<<

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer and continuing from left to right in equal sections with small overlaps. Each original is also photographed in one exposure and is included in reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6" x 9" black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order.

A Bell & Howell Information Company 300 North Zeeb Road. Ann Arbor. Ml 48106-1346USA 313/761-4700 800/521-0600

SIGMA RECEPTORS AND THE IMMUNE SYSTEM

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Graduate School of the Ohio State University

By

Yuhong Liu, B.M. (Bachelor of Medicine)

The Ohio State University 1995

Dissertation Committee: Approved by

Dr. S. A. Wolfe, Jr. Dr. M. S. O’Dorisio Research Advisor Dr. R. H. Fertel Dr. W. P. Lafuse ft)/I Dr. J. F. Sheridan Aavisorof Record Department of Medical Microbiology and Immunology UMI Number: 9612230

UMI Microform 9612230 Copyright 1996/ by OMI Company. All rights reserved.

This microform edition is protected against unauthorized copying under Title 17, United States Code.

UMI 300 North Zeeb Road Ann Arbor, MI 48103 To My Parents,

My Husband Willie and My Daughter Ellen,

For Your Unfailing Love, Support, and Understanding.

ii ACKNOWLEDGMENTS

I would like to express my sincere appreciation to my

research advisor, Dr. Seth Wolfe, for his guidance and insight

throughout this project, his unshakable belief in me and this

research endeavor, and his and Dr. Charity Fox’s support and

friendship. I offer my sincere thanks to the other members of my

advisory committee, Dr. Richard Fertel, Dr. William Lafuse, Dr. Sue

O’Dorisio, and Dr. John Sheridan, for their invaluable suggestions

and comments. Special thanks go to Dr. O’Dorisio for her support

and encouragement when they were most needed, and to Dr. Fertel

for his professional insight and intellectual support. Also, I would

like to express gratitude to Joseph Pultz and Dr. Dennis Pearl for

their assistance in statistical analysis of our data. Finally, I would

like to express my deepest appreciation to my husband and

daughter, my parents, my brother, my father- and mother-in-law, for their support and understanding throughout this experience. VITA

February 20, 1964 ...... Born, Changchun, P.R. China

1987 ...... Bachelor of Medicine, Basic Medicine, Beijing Medical University, Beijing, P.R. China

1987-1989 ...... Teaching Assistant, Department of Immunology, Beijing Medical University, Beijing, P.R. China

1989-1995 ...... Graduate Research Associate Department of Medical Microbiology and Immunology, The Ohio State University, College of Medicine, Columbus, Ohio

PUBLICATIONS

Liu, Y. and Y. Qian. 1989. The experimental studies of IL-3. Progress in Physiological Sciences 20(1 ):17

Liu, Y., Q. Gong and T. Tong. 1989. Isolation and characterization of high molecular weight DNA binding protein from serum of mice with H22 ascites hepatoma. J. Beijing Med. Univ. 21(5):372

Liu, Y., Q. Gong and T. Tong. 1990. Isolation and characterization of high molecular weight DNA binding protein from ascites of mice with Ehrlich ascites carcinoma. Chinese Biochem. J. 6(3):264

Wang, D., S. Yang, Y. Qian and Y. Liu. 1990. Establishment and characterization of an acute myelogenous leukemia cell line. J. Beijing Med. Univ. 22(5):331

iv Liu, Y., B. B. Whitlock, J. A. Pultz and S. A. Wolfe, Jr. 1995. Sigma-1 receptors modulate functional activity of rat splenocytes. J. Neuroimmunology 59: 143-154

Liu, Y. and S. A. Wolfe, Jr. 1995. and potentiate murine splenic B cell proliferation. Imuunopharmacology submitted

FIELDS OF STUDY

Major Field: Medical Microbiology and Immunology Studies in sigma receptors in immune system with Dr. Seth A. Wolfe, Jr. TABLE OF CONTENTS

DEDICATION...... ii

ACKNOWLEDGMENTS ...... iii

VITA ...... iv

LIST OF TABLES ...... vii

LIST OF FIGURES ...... viii

CHAPTER PAGE

I INTRODUCTION ...... 1

II MATERIALS AND METHODS ...... 13

III RESULTS ...... 23

IV DISCUSSION ...... 75

LIST OF REFERENCES ...... 94 LIST OF TABLES

TABLE PAGE

1. Pharmacological specificity of [3H]haloperidol- labeled sigma-1 receptors in rat spleen ( K j ) and potency of drugs suppressing ConA-induced proliferation of rat splenocytes (EC 50) ...... 29

