18.5 pA at 20 Hz (11 = 6) and by 54.6 % from 7.3 to 11.3 pA at 3 Hz (11 = 6) for DOCA-salt MA. The apparent increase in NE overflow caused elevated constrictions in sham and DOCA-salt MA at low frequencies (1 and 3 Hz) and the contractile response curves in the presence of yohimbine were lefi-shified from the curve without (control) the drug as seen in Figure 6.2C and Table 6.2. 141 141 if} 35 35 11:181unMMn'5) I:ISIunMV(n-4) 30.-oocAMA(n=6) -DOCAMV(n-5) A 1 Emmmmrfi) Eamwnmmmc) é 25‘ -mCAMNYolintim(n*-6) at 2 -mmNMm(n-5) :20, _# E L920 - , 515‘ g 2: E15 xg 4%” = 1 E a * 7g 5 10* g: g: 10* Z: Z: 1 = a 1 a 5 54 g; g; 54 g; f; .5 O 1—A—ShunMV(n-6) -o— oocnlmwme) H—sIunMVIYotinumn-c) .ro- DOCAMVIYdinumn-e) %Constr1ctlon 991-593.818.913 % Constrlction owéfififi 8425 5 Figure 6.2. Contribution of aZ-adrenergic receptor-mediated components to NE overflow and neurogenic constriction of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of yohimbine (1.0 uM) in sham and DOCA-salt MA (A) and MV (B). Frequency-response curves for the contractile response before (control) and after application of yohimbine in sham and DOCA-salt MA(C) and MV (D). *indicates significantly different from NE oxidation current in control sham MA (MV) (P<0.05). #indicates significantly different from NE oxidation current in control DOCA-salt MA(MV) (P<0.05). Data are presented as mean 3: S.E.M. 142 Table 6.2. Maximum constriction (Em) and half maximum stimulation frequency (S50) for MA and MV from sham and DOCA-salt rats in the absence (control) and presence of yohimbine (Yo, 1.0 uM) and UK 14,304 (UK, 1.0 uM). All data are expressed as the mean i S.E.M and “11” value refers to the number of animals from which the data were obtained. Sham Artery DOCA-salt Artery Control (n=5) Yo (n=5) Control (n=6) Yo (n=6) S50 (Hz) 5.4 a 1.1 2.4 i 07“ 5.6 :t 0.9 2.5 3. 03* Em, (%) 33.9 :1: 1.9 30.0 :1: 2.3 33.0 i 2.7 29.4 d: 1.9 Control (n=5) UK (n=5) Control (n=6) UK (n=6) S50 (Hz) 5.8 :1: 0.6 11.3 :t 07“ 6.5 :1 0.9 7.9 :1: 0.7 ” 5mm) 36.9 d: 2.3 13.5 i 2.3“ 33.5 d: 2.5 20.4 :1: 1.9 *" Sham Vein DOCA-salt Vein Control (n=5) Yo (n=5) Control (n=6) Yo (n=6) S50 (Hz) 3.0 :1: 0.7 3.4 a: 0.8 1.5 :1: 0.4 1.4 i 0.7 Em, (%) 27.3 3: 2.3 21.9 a 2.9 21.7 :1: 1.9 20.0 3: 1.7 Control (n=5) UK (n=5) Control (n=4) UK (n=4) S50(Hz) 1.9308 10.1 a 1.1“ 2,431.1 9.341.5* 15mm) 31.9 a 2.3 11.9 i 1.6“ 28.9 :1: 1.7 15.9 :1: 2.6* &Indicates Significantly different from the S50 or EmaJlr in control sham MA(MV) (P<0.05). *lndicates significantly different fi'om the S50 or Em, in control DOCA-salt MA(MV) (P<0.05). ”Indicates significantly different from the S50 or Em in the UK 14,304 treated sham MA (P<0.05). 143 However, there was no difference between the contractile responses for sham and DOCA-salt MA. Yohimbine also significantly increased NE overflow in sham and DOCA-salt MV at low frequencies (1 to 7 Hz) but not at high frequencies in Figure 6.2B. In the presence of yohimbine, the elevated NE overflow for sham and DOCA-salt MV was not Significantly different. The nominal oxidation current increased by 38.3 % from 12.5 to 17.3 pA at 7 Hz (11 = 5) and by 76.7 % from 3.8 to 6.6 pA at 1 Hz (11 = 6) for sham MV, and by 39.6 % from 11.9 to 16.7 pA at 7 Hz (n = 6) and by 81.9 % from 4.2 to 7.6 pA at 1 Hz (n = 6) for DOCA-salt MV. Interestingly, the elevated NE overflow did not alter neurogenic constriction for both sham and DOCA-salt MV as seen in Figure 6.2D and Table 6.2. Figure 6.3 shows that UK 14,304, an orZ-adrenergic receptor agonist significantly inhibited NE overflow at all frequencies in sham and DOCA-salt MA and MV, which resulted in decreased constriction. However, NE overflow and the elicited contractile response for DOCA-salt MA were much less affected by UK 14,304 compared to the responses for sham MA (Figure 6.3A and C). In the presence of UK 14,304, the nominal oxidation current decreased by 70.0 % from 12.5 to 3.8 pA at 20 Hz for sham MA, whereas the nominal oxidation current decreased by only 31.1 % from 17.6 to 12.1 pA for DOCA-salt MA. The Ema, for DOCA-salt MA was also significantly greater than that for sham MA in the presence of UK 14,304 but there were no differences between the frequency response curves obtained for sham and DOCA-salt MV as seen in Figure 6.3D and Table 6.2. 144 A B 30 . Gammon-1) ao‘mamwm) 25. -mMA(n=5) 25q-mMV0F4) , @mmuwumo A Emwm1mw6) -20] -[XI:AMMK14.304(11=6) < ‘-oocAMwm1m(n=4) < 9.20 9; . , ‘é’ . 7 *' 15 Z % 5 10- f 0 f o 7% é 10‘ f . z 4 . 4 54 f f / 1 % g ‘ g 0. % % 0‘ g 1 FWa-lz) w-A—Slunmm-E) m—A—Sunwovfi) .—o—mm(n=5) i—O—mwflfi‘l fi—A—mmuuwlr—o c 351-A—Stunwm1mm-5) 5 30.-*-DOGAMK14304(n-6) o aol—o—mm‘lqu 1a 1 l a 2 20 2 20 o 1 o « o 15: O 15 3 10- 9" 1o. 5 5 - - W 0‘ W010 1° Figure 6.3. Contribution of (112-adrenergic receptor-mediated components to NE overflow and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). F requency-response for NE oxidation current before (control) and afier application of UK 14,304 (1.0 uM) in sham and DOCA-salt MA (A) and MV (B). Frequency-response curves for contractile response before (control) and after application of UK 14,304 in sham and DOCA-salt MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control Sham MA (MV) (P<0.05). “Indicates significantly different from NE oxidation current in control DOCA-salt MA (MV) (P<0.05). &Indicates significantly different from NE oxidation current in the drug treated sham MA (P<0.05). Data are mean i S.E.M. 145 Effects of NET on NE clearance and vasoconstriction in DOCA-salt MA and MV. Reuptake of NE by the norepinephrine transporter (NET) is the primary clearance mechanism responsible for removing the neurotransmitter from the junctional sites. To test whether NET was altered in DOCA-salt MA or MV, the effect of cocaine on NE overflow and the contractile response was investigated. Figures 6.4A and B show that cocaine (10 uM) increased the NE oxidation current for sham and DOCA-salt MA but was less influence on NE overflow for sham and DOCA-salt MV. Interestingly, the NE oxidation current of the DOCA-salt MA was greater than that for sham MA after cocaine treatment. The nominal oxidation current increased by 72.8 % from 13.7 to 23.7 pA at 20 Hz and by 73.4 % from 12.6 to 21.8 pA at 10 Hz for DOCA-salt MA, whereas the nominal oxidation current increased by 32.2 % from 12.7 to 16.8 pA at 20 Hz and by 41.2 % from 11.6 to 16.4 pA at 10 Hz for sham MA in the presence of cocaine. For sham and DOCA-salt MA, the slower clearance of NE from junctional sites in the presence of cocaine enhanced the contractile response, particularly at low frequencies. The contractile response curves in the presence of cocaine were left-shifted from the curve without the drug (control) (Figure 6.4C and Table 6.3). However, there was little effect on the contractile response of sham and DOCA-salt MV, since the negligible effect of cocaine on NB clearance of MV (Figure 6.4D and Table 6.3). Corticosterone, an extraneuronal transporter (uptake 2) blocker, also had no significant effect on NB overflow or the contractile response for either MA or MV (see also Chapter 5). 146 A B 40 40 Dammwe) 1 -ooc41m n= .Clslamwlwn EMMA/grams) WWIM g 301 -[II?AMN(bca'ne(n=7) ‘ QmESnmwmtw-n " . s #8. 3 J-WMWCocdnemfl) E a g 20- E20. 0 i (E 5 7/5 Z5 %E %E _ %= %= %= 10‘ %= 10- , = /= %5 ¢E as 1 4% 4E 45 25 f5 %= /= %= /= 0 £5 0 £5 45 £5 65 0 A 4o. 351 ,5 30 g *7 1 '4: 2 20' g 815 8 =3 «H as ‘1 O"' "'I 1 10 Figure 6.4. Contribution of reuptake transporter-mediated components to NE clearance and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). F requency-response for NE oxidation current before (control) and after application of cocaine (10 11M) in sham and DOCA-salt MA (A) and MV (B). Frequency-response curves for contractile response before (control) and after application of cocaine in sham and DOCA-salt MA (C) and MV (D). *lndicates significantly different from NE oxidation current in control sham MA (P<0.05). ”Indicates significantly different from NE oxidation current in control DOCA-salt MA (P<0.05). &Indicates Significantly different from NE oxidation current in the drug treated sham MA (P<0.05) Data are mean i S.E.M. 147 Table 6.3. Maximum constriction (Em) and half maximum stimulation frequency (S50) for MA and MV from sham and DOCA-salt rats in the absence (control) and presence of cocaine (Co, 10 nM) and the combined drug (cocaine + yohimbine). All data are expressed as the mean .+_ S.E.M and “11” value refers to the number of animals from which the data were obtained. Sham Artery DOCA-salt Artery Control (n=5) Cocaine (n=5) Control (n=6) Cocaine (n=6) S50 (Hz) 5.7 3 0.3 2.3 3 07“ 5.5 3 0.4 1.6 3 1.2* Emax (%) 35.7 is 1.6 33.3 i 2.1 31.3 3: 1.3 27.9 3: 2.7 Control (n=5) Co + Yo (n=5) Control (n=5) Co + Yo (n=5) S50 (Hz) 5.9 3 1.3 1.8 3 09“ 6.3 3 2.3 1.8 3 02* Emax (%) 29.6 :1: 1.7 28.0 :t 2.9 25.9 :h 2.3 21.6 :1: 1.9 Sham Vein DOCA-salt Vein Control (n=5) Cocaine (n=5) Control (n=5) Cocaine (n=5) S50 (Hz) 1.3 3 0.6 1.9 3 0.4 2.5 3 0.8 1.8 3 0.5 15...... (%) 24.6 3 1.3 24.8 3 1.9 23.9 3 1.7 20.6 3 2.1 Control (n=5) Co + Yo (n=5) Control (n=5) Co + Yo (n=5) S50 (Hz) 1.8 3 0.6 2.4 3 0.6 1.6 3 0.9 — Em...(%) 33.3 3 3.1 24.8 3 2.9“ 21.9 3 1.2 16.2 3 07* &indicates significantly different from the S50 or Em in control Sham MA(MV) (P<0.05). *Indicates Significantly different from the S50 or Emax in control DOCA-salt MA(MV) (P<0.05). 148 Effect of combined application of cocaine and yohimbine. The interrelationship between reuptake and prejunctional orZ-adrenergic receptors in DOCA-salt MA and MV was studied using the combined drug, cocaine (10 11M) and yohimbine (1.0 uM). Figure 6.5A shows that the combined application of cocaine and yohimbine markedly increased the oxidation current in sham and DOCA-salt MA compared with the effects of either of the individual drugs. There were no differences between the elevated oxidation currents obtained from sham and DOCA-salt MA. In the presence of the combined drug, the nominal oxidation current increased by 74.3 % from 16.0 to 28.0 pA at 20 Hz and by 316 % from 6.4 to 25.6 pA at 3 Hz for sham MA, and by 102 % from 16.3 to 33.0 pA at 20 Hz and by 279 % from 7.8 to 29.6 pA at 3 Hz for DOCA-salt MA. The increased junctional NE concentration caused elevated constrictions in sham and DOCA-salt MA, particularly at low frequencies, with left-shifted from the curve without the drug (control) (Figure 6.5C and Table 6.3). The elevated contractile responses for sham and DOCA-salt MA were not significantly different. The S50 value from Sham and DOCA-salt MA were 5.9 i 1.3 Hz and 6.3 :1: 2.3 Hz, before the combined drug treatment and 1.8 i 0.9 Hz and 1.8 i 0.2 Hz, after the combined drug treatment, respectively (Table 6.3.) However, the NE oxidation current increased significantly only at low frequencies in sham and DOCA-salt MV after the drug treatment but there were no differences between the elevated oxidation currents obtained from Sham and DOCA-salt MV. Though the NE concentration was increased significantly at low frequencies in MV, the contractile responses were significantly decreased in the presence of the drug (Figure 6.5D and Table 6.3). 149 A B ' 1383115511145) 60 m”(m 60« [jargon/(n=5) ‘ Eammvoqu 50- -DOCAMV(1=6) 250 -mWY°(,F6) .. Esra-nwwvows) 340: 3* * * E40: -mm1-Yo(m6) ‘5' 1 2 1 t 30 g 30 * :1 ‘ E :1 1 ’ ' o 20 E o 29 .=. 10 E 10 E 0 5. 0 5 mm mm CC 45 4‘" —A—SBHM(I‘F6’ J—A-Smflflmq 40: _O_mm(m 401—o-mwm-15) 35. —A—smnmc6+vo(u=5) c 35*“930WY01M 1 —o— mmvowq o 1—'— WWW”) 5 3°“. “*3 3° W “3% '5 25 l l 1 “fl , a) , A/ 1324 52* 8151 .\° 151 .\° 101 10 5 5; - ° '1 11 ' * * 7716 W01!) Figure 6.5. Contribution of uptake transporter/or2-adrenergic receptor-mediated components to NE clearance/release and neurogenic constrictions of sham and DOCA- salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of combined drug (cocaine (10 uM) + yohimbine (1.0 uM)) in sham and DOCA-salt MA (A) and MV (B). Frequency- response curves for contractile response before (control) and after application of the combined drug in sham and DOCA-salt MA (C) and MV (D). *Indicates significantly different from NE oxidation current in control sham MA(MV) (P<0.05). I’Indicates significantly different from NE oxidation current in control DOCA-salt MA(MV) (P<0.05). Data are mean i S.E.M. 150 Purinergic and adrenergic contribution to neurogenic constriction in DOCA-salt MA and MV. NE and ATP are major vasoconstrictor neurotransmitters released from sympathetic nerves. For the study of contributions of ATP and NE to neurogenic constrictions of DOCA-salt MA and MV, NE oxidation current and contractile response curves in sham and DOCA-salt tissues were constructed in the absence (control) and presence of prazosin (0.1 uM), an al-adrenergic receptor antagonist and pyridoxal- phosphate-6-azophenyl-2',4'-disulfonic acid (PPADS) (10 HM), a P2X receptor antagonist. Prazosin and PPADS block the constrictor effects of NE and ATP. Figures 6.6A and B show the evidence that the postjunctional receptor blockers did not alter NE oxidation current in both sham and DOCA-salt MA and MV. Prazosin reduced part of neurogenic constriction of sham MA but markedly inhibited the contractile response in DOCA-salt MA (Figure 6.6C). In Table 6.4, the Em in sham and DOCA-salt MA was 31.6 i 2.1 % and 33.7 i 1.9 %, before prazosin treatment and 23.4 i 2.4 % and 13.7 i 3.4 % afier prazosin treatment, respectively. The S50 in sham and DOCA-salt MA was 5.7 i 0.4 Hz and 5.8 i 0.9 Hz, before prazosin treatment and 5.9 i 1.0 and 11.3 i 2.7 Hz after prazosin treatment, respectively. Subsequent addition of PPADS completely blocked the neurogenic response remaining in the presence of prazosin. PPADS alone greatly reduced neurogenic constrictions in sham MA but PPADS had a little effect on responses in DOCA-salt MA. In contrast, prazosin mostly blocked constriction caused by NE during electrical nerve stimulation in sham and DOCA-salt MV but PPADS alone had no effect on the contractile responses in sham and DOCA-salt MV (Figure 6.6D). 