2. Splenic [3H]haloperidol binding sites are present on isolated splenocytes ...... 39

3. Drugs that do not enhance anti-p induced murine splenic B cell proliferation ...... 74

vii LIST OF FIGURES

FIGURE PAGE

1. ConA induced rat splenocyte proliferation dose response cuive ...... 25

2. Three classical sigma agonists suppress ConA-induced ratsplenocyte proliferation ...... 27

3. (-)-Sulpiride has no effect on ConA-induced ratsplenocyte proliferation ...... 31

4. Specific binding sites for pHJhaloperidol are piesenton membranesofratspleen(A)andratsplenocytes(B) ...... 3 3 , 34

5. Isolated ratsplenocytes (A) had higher densities (Bma*) of pH]haloperidol binding sites than the rat spleens (B)ftom which they were obtained ...... 3 7 , 38

6. Competition binding assaysofthree prototype sigma ligands against pH]haloperidol in ratspleen ...... 41

7. The ability of seven reference sigma agonists, to suppress Con A-induced proliferation correlated highly with theirpotency in binding pHJhaloperidol-labeled splenic sigma-1 receptors ...... 46

8. Correlation between EC 50 and Kjofall fifteen drugstested ...... 48

9. Speafic binding sites for pH]haloperidol are piesenton membranes of mouse B lymphoma cell line, A20 cells ...... 50

10. Rosenthal plotsofthe same data as in Rgure 9 ...... 52

11. Haloperidol and spiperone enhance anti-ju. induced B cell proliferation in a dose-dependent manner ...... 55

v iii LIST OF FIGURES

FIGURE PAGE

12. Haloperidol and spiperone lowerthe threshold of anti-jj. needed to trigger Boell proliferation ...... 5 8 , 59

13. inhibits anti-p. induced Bcell proliferation ...... 62

14. Norepinephrineinhibitsanti-pinduced Bcellproliferation ...... 6 4

15. enhancesanti-pinduoed Bcell proliferation at 104 M ...... 66

16. Spiperone and dopamine antagonize each other’s effects on anti-p induced Bcell proliferation ...... 6 8 , 69

17. Spiperone and antagonize each other’s effects on anti-p induced B cell proliferation ...... 71,72

ix CHAPTER I INTRODUCTION

The Nature of Sigma Receptors

The existence of sigma receptors was first postulated by

Martin et al. in 1976 as a result of canine physiological studies with racemic mixtures of the opiate benzomorphan, N-allynormetazocine

(SKF 10,047, or NANM) (1). SKF 10,047 and related benzomorphans cause canine delirium and produce psychotomimetic effects in humans (2). It was originally thought that a novel type of ‘opiate’ receptor, sigma (a) opiate receptors, mediated the phenomena. However, further studies revealed that naltrexone, a classical antagonist, could not block the effects of SKF 1 0,047 in dogs (3). Therefore, it was clear that the sigma ‘opiate’ receptor should not be classified as a type of opiate receptor after all. But the designation was retained.

[3H]SKF 10,047 labeled etorphine-inaccessible sigma sites were subsequently described by Su in guinea pig brain (4,5).

Sigma receptors bind psychotomimetic drugs. For example:

1 2

1. Opiate-related compounds. The opiate benzomorphans

N-allylnormetazocine (SKF 10,047), ,

ethylketocyclazocine and , bind to both opiate receptors

and sigma receptors. But these two kinds of receptors have

opposite stereoselectivity. Opiate receptors bind to

(-)-benzomorphans better, whereas sigma receptors prefer

(+)-benzomorphans;

2. -related compounds. In some early studies

sigma receptors were believed to be identical to PCP receptors

based on the displacement of [3H]PCP binding by the sigma ligand

(+)-SKF 1 0,047 (6, 7). Later it was found that the drug selectivity

and the anatomical distribution of sigma receptors and PCP

receptors are distinct from each other. It is now known that PCP

and its derivative, TCP, bind to PCP receptors (in the ion channels

of NMDA receptors) with high affinities, while their binding affinity to sigma receptors is relatively low (8);

3. Guanidines. 1,3-Di(2-tolyI)guanidine (DTG) binds to sigma

receptors selectively and with high affinity. Ligands for other

receptors are very weak in displacing [3H]DTG binding (9);

4. 3-Phenylpiperidines.

3-(3-hydroxyphenyl)-N-propylpiperidine (3-PPP) binds to both sigma receptors (10) and dopamine D 2 receptors (11). But (+)-3-PPP binds to sigma receptors with higher affinity, (-)-3-PPP is more potent at dopamine D 2 receptors (11);

5. Miscellaneous compounds. Haloperidol binds to sigma receptors, serotonin 5-HT2 and dopamine D2 receptors with high 3

affinity (12). (-)- binds to sigma receptors with high

affinity, while its (+)-enantiomer binds to D 2 receptors and

serotonin receptors with high affinity (13). A novel

drug, rimcazole, also binds to sigma receptors, but with relatively

low affinity (14).