151 A B aoLCIMMMfl") sailfismnwowe) ‘-DOGAMA(nfl ‘-IIEAW(II=II) 5&9:anan 5EMW(M3) 2 .-mGAWAIB(n-6) 2 ,-mmmn(mq 9; 20 5 20 E . ‘é’ . V = g 15' Z g 15 5 .5: 3 ‘ Z 3 1 = Z: / = /= o 10‘ Z 0 1o 5 7: Z: ‘ %= g d E g! g; 5‘ Z: Z 5. 5 ZE ZE E / E /= /'=' i Z: Z . = /= /= 0- Z: Z 0 ,5 Z5 Z5 45 ,_A_ Sunmumfl) 401—0— mumme) 35.—A—sranuwnmmme) 30"" mwnanimmn .—-— Minoan» % Constriction % Constriction <9- 91- <3- <75- 83?. 8 "| v V V V V 1 mat)“ Figure 6.6. Contribution of oil-adrenergic receptor and P2X-mediated components to NE release and neurogenic constrictions of sham and DOCA-salt mesenteric arteries (MA) and mesenteric veins (MV). Frequency-response for NE oxidation current before (control) and after application of prazosin (0.1 uM) or/and PPADS (10 pM) in sham and DOCA-salt arteries (A) and veins (B). Frequency-response curves for contractile response before (control) and after application of the drug in sham and DOCA-salt MA (C) and MV (D). Data are mean t S.E.M. 152 Table 6.4. Maximum constriction (Em) and half maximum stimulation frequency (S50) in MA and MV from sham and DOCA-salt rat in the absence (control) and presence of prazosin (0.1 uM) and PPADS (10 uM) treatment. All data are expressed as the mean i SEM and “n” values refer to the number of animals from which the data were obtained. Sham Artery DOCA-salt Artery Control (n=5) Prazosin (n=5) Control (n=5) Prazosin (n=5) 350012) 5.7 i 0.4 5.9 a: 1.0 5.8 :l: 0.9 11.7 :h 22*” 15mm) 31.6 :t 2.1 23.4 i 2.4“ 33.7 2!: 1.9 13.7 :h 3.4“ I Sham Vein DOCA-salt Vein Control (n=5) Prazosin (n=5) Control (n=5) Prazosin (n=5) S50 (Hz) 1.55:0.6 — 1.9105 11.5:t3.1* Ew(%) 27.4 :t 1.7 5.7 :l: 3.1“ 24.9 d: 1.9 8.5 :t 1.7* &Indicates significantly different from the S50 or Em in control sham MA(MV) (P<0.05). *Indicates significantly different fiom the S50 or Em in control DOCA-salt MA(MV) (P<0.05). ”Indicates significantly different from the S50 or Em, in the prazosin treated sham MA (P<0.05). 153 6.3 Discussion All NE overflow and evoked contractile responses for both MA and MV induced by electrical stimulation were blocked completely by TTX; a voltage-gated sodium channel blocker (see Chapter 4). This result indicates that NE overflow and contractile response were nerve-mediated. In this Chapter, NE overflow and vasoconstriction for DOCA-salt MA and MV were investigated in order to assess changes that might be associated with hypertension. A carbon fiber was used to monitor NE in this work rather than a diamond microelectrode because it was found that several of the drugs employed were oxidized at the positive detection potential needed to detect NE at the diamond micro electrode. Increase NE overflow in DOCA-salt MA and altered neurogenic responses in DOCA-salt MV. In the early stage of certain forms of experimental hypertension [69] as well as in hypertensive patients [203, 17, 48], sympathetic traffic to the cardiovascular system is elevated. Previous studies have shown that there was increased sympathetic neuronal activity and NE overflow from sympathetic nerves in DOCA-salt rats [77, 69, 68]. However, NE overflow was not monitored in real time and not measured directly in these studies. The present study revealed that NE overflow was significantly increased in DOCA-salt MA compared with sham MA. However, little change was seen in the overflow for DOCA-salt MV. This result is different from previous in vitro work that showed an increased NE overflow in DOCA-salt MV [68]. This difference may be accounted for by methodological differences. In this study, NE overflow was measured locally at the vessel surface near the point of focal stimulation rather than as accumulative overflow from an entire tissue bed. 154 The increased NE overflow in DOCA-salt MA could result from increased NE concentrations released from nerve endings, reduced clearance by NE transporter (NET), dysfunction of the prejunctional aZ-adrenergic autoreceptors, reduced extracellular enzymatic metabolism or some contribution of these. It was hypothesized that dysfunction of the prejunctional aZ-adrenergic autoreceptors and/or NET located at the sympathetic nerve terminals is the cause for the increased NE overflow in DOCA-salt MA. Impaired prejunctional (1.2-adrenergic autoreceptors in DOCA-salt MA. Prejunctional aZ-adrenergic autoreceptors mediate the inhibition of NE release from sympathetic terminals, constituting an important autoregulatory system for sympathetic tone [211, 52]. Previous studies showed that prejunctional regulation of transmitter release by a2-adrenergic autoreceptor was altered in experimental rat models of hypertension [77, 80]. However, either impaired [79, 68] or unaltered [28] a2-adrenergic autoreceptors in the isolated MA of DOCA-salt rats have been reported. My study shows that blockade of these receptors with an a2-adrenergic receptor antagonist, yohimbine, resulted in increased NE overflow in both sham and DOCA-salt MA, because the negative feedback system was interrupted by the drug. However, yohimbine was less effective of increasing NE overflow in DOCA-salt MA compared to sham MA. Moreover, an aZ-adrenergic receptor agonist, UK 14,304 inhibited NE overflow in sham MA more than DOCA-salt MA. This result is consistent with studies showing a failure of yohimbine to increase plasma NE level [80, 81] and nerve stimulation—induced NE overflow in the mesenteric vasculature of DOCA-salt rats [49, 79, 68]. Taken together 155 these data suggest that aZ-adrenergic autoreceptors may be impaired in DOCA-salt MA. However, yohimbine or UK 14,304 did affect NE overflow less in sham and DOCA-salt MV compared with sham MA. Also, there were no differences in NE overflow in the presence of the drugs between sham and DOCA-salt MV. These experiments indicate that NE release and its regulation through the prejunctional a2-adrenergic autoreceptor (prejunctional negative feedback system) does not seem to be altered in the isolated MV of the DOCA-salt hypertensive rat. Westfall et al. also observed that there was no significant elevated NE release from the portal vein and no attenuation of the yohimbine effect in DOCA-salt hypertensive rats [77]. However, the portal vein may not be an appropriate blood vessel to study hypertension associated with changes in the function of peri-venous sympathetic nerves because the portal vein makes only a small contribution to total vascular capacitance. Upregulation of NE reuptake in DOCA-salt MA. NET activity is responsible for removing NE from the synaptic cleft once released from the nerve terminal [72]. Reduced uptake has been proposed to explain the elevated NE associated with DOCA-salt hypertension [87]. However, unaltered neuronal reuptake of NE in DOCA-salt hypertensive rats [67] or enhanced NE clearance rate in chronically hypertensive animals [212, 92] and in human with essential hypertension [17] have been reported. The unaltered or enhanced neuronal uptake reflected that elevated NE overflow was due to increased NE secretion rate subsequent to attenuation of the a2-presynaptic or local inhibitory mechanism [205, 69]. However, the explanation for reduced or increased NE reuptake in smooth muscle of DOCA-salt blood vessels afier the development of hypertension is still not clear. The present study showed that blockade of NET, by 156 cocaine, resulted in an increased NE overflow in both sham and DOCA-salt MA but the magnitude of increase in DOCA-salt MA was bigger than sham MA at all frequencies and significantly greater at high fiequencies (10 and 20 Hz). However, when 0L2- adrenergic autoreceptor and NET were blocked by the combined drug (yohimbine + cocaine), NE overflow markedly increased in both sham and DOCA-salt MA but not significantly different between sham and DOCA-salt MA. The results suggest that cocaine-sensitive reuptake mechanism seems to be upregulated in DOCA-salt MA. Altered neuronal reuptake of NE in DOCA-salt MA may be caused by increased NE release due to impaired a2-adrenergic autoreceptors. The increased NE release may stimulate increased NET expression in DOCA-salt MA [31]. However, cocaine affected NE clearance less in sham and DOCA-salt MV. The combined drug increased NE overflow at low frequencies in sham and DOCA-salt MV due to blockade of autoinhibition by yohimbine not by cocaine. The result indicates that NE clearance for seems to be mainly by diffusion instead of through NET for sham and DOCA-salt MV. Postjunctional (1.2-adrenergic receptors in sham and DOCA-salt MV. The contractile responses of sham and DOCA-salt MA at low frequencies (< 10 Hz) in the presence of yohimbine and/or cocaine increased due to the elevated NE overflow and slow rate of NE clearance in junctional cleft by yohimbine and cocaine, respectively while the contractile response at high frequency ( S 10 Hz) relatively was unaffected by the drug treatment in both sham and DOCA-salt MA. This result suggests that the junctional NE concentration at the higher frequency may be sufficient to saturate receptor sites. 157 Elevated NE overflow in the presence of yohimbine, had little effect on neurogenic constriction or even decreased the maximum constriction (Em) in MV from sham and DOCA-salt hypertensive rats. Moreover, UK 14,304 preconstricted over 10 % in both sham and DOCA-salt MV. The result suggests that postjunctional a2-adrenergic receptors may also regulate neurogenic constriction of MV in sham and DOCA-salt rats. Previous studies have proposed that a2-adrenergic receptors also located postj unctionally and mediate contractile response in blood vessels [213-215]. However, UK 14,304 did not preconstrict sham and DOCA-salt MA, suggesting that a2-adrenergic receptors may not be expressed by MA and are not involved in contractile response to NB. This was agreement with previous studies in rat and mice MA [31, 216]. Alternation vasculature and/or postsynaptic adrenergic dysfunction in DOCA-salt MA and MV. The present and previous work showed that first, MV were more sensitive than MA to nerve stimulation, which was based on the observation that the S 50 was lower in sham and DOCA-salt MV compared with sham and DOCA MA [22, 31]. Therefore, changes in neuroeffector transmission to vein in hypertension are important because veins are more sensitive to the vasoconstrictor effects of sympathetic nerve stimulation than arteries [22, 31]. Second, contractile response in MA was mediated mainly by ATP because PPADS, P2X receptor antagonist greatly reduced neurogenic constriction of sham MA while NE makes a major contribution to neurogenic constriction of DOCA-salt MA because the constriction of DOCA-salt MA was greatly reduced by prazosin. There was an increase in the adrenergic contribution to neurogenic constrictions in DOCA-salt MA compared to that occurring in sham MA. This result is consistent with other studies showing that postjunctional al-adrenergic function become dominant in DOCA-salt and 158 SHR rats [80] and strikingly increased in density of or] binding sites in DOCA-salt MA [217, 31]. Previous study suggested that the increased adrenergic transmission may be due to increased NE release but there was no changes in postjunctional sensitivity of DOCA-salt MA to NE [31]. The presence of an imbalance in postjunctional adrenoceptor functions may be associated with a change in neuroeffector mechanisms in DOCA-salt MA and promotes the pressor effects of the sympathetic system [80]. However, elevated NE overflow did not alter the contractile response in DOCA-salt MA, though NE made a major contribution to neurogenic constriction of DOCA-salt MA and even, there was no change in reactivity of al-adrenergic receptors to NE in DOCA-salt MA. The elevated NE overflow caused an increase in NE uptake in DOCA-salt MA, the increased uptake may offset the vasoconstrictor effects of elevated NE release in DOCA-salt MA. Prazosin alone mostly blocked neurogenic constriction in both sham and DOCA— salt MV. Therefore, NE mediates neurogenic constriction of sham and DOCA-salt MV. Interestingly, my study shows that the E,,m value was significantly lower in DOCA-salt MV compared with sham MV, though NE overflow was not significantly different between sham and DOCA-salt MV. Therefore, the reduced contractile response was not due to the decreased NE overflow in DOCA-salt MV from my data. Luo et al. suggested that Oil-adrenergic receptor sensitivity to NE was decreased in DOCA-salt MV due to downregulation of Oil-adrenergic receptors in DOCA-salt veins caused by the increased sympathetic nerve activity [80, 176, 31]. Also, similar decrease in NE sensitivity occurs in hand veins taken from human hypertensive subjects [209]. Other mechanisms include altered vasculature structure in DOCA-salt rats may also account for this result. The decreased compliance of DOCA-salt MV may cause redistribution of blood toward the 159 heart, leads to increased cardiac output to arterial side and finally will result in increased arterial blood pressure [218, 81]. 6.4 Conclusion The present study demonstrates that prej unctional a2-adrenergic autoreceptors and NE reuptake play a greater role in regulating NE availability at the neuroeffector junction in MA than in MV. The elevated NE overflow from perivascular sympathetic nerves in DOCA-salt MA is likely due to altered a2-adrenergtic autoreceptors. The elevated NE overflow caused an increase in NE reuptake in DOCA-salt MA and the increased uptake may offset the vasoconstrictor effects of elevated NE release in these tissues. NE mediated venous constriction by acting at postjunctional Oil-adrenergic receptors in MV from sham and DOCA-salt rats but contractile response was decreased in DOCA-salt MV due to desensitized al-adrenergic receptors. ATP, acting at P2 receptors, was the dominant vasoconstrictor transmitter in small MA from sham rats. In contrast, NE made a major contribution to neurogenic constriction of DOCA-salt MA due to an increase in the adrenergic contribution to neurogenic constrictions. However, changed contractile response in DOCA-salt rat may be also due to the altered vasculature in DOCA-salt rats. Therefore, there may be altered vasculature and/or postjunctional dysfunction in DOCA-salt MA and MV. All the observations suggest that there was differential neural control of MA and MV and the neural control of MA artery and MV were altered in DOCA-salt rats. The alternation in the local mechanisms modulating sympathetic neuroeffector mechanisms in hypertension will impact overall hemodynamics. 160 Chapter 7 Conclusions 1. Advantages of Diamond Microelectrode for the In Vitro Electroanalytical Measurement. The fabrication, electrochemical characterization and application of boron-doped diamond microelectrodes for the in vitro continuous amperometric measurement of the vasoconstrictor neurotransmitter, norepinephrine (N E), in Chapters 2 - 4. The diamond microelectrode is formed by depositing a high quality, thin film of electrically conducting diamond on a sharpened 76 um diam Pt wire. The electrode exhibits a low and stable background current over a wide potential range that is independent of the solution pH. Additionally, the bare electrode provides high resistance to deactivation and fouling during exposure to laboratory atmosphere and complex biological environments, at least in part, because of the non-polar, low oxygen, sp3-bonded carbon surface on which weak adsorption of polar biomolecules and other contaminants occurs. In contrast, the carbon fiber microelectrode exhibits a pH-dependent background current response with evidence for electroactive surface carbon-oxygen functional groups, and deactivates irreversibly during laboratory air exposure and contact with tissue. Diamond clearly provides superior response sensitivity precision and stability making it useful for sensitive and stable electroanalytical measurements in complex biological environments. 161 II. In Vitro Monitoring of Endogenous NE Overflow with a Diamond Microelectrode and Vasoconstriction. Continuous amperometry with the diamond microelectrode and video imaging were used the elicited for the first time to simultaneously record endogenous NE overflow and the evoked contractile response of a rat mesenteric artery. Electrical stimulation of sympathetic nerve endings elicited a transient oxidation current that was attributed to endogenous NE. NE overflow was elicited by electrical stimulation at frequencies between 1-60 Hz, with a maximum response seen at 20 Hz. Confirmation that the oxidation current was in fact associated with endogenous NE came fi'om the results of several drugs. Tetrodotoxin (TTX), a voltage-dependent sodium channel antagonist which blocks nerve conduction, abolished both the oxidation current and the arterial constriction. The a2-adrenergic autoreceptor antagonist, yohimbine, caused an increase in the oxidation current and the corresponding constriction. The addition of cocaine, a monoamine reuptake blocker, caused both the oxidation current and the contractile response to increase. These results, combined with the fact that the hydrodynamic voltammetric Em for endogenous NE was identical to that found for a standard solution confirmed that the oxidation current was due to the neurotransmitter. The TTX result also suggested that NE overflow and the contractile response were mediated neurogenically. The uncoated diamond microelectrode provided a sensitive, reproducible and stable oxidation current response that varied with the stimulation frequency and pulse number in a predictable manner, similar to that for a carbon fiber microelectrode. Furthermore, this study demonstrated that continuous amperometric monitoring of NE with a diamond microelectrode and video imaging of vascular diameter allow for real time local 162 measurement of the temporal relationship between nerve stimulated NE overflow and arterial constriction. III. Differential Sympathetic Neuroeffector Transmission to Mesenteric Arteries (MA) and Veins (MV) from Rats. The dual measurement technique was applied to investigate differences in sympathetic neuroeffector transmission in normal rat MV and MA. A better understanding of the firnctional differences is important because arteries and veins make different contributions to overall hemodynamics. The results indicated the NE transporter (NET) and prej unctional a2-adrenergic autoreceptors play a more prominent role in controlling NE release at neuroeffector junction in MA than in MV. Therefore, NE overflow in MV exceeded that in MA. These conclusions are based on following results: i) NE overflow in MA was more strongly enhanced by the a2-adrenergic receptor antagonist, yohimbine and significantly attenuated by a2-adrenergic receptor agonist, UK 14,304, compared to MV; and ii) NE overflow in MA increased more significantly in the presence of cocaine or the combined drug (cocaine + yohimbine) compared to MV. Neurogenic constriction of MV was mediated by NE acting at al-adrenergic receptors while that of MA was mediated by ATP acting at P2X receptors as well as NE. The conclusion is based on the observation that the P2 receptor antagonist, PPADS, reduced the contractile response in MA but not in MV while the al-adrenergic receptor antagonist, prazosin, blocked the contractile response in MV but only slightly in MA. Glyoxylic acid (GA)-induced fluorescence images also showed that there are differences in the arrangement of sympathetic nerves in MA and MV. The perisympathetic nerve 163 plexus in MV consisted of individual varicose axons largely arranged circumferentially along the vessel, denser, more mash-like network was observed for MA. Taken together, these results suggest that there are fundamental differences in sympathetic neuroeffector transmission to arteries and veins, which contribute to their different hemodynamic functions in regulation of arterial and venous tone. IV. Altered Sympathetic Neuroeffector Transmission to MA and MV in DOCA-salt Hypertensive Rats Small MA and MV in vitro preparations from DOCA-salt hypertensive rats were used what junctional differences are associated with the disease state. NE overflow in DOCA-salt MA was less attenuated by UK 14,304, an a2-adrenergic receptor agonist, and less elevated by yohimbine while cocaine increased NE overflow to a greater extent compared to sham MA. However, the combined drug (cocaine + yohimbine) produced a greater increase in NE overflow in both sham and DOCA-salt MA and elevated NE overflow by the drug was not different between sham and DOCA-salt MV. These results suggest that sympathetic nerves associated with DOCA-salt MA released more NE than sham MA due to impaired prejunctional a2-adrenergic autoreceptors. The elevated NE overflow caused an increase in NE reuptake in DOCA-salt MA and the increased clearance may offset the vasoconstrictor effects of elevated NE release in these tissues. In contrast, NE overflow was unchanged between sham and DOCA-salt MV. PPADS greatly reduced neurogenic responses in sham but not DOCA-salt MA, whereas prazosin inhibited response more in DOCA-salt MA compared to sham MA. The result suggests that NE made a major contribution to neurogenic constriction of DOCA- 164 salt MA because the elevated NE release increased the adrenergic contribution to the constriction of DOCA-salt MA. Contractile response in sham and DOCA-salt MV was mostly blocked by prazosin but not by PPADS. Neurogenic constriction in MV from sham and DOCA-salt rat was more sensitive than that in sham and DOCA-salt MA. However, the maximum contractile response was decreased in DOCA-salt MV compared to sham MV. It may be due to desensitization of al-adrenergic receptors and/or due to the altered vasculature in DOCA-salt MV. The observations suggest that the altered sympathetic neural control of both arteries and veins in hypertension will impact overall blood pressure. 165 References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] MacMahon, S.; Peto, R.; Cutler, J .; Collins, R.; Sorlie, P.; Neaton, J .; Abbott, R.; Godwin, J .; Dyer, A.; Stamler, J. Lancet 1990, 335, 765-74. Burt, V. L.; Whelton, P.; Roccella, E. J.; Brown, C.; Cutler, J. A.; Higgins, M.; Horan, M. J .; Labarthe, D. Hypertension 1995, 25, 305-13. Westfall, T. C.; Meldrum, M. J. Ann Rev Pharmacol Toxicol 1985, 25, 621-41. Beevers, G.; Lip, G. Y.; O'Brien, E. ij 2001, 322, 912-6. Guyton, A. C. Textbook of Medical Physiology (8” ed. ); Saunders: Philadelphia, PA, 1991. Tomanek, R. J .; Barlow, P. A. Hypertension 1990, 15, 225-33. Lund-Johansen, P. J Cardiovasc Pharmacol 1989, 14 Suppl 10, S7-13. McDonough, A. A.; Leong, P. K.; Yang, L. E. Ann N Y Acad Sci 2003, 986, 669- 77. Abboud, F. M.; Floras, J. S.; Aylward, P. E.; Guo, G. B.; Gupta, B. N.; Schmid, P. G. Blood Vessels 1990, 27, 106-15. Katusic, Z. 8.; Shepherd, J. T. Hypertension 1991, 18, 86-92. Burton, A. C. Physiol Rev 1954, 34, 619-42. Nilsson, H. Acta Physiol Scand Suppl 1985, 541, 1-34. Hainsworth, R. Rev Physiol Biochem Pharmacol 1986, 105, 101-73. Mark, A. L. Hypertension 1984, 6, 69-73. Schiffrin, E. L.; Lariviere, R.; Li, J. S.; Sventek, P. J Vasc Res 1996, 33, 235-48. Heagerty, A. M.; Aalkjaer, C.; Bund, S. J.; Korsgaard, N.; Mulvany, M. J. Hypertension 1993, 21, 391-7. Aalkjaer, C.; Heagerty, A. M.; Petersen, K. K.; Swales, J. D.; Mulvany, M. J. Circ Res 1987, 61, 181-6. Vial, J. H.; Yong, A. C.; Boyd, G. W. Clin Exp Pharmacol Physiol 1982, 9, 309- 13. 166 [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [311 [32] [33] [34] [35] [36] [37] Lee, R. M.; Smeda, J. S. Can J Physiol Pharmacol 1985, 63, 392-401. Dampney, R. A.; Coleman, M. J .; Fontes, M. A.; Hirooka, Y.; Horiuchi, J .; Li, Y. W.; Polson, J. W.; Potts, P. D.; Tagawa, T. Clin Exp PharmacolPhysi012002, 29, 261-8. Lefltowitz, R. J .; Hoffman, B. B.; Taylor, P. Goodman & Gilman's The Pharmacological Basis of Therapeutics, 9th ed.; McGraw-Hill, 1996. Kreulen, D. L. Am J Physiol 1986, 251, H1267-75. Burnstock, G.; Costa, M. Adrenergic neurons. Their organization, function and development in the peripheral nervous system; Chapman and Hall: London, 197 5. Luff, S. E.; McLachlan, E. M.; Hirst, G. D. J Comp Neurol 1987, 25 7, 578-94. Luff, S. E.; McLachlan, E. M. Blood Vessels 1989, 26, 95-106. F urness, J. B.; Marshall, J. M. J Physio] 1974, 239, 75-88. Guyton, A. C.; Coleman, T. G.; Granger, H. J. Annu Rev Physiol 1972, 34, 13-46. Ekas, R. D., Jr.; Lokhandwala, M. F. Am J Physiol 1980, 239, R303-8. Esler, M.; Jennings, G.; Komer, P.; Willett, 1.; Dudley, F.; Hasking, G.; Anderson, W.; Lambert, G. Hypertension 1988, 11, 3-20. Greenway, C. V. Fed Proc 1983, 42, 1678-84. Luo, M.; Hess, M. C.; Fink, G. D.; Olson, L. K.; Rogers, J.; Kreulen, D. L.; Dai, X.; Galligan, J. J. Auton Neurosci 2003, 104, 47-57. Rosenblum, J. D.; Boyle, C. M.; Schwartz, L. B. Surg Clin North Am 1997, 77, 289-306. Jacobson, E. D. Physiologist 1982, 25, 439-43. Adams, R. N. Anal Chem 1976, 48, 1126A-38A. Westfall, T. C.; Carpentier, S.; Chen, X.; Beinfeld, M. C.; Naes, L.; Meldrum, M. J. J Cardiovasc Pharmacol 1987, 10, 716-22. Gitterman, D. P.; Evans, R. J. Br J Pharmacol 2001, 132, 1201-8. Kitajima, S.; Ozaki, H.; Karaki, H. Br J Pharmacol 1993, 110, 263-8. 167 [38] [39] [40] [41 l [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] Pablo Huidobro-Toro, J .; Veronica Donoso, M. Eur J Pharmacol 2004, 500, 27- 35. Phillips, J. K.; McLean, A. J .; Hill, C. E. Br J Pharmacol 1998, 124, 1403-12. Tatemoto, K. Proc Natl Acad Sci U S A 1982, 79, 5485-9. Tatemoto, K.; Carlquist, M.; Mutt, V. Nature 1982, 296, 659-60. Fried, G.; Lundberg, J. M.; Theodorsson-Norheim, E. Acta Physiol Scand 1985, 125, 145-54. Lundberg, J. M. Pharmacol Rev 1996, 48, 113-78. Bradley, E.; Law, A.; Bell, D.; Johnson, C. D. Am J Physiol Heart Circ Physiol 2003, 284, H2007-14. Burnstock, G. Curr Opin Pharmac012004, 4, 47-52. Eisenhofer, G.; Saigusa, T.; Esler, M. D.; Cox, H. S.; Angus, J. A.; Dorward, P. K. Am J Physiol 1991, 260, R824-32. Eisenhofer, G. Pharmacol T her 2001, 91, 35-62. Esler, M.; Jennings, G.; Lambert, G.; Meredith, 1.; Home, M.; Eisenhofer, G. Physiol Rev 1990, 70, 963-85. Langer, S. Z. Pharmacol Rev 1980, 32, 337-62. Stjarne, L.; Astrand, P.; Bao, J. X.; Gonon, F.; Msghina, M.; Stjarne, E. Adv Sec Mess Phosphoprotein Res 1994, 29, 461-96. Westfall, D. P.; Todorov, L. D.; Mihaylova-Todorova, S. T. J Pharmacol Exp Ther 2002, 303, 439-44. Starke, K. J Neurochem 2001, 78, 685-93. Westfall, T. C.; Yang, C. L.; Rotto-Perceley, D.; Macarthur, H. J Auton Pharmacol 1996, 16, 345-8. Bobalova, J.; Mutafova-Yambolieva, V. N. Clin Exp Pharmacol Physiol 2001, 28, 397-401. Donoso, M. V.; Carvajal, A.; Paredes, A.; Tomic, A.; Koenig, C. S.; Huidobro- Toro, J. P. Peptides 2002, 23, 1663-71. 168 [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] Donoso, M. V.; Miranda, R.; Briones, R.; Irarrazaval, M. J .; Huidobro-Toro, J. P. Peptides 2004, 25, 53-64. Shi, A. G.; Ahmad, S.; Kwan, C. Y.; Daniel, E. E. J Cardiovasc Pharmacol 1990, 15, 515-26. Minneman, K. P. Pharmacol Rev 1988, 40, 87-119. Piascik, M. T.; Soltis, E. E.; Piascik, M. M.; Macmillan, L. B. Pharmacol Ther 1996, 72, 215-41. Evans, R. J .; Surprenant, A. Br J Pharmacol 1992, 106, 242-9. Todorov, L. D.; Mihaylova-Todorova, S.; Craviso, G. L.; Bjur, R. A.; Westfall, D. P. J Physiol 1996, 496 (Pt 3), 731-48. Okamoto, K.; Aoki, K. Jpn Circ J 1963, 27, 282-93. Friedman, R. Health Psychol 1988, 7, 149-58. Schenk, J .; McNeill, J. H. J Pharmacol Toxicol Methods 1992, 2 7, 161-70. Hebden, R. A.; Todd, M. B.; Tang, C.; Gowen, B.; McNeil], J. H. Am J Physiol 1990, 258, R1042-50. Seeley, R. R.; Stephens, T. D.; Tate, P. In : Anatomy & Physiology; McGraw- Hill: Boston, MA, 1998. Drolet, G.; Bouvier, M.; de Champlain, J. J hypertens 1989, 7, 237-42. Luo, M.; Fink, G. D.; Lookingland, K. J.; Morris, J. A.; Galligan, J. J. Am J Physiol Heart C irc Physiol 2004, 286, H1558-64. De Champlain, J.; Eid, H.; Drolet, G.; Bouvier, M.; Foucart, S. Canadian J Physio] Pharmacol 1989, 6 7, 1140-5. Rumantir, M. S.; Kaye, D. M.; Jennings, G. L.; Vaz, M.; Hastings, J. A.; Esler, M. D. Hypertension 2000, 36, 824-9. Anderson, E. A.; Sinkey, C. A.; Lawton, W. J.; Mark, A. L. Hypertension 1989, 14, 177-83. Cabassi, A.; Vinci, 8.; Quartieri, F.; Moschini, L.; Borghetti, A. Hypertension 2001, 37, 698-702. 169 [73] [74] [75] [76] [77] [78] [79] [30] [81] [82] [33] [84] [85] [36] [87] [88] [39] Esler, M.; Lambert, G.; Jennings, G. Clin Exp Hypertens A 1989, 11 Suppl 1, 75- 89. Ferrier, C.; Jennings, G. L.; Eisenhofer, G.; Lambert, G.; Cox, H. S.; Kalff, V.; Kelly, M.; Esler, M. D. J Hypertens 1993, 1 1, 1217-27. Brody, M. J .; Dorr, L. D.; Shaffer, R. A. Am J Physiol 1970, 219, 1746-50. Tsuda, K.; Kuchii, M.; Nishio, 1.; Masuyama, Y. Clin Exp Hypertens A 1986, 8, 259-75. Westfall, T. C.; Carpentier, S.; Naes, L.; Meldrum, M. J. Clin Exp hypertens 1986, 8, 221-37. Esler, M.; Jennings, G.; Biviano, B.; Lambert, G.; Hasking, G. J Cardiovasc Pharmacol 1986, 8 Suppl 5, 839-43. Tsuda, K.; Tsuda, S.; Nishio, 1.; Masuyama, Y. Jpn Heart J 1989, 30, 231-9. de Champlain, J. J Hypertens Suppl 1990, 8, 877-85. Moreau, P.; Drolet, G.; Yamaguchi, N.; de Champlain, J. Am j hypertens 1995, 8, 287-93. Westfall, T. C.; Meldrum, M. J.; Badino, L.; Earnhardt, J. T. Hypertens 1984, 6, 267-74. Tsuda, K.; Kuchii, M.; Nishio, I.; Masuyama, Y. Jpn Circ J 1987, 51, 25-32. Foucart, S.; Nadeau, R.; de Champlain, J. Can J Physio] Pharmacol 1987, 65, 550-7. Bobalova, J .; Mutafova-Yambolieva, V. N. J Auton Pharmacol 2001, 21, 47-55. Esler, M.; Jackman, G.; Bobik, A.; Leonard, P.; Kelleher, D.; Skews, H.; Jennings, G.; Komer, P. Hypertension 1981, 3, 149-56. Rascher, W.; Schomig, A.; Dietz, R.; Weber, J .; Gross, F. Eur J Pharmacol 1981, 75, 255-63. Lin, M. S.; McNay, J. L.; Shepherd, A. M.; Keeton, T. K. Clin Pharmacol Ther 1983, 34, 466-73. Lambert, G. W.; Ferrier, C.; Kaye, D. M.; Kalff, V.; Kelly, M. J.; Cox, H. S.; Turner, A. G.; Jennings, G. L.; Esler, M. D. Blood Press 1994, 3, 55-66. 170 [90] [91] [92] [93] [94] [95] [96] [97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] de Champlain, J .; Krakoff, L. R.; Axelrod, J. Life Sci 1966, 5, 2283-91. Garrett, J.; Castro-Tavares, J .; Branco, D. Blood Vessels 1981, 18, 100-9. Whall, C. W., Jr.; Myers, M. M.; Halpem, W. Blood Vessels 1980, 17, 1-15. Zsoter, T. T.; Wolchinsky, C. Clin Invest Med 1983, 6, 191-5. Longhurst, P. A.; Rice, P. J .; Taylor, D. A.; Fleming, W. W. Hypertension 1988, 12, 133-42. Goldstein, D. S.; Zimlichman, R.; Stull, R.; Keiser, H. R.; Kopin, I. J. Hypertension 1986, 8, 471-5. Izzo, J. L., Jr.; Licht, M. R.; Smith, R. J .; Larrabee, P. S.; Radke, K. J .; Kallay, M. C. Am J Cardiol 1987, 60, 303-8. Grassi, G.; Colombo, M.; Seravalle, G.; Spaziani, D.; Mancia, G. Hypertension 1998, 31 , 64-7. Ramme, D.; Illes, P.; Spath, L.; Starke, K. Naunyn Schmiedebergs Arch Pharmacol 1986, 334, 48-55. Cox, S. L.; Story, D. F.; Ziogas, J. Br J Pharmacol 1996, 119, 976-84. Smyth, L.; Bobalova, J .; Ward, S. M.; Keef, K. D.; Mutafova-Yambolieva, V. N. Auton Neurosci 2000, 86, 18-29. McAllen, R. M.; Malpas, S. C. Clin Exp Pharmacol Physiol 1997, 24, 791-9. Stemson, A. W.; McCreery, R.; Feinberg, B.; Adams, R. N. Journal of Electroanal Chem Interfacial Electrochem 1973, 46, 313-21. Wightman, R. M. Science 1988, 240, 415-20. Kawagoe, K. T.; Zimmerman, J. B.; Wightman, R. M. J Neurosci Methods 1993, 48, 225-40. Wightman, R. M.; Robinson, D. L. J Neurochem 2002, 82, 721-35. Suaud-Chagny, M. F.; Buda, M.; Gonon, F. Advances in the Biosciences (Oxford) 1991, 82, 341-2. Brun, P.; Suaud-Chagny, M. F.; Gonon, F.; Buda, M. Neuroscience 1993, 52, 961-72. 171 [108] [109] [110] [111] [112] [113] [114] [115] [116] [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] Robinson, D. L.; Venton, B. J.; Heien, M. L.; Wightman, R. M.; Troyer, K. P.; Zhang, H.; Garris, P. A.; Phillips, P. B.; Sulzer, D.; Michael, D. J.; Budygin, E. A.; Bergstrom, B. P.; Rebec, G. V. Clin Chem 2003, 49, 1763-73. Perez, X. A.; Andrews, A. M. Anal Chem 2005, 77, 818-26. Troyer, K. P.; Heien, M. L.; Venton, B. J.; Wightman, R. M. Curr Opin Chem Bi012002, 6, 696-703. Leszczyszyn, D. J.; Jankowski, J. A.; Viveros, O. H.; Diliberto, E. J ., Jr.; Near, J. A.; Wightman, R. M. J Neurochem 1991, 56, 1855-63. Travis, E. R.; Wightman, R. M. Annual Review of Biophysics and Biomolecular Structure 1998, 27, 77-103. Bruns, D.; Jahn, R. Nature 1995, 377, 62-5. Kovach, P. M.; Ewing, A. G.; Wilson, R. L.; Wightman, R. M. J Neurosci Methods 1984, 10, 215-27. Mermet, C.; Suaud-Chagny, M. F.; Gonon, F. Neuroscience 1990, 34, 423-32. Gonon, F. Neuromethods 1995, 27, 153-77. Dugast, C.; Cespuglio, R.; Suaud-Chagny, M. F. J Neurochem 2002, 82, 529-37. Brock, J. A.; Tan, J. H. Br J Pharmacol 2004, 142, 267-74. Gonon, F.; Bao, J. X.; Msghina, M.; Suaud-Chagny, M. F.; Stjarne, L. J Neurochem 1993, 60, 1251-7. Gonon, F.; Msghina, M.; Stjaeme, L. Neuroscience 1993, 56, 535-8. Dunn, W. R.; Brock, J. A.; Hardy, T. A. Br J Pharmacol 1999, 128, 174-80. Gonon, F. Bioelectrochem Bioenerg 1995, 38, 247-9. Brock, J. A.; Bridgewater, M.; Cunnane, T. C. Br J Pharmacol 1997, 120, 769-76. Brock, J. A.; Dunn, W. R.; Boyd, N. S.; Wong, D. K. Br J Pharmacol 2000, 131, 1507-11. Clark, R. A.; Zerby, S. B.; Ewing, A. G. In Electroanalytical Chemistry; Israel, R., Ed.; Marcel Dekker, INC.: New York, 1996; Vol. 20. Wightman, R. M. Anal Chem 1981, 53, 1125A-6, 8A, 30A, 32A, 34A. 172 [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] [142] [143] [144] [145] Fleischmann, M.; Pons, S. J Electroanal Chem Interfacial Electrochem 1987, 222, 107-15. Lane, R. F .; Hubbard, A. T. Anal Chem 1976, 48, 1287-92. Lane, R. F.; Hubbard, A. T.; F ukunaga, K.; Blanchard, R. J. Brain Res 1976, 114, 346-52. McCreery, R. L. Electroanal Chem 1990, 1 7, 221-374. Capella, P.; Ghasemzadeh, B.; Mitchell, K.; Adams, R. N. Electroanalysis 1990, 2, 175-82. Fox, K.; Arrnstrong-James, M.; Millar, J. J Neurosci Methods 1980, 3, 37-48. Gonon, F.; Buda, M.; Cespuglio, R.; Jouvet, M.; Pujol, J. F. Brain Res 1981, 223, 69-80. Suaud-Chagny, M. F.; Cespuglio, R.; Rivot, J. P.; Buda, M.; Gonon, F. J Neurosci Methods 1993, 48, 241-50. Suaud-Chagny, M. F. Methods 2004, 33, 322-9. Yavich, L.; Jakala, P.; Tanila, H. J Neurochem 2005, 95, 641-50. Gonon, F .; Buda, M.; Pujol, J. F. IBRO Handbook Series 1984, 6, 153-71. Stamford, J. A. Brain Res 1985, 357, 119-35. Runnels, P. L.; Joseph, J. D.; Logrnan, M. J.; Wightman, R. M. Anal Chem 1999, 71, 2782-9. Venton, B. J .; Michael, D. J .; Wightman, R. M. J Neurochem 2003, 84, 373-81. Kawagoe, K. T.; Jankowski, J. A.; Wightman, R. M. Anal Chem 1991, 63, 1589- 94. Han, 8.; Pan, L. S.; Kania, D. R. Diamond: Electronic Properties and Applications; Kluwer Academic: Norwell, MA, 1995. Collins, A. T. J. Phys. .' Condense. Matter 2002, 14, 3743-50. Nishimura, K.; Das, K.; Glass, J. T. J Appl Phys 1991, 69, 3142-8. Nishitani-Gamo, M.; Yasu, E.; Xiao, C.; Kikuchi, Y.; Ushizawa, K.; Sakaguchi, 1.; Suzuki, T.; Ando, T. Diamond Relat Mater 2000, 9, 941-7. 173 [146] [147] [148] [149] [150] [151] [152] [153] [154] [155] [156] [157] [158] [159] [160] [161] [162] Xu, J.; Granger, M. C.; Chen, Q.; Strojek, J. W.; Lister, T. E.; Swain, G. M. Anal Chem 1997, 69, 59lA-7A. Stoffer, J.; Haymond, S.; Zak, J. K.; Show, Y.; Cvackova, Z.; Swain, G. M. Electrochem Soc Interface 2003, 12, 33-8. Fischer, A. E.; Show, Y.; Swain, G. M. Anal Chem 2004, 76, 2553-60. Manivannan, A.; Rarnakrishnan, L.; Seehra, M. S.; Granite, E.; Butler, J. B.; Tryk, D. A.; Fujishima, A. J Electroanal Chem 2005, 577, 287-93. Muna, G. W.; Quaiserova-Mocko, V.; Swain, G. M. Electroanalysis 2005, 17, 1160-70. Swain, G. M.; Anderson, A. B.; Angus, J. C. MRS Bulletin 1998, 23, 56-60. Swain, G. M. J Electrochem Soc 1994, 141, 3382-93. Granger, M. C.; Witek, M.; Xu, J .; Wang, J .; Hupert, M.; Hanks, A.; Koppang, M. D.; Butler, J. E.; Lucazeau, G.; Merrnoux, M.; Strojek, J. W.; Swain, G. M. Anal Chem 2000, 72, 3793-804. Swain, G. M. In Electroanalytical Chemistry; Bard, A. J ., Rubinstein, 1., Eds., 2004; Vol. 22, pp 181-277. Cvacka, J.; Quaiserova, V.; Park, J.; Show, Y.; Muck, A., Jr.; Swain, G. M. Anal Chem 2003, 75, 2678-87. Shin, D.; Sarada, B. V.; Tryk, D. A.; Fujishima, A.; Wang, J. Anal Chem 2003, 75, 530-4. Park, J .; Quaiserova-Mocko, V.; Peckova, K.; Galligan, J. J .; Fink, G. D.; Swain, G. M. Diamond Relat Mater 2006, 15, 761-72. Muna, G. W.; Quaiserova-Mocko, V.; Swain, G. M. Anal Chem 2005, 77, 6542-8. Park, J.; Show, Y.; Quaiserova, V.; Galligan, J. J.; Fink, G. D.; Swain, G. M. J Electroanal Chem 2005, 583, 56-68. Muna, G. W.; Tasheva, N.; Swain, G. M. Environ Sci and Technol 2004, 38, 3674-82. Dennison, J. R.; Holtz, M.; Swain, G. Spectroscopy 1996, 11, 38-46. Koppang, M. D.; Witek, M.; Blau, J.; Swain, G. M. Anal Chem 1999, 71, 1188- 95. 174 [163] [164] [165] [166] [167] [168] [169] [170] [171] [172] [173] [174] [175] [176] [177] [178] [179] [180] [181] Bennett, J. A.; Wang, J.; Show, Y.; Swain, G. M. J Electrochem Soc 2004, 151, E306-E13. Mermoux, M.; Marcus, B.; Swain, G. M.; Butler, J. E. J Phys Chem B 2002, 106, 10816-27. Chow, R. H.; von Ruden, L.; Neher, E. Nature 1992, 356, 60-3. Zhou, Z.; Misler, S. J Biol Chem 1996, 2 71, 270-7. Ranganathan, S.; Kuo, T.-C.; McCreery, R. L. Anal Chem 1999, 71, 3574-80. Chen, Q.; Swain, G. M. Langmuir 1998, 14, 7017-26. Libioulle, L.; Houbion, Y.; Gilles, J. M. Rev Sci Instrum 1995, 66, 97-100. Hu, 1. F.; Karweik, D. H.; Kuwana, T. J Electroanal Chem Interfacial Electrochem 1985, 188, 59. Rice, R. J.; McCreery, R. L. J Electroanal Chem Interfacial Electrochem 1991, 310, 127-38. Granger, M. C.; Swain, G. M. J Electrochem Soc 1999, 146, 4551-8. Swain, G. M. Electrically conducting diamond thin films: Advanced electrode materials for electrochemical technologies, 2004. Cline, K. K.; McDermott, M. T.; McCreery, R. L. J Phys Chem 1994, 98, 5314- 19. Chen, P.; McCreery, R. L. Anal Chem 1996, 68, 3958-65. Fink, G. D.; Johnson, R. J .; Galligan, J. J. Hypertension 2000, 35, 464-9. Capella, P.; Ghasemzadeh, M. B.; Adams, R. N.; Wiedemann, D. J.; Wightman, R. M. J Neurochem 1993, 60, 449-53. Stamford, J. A.; Justice, J. 3., Jr. Anal Chem 1996, 68, 359A-63A. Jolley, S.; Koppang, M.; Jackson, T.; Swain, G. M. Anal Chem 1997, 69, 4099- 107. Show, Y.; Witek, M.-l. A.; Sonthalia, P.; Swain, G. M. Chem Mater 2003, 15, 879-88. Kissinger, P. T.; Hart, J. B.; Adams, R. N. Brain Res 1973, 55, 209-13. 175 [182] [183] [184] [185] [186] [187] [188] [189] [190] [191] [192] [193] [194] [195] [196] [197] [198] [199] [200] [201] DuVall, S. H.; McCreery, R. L. J Am Chem Soc 2000, 122, 6759-64. Cahill, P. S.; Walker, O. D.; Finnegan, J. M.; Mickelson, G. E.; Travis, E. R.; Wightman, R. M. Anal Chem 1996, 68, 3180-6. Pihel, K.; Walker, Q. D.; Wightman, R. M. Anal Chem 1996, 68, 2084-9. Wightman, R. M.; May, L. J .; Michael, A. C. Anal Chem 1988, 60, 769A-79A. Msghina, M.; Gonon, F.; Stjarne, L. J Physiol 1999, 515 (Pt 2), 523-31. Dunn, W. R.; Hardy, T. A.; Brock, J. A. Br J Pharmacol 2003, 140, 231-8. Nilsson, H.; Ljung, B.; Sjoblom, N.; Wallin, B. G. Acta Physiol Scand 1985, 123, 303-9. Nilsson, H.; Goldstein, M.; Nilsson, O.; Ljung, B.; Sjoblom, N.; Wallin, B. G. Acta Physiol Scand 1986, 126, 121-33. Cahill, P. S.; Wightman, R. M. Anal Chem 1995, 67, 2599-605. Daugirdas, J. T. Am J Kidney Dis 2001, 38, 81 1-7. Kreulen, D. L. J Smooth Muscle Res 2003, 39, 269-79. Hottenstein, O. D.; Kreulen, D. L. J Physiol 1987, 384, 153-67. Hirst, G. D.; Jobling, P. Br J Pharmacol 1989, 96, 993-9. Segal, S. S. Acta Physio] Scand 2000, I68, 51 1-8. Tani, H.; Singer, W.; McPhee, B. R.; Opfer-Gehrking, T. L.; Haruma, K.; Kajiyama, G.; Low, P. A. Auton Neurosci 2000, 86, 107-13. Bjorklund, A.; Lindvall, 0.; Svensson, L. A. Histochemie 1972, 32, 113-31. Klemm, M. F.; Van Helden, D. F.; Luff, S. E. J Comp Neurol 1993, 334, 159-67. Browning, K. N.; Zheng, Z.; Kreulen, D. L.; Travagli, R. A. Am J Physiol 1999, 276, H1263-H72. Stjaeme, L.; Bao, J. X.; Gonon, F.; Msghina, M. Neuroscience 1994, 60, 1021-38. Galligan, J. J.; Hess, M. C.; Miller, S. B.; Fink, G. D. J Pharmacol Exp Ther 2001, 296, 478-85. 176 [202] [203] [204] [205] [206] [207] [208] [209] [210] [211] [212] [213] [214] [215] [216] [217] [218] Reid, J. L.; Zivin, J. A.; Kopin, I. J. Circ Res 1975, 37, 569-79. de Champlain, J. Clin Endocrinol Metab 1977, 6, 633-55. Bouvier, M.; De Champlain, J. Clin Exp Hypertens, 1985, A 7, 1629-45. Bouvier, M.; de Champlain, J. J A uto Nerv Syst 1986, 15, 191-5. Esler, M.; Lambert, G.; Jennings, G. J Hypertens Suppl 1990, 8, 853-7. de Champlain, J.; Farley, L.; Cousineau, D.; van Ameringen, M. R. Circ Res 1976, 38, 109-14. Micalizzi, E. R.; Pals, D. T. Life Sci 1979, 24, 2071-6. Sudhir, K.; Angus, J. A.; Esler, M. D.; Jennings, G. L.; Lambert, G. W.; Komer, P. I. J Hypertens 1990, 8, 1119-28. Monos, E.; Berczi, V.; Nadasy, G. Physiol Rev 1995, 75, 611-66. Langer, S. Z. J Physiol 1970, 208, 515-46. Vanhoutte, P. M.; Webb, R. C.; Collis, M. G. Clin Sci 1980, 59 Suppl 6, 2115-233. Polonia, J. J .; Guerreiro, M.; Guimaraes, 8.; Garrett, J. Arch Int Pharmacodyn T her 1986, 279, 5-16. Daly, C. J .; McGrath, J. C.; Wilson, V. G. BrJ Pharmacol 1988, 95, 485-500. Civantos Calzada, B.; Aleixandre de Artinano, A. Pharmacol Res 2001, 44, 195- 208. Perez-Rivera, A. A.; Fink, G. D.; Galligan, J. J. J Pharmacol Exp Ther 2004, 308, 350-7. Meggs, L. G.; Stitzel, R.; Ben-Ari, J .; Chander, P.; Gammon, D.; Goodman, A. 1.; Head, R. Clin Exp Hypertens A 1988, 10, 229-47. Nyhof, R. A.; Laine, G. A.; Meininger, G. A.; Granger, H. J. Fed Proc 1983, 42, 1690-3. 177 ll]ll]1j[1]l|]jlj][1111111111