Recently, it has been found that there is more than one type

of sigma receptor. A second type of sigma receptor was first found

in rat PC 1 2 cells, a tumor cell line that has the phenotype of

sympathetic neurons (15). This new sigma site has an overlapping

pharmacology with the previously described ( 0 1 ) sigma receptors.

It has substantially lower affinity for (+)-benzomorphans, and its

affinity for (-)-benzomorphans is greater than for

(+)-benzomorphans. Researchers have come to agreement on the

nomenclature and labeling conditions for the two subtypes of sigma

receptors, which are now known as sigma-1 ( 0 1 ) and sigma-2 ( 0 2 )

(16). These can coexist in tissues. For example, rat liver contains approximately 25% sigma-1 sites and 75% sigma-2 sites (17); rat brain also contains both sites (18). Furthermore, studies done in our laboratory indicated that there might be a third type of saturable and specific binding site for sigma ligands in rat spleen and human peripheral blood leukocytes (Whitlock and Wolfe, unpublished data).

Little is known about the structure of sigma receptors. Sigma binding sites are sensitive to temperature, pH, and protein-modifying agents, which suggests that sigma receptors are proteins (5, 19). Several attempts have been made to determine the 4 molecular weight of sigma receptors (15,20-22). A single polypeptide of 29 KD was found in guinea pig brain using azido-DTG to label sigma-1 sites, while two polypeptides with molecular weights of 18 KD and 21 KD were discovered in PC1 2 cells. Because no or very few sigma-1 receptors are found on

PC12 cells, these two polypeptides may represent sigma-2 receptors. In addition, McCann and Su identified a

CHAPS-solubilized sigma-1 complex with a molecular weight of

450 KD from rat liver membrane (23). Another group has isolated and sequenced a 28 KD, cyclophilin-like protein from rat liver membrane that may be a component of sigma receptors (24). These discrepancies regarding the molecular weight of sigma-1 receptors might be due to species differences or the difference between a single subunit and a whole receptor complex. So far, sigma receptors have not been cloned.

There have been several studies on the signal transduction mechanisms of sigma receptors. One group reported that sigma receptors might be coupled to guanine nucleotide-binding proteins

(25, 26). As observed with other G protein-coupled receptors, guanosine triphosphate and Gpp(NH)p, a non-hydrolyzable analog of GTP, inhibited the binding of [3H](+)-3-PPP to rat brain membranes. Pertussis toxin and cholera toxin also decreased binding by [3H](+)-3-PPP (26,27), which indicated that sigma receptors may interact both with Gj and G0, as well as Gs proteins.

Further studies also showed that in rat brain only [3H](+)-3-PPP binding sites were sensitive to guanine nucleotides, while [3H]DTG 5

binding sites were not (28). This suggests that subtypes of sigma

receptors may have different signal transduction mechanisms.

Also, Bowen and coworkers found that although sigma ligands did

not have direct effects on phosphoinositide (PI) turnover in rat

brain, they did inhibit inositol phosphate (IP 1 ) production following

the stimulation of muscarinic receptors by either carbachol or

oxotremorine-M in synaptoneurosomes from whole rat brain (29, 30).

This implies that, in brain, sigma receptors can modulate the

phosphoinositide secondary messenger system.

Physiological Functions of Sigma Receptors

Evidence for the existence of endogenous sigma ligands has

been observed in several laboratories. Preliminary reports of

partially purified extracts from brain and liver have shown both

peptides and non-peptides with the binding characteristics of sigma

ligands (31-33). Many known endogenous substances have been

screened and proven to be ineffective in inhibiting binding at sigma

receptors (reviewed in 34). But there are a few exceptions. Roman

et al. reported that the peptides, (NPY) and peptide

YY (PYY), have high affinity for rat brain sigma receptors labeled with [3H](+)-SKF 10,047 (35), although others have reported that they have no such activity (23, 36). , another candidate for sigma endogenous ligand, was reported to inhibit [3H](+)-SKF

10,047 binding in guinea pig brain and spleen, with a K j of 268nM

(37). Also, Zn++ has been proposed to be an endogenous ligand for sigma-2 receptors in rat brain (38). However, none of these has been conclusive, and the true identity of the sigma endogenous

ligand(s) remains unknown.

Sigma receptors are widely distributed in the central nervous

system (39), endocrine tissues (40-42), gastrointestinal tract (43),

liver (44) and immune system (45, 46). However, the physiological

functions of sigma receptors are poorly understood, mainly due to

the lack of selective ligands for sigma receptors. For example,

conclusions about the functions of sigma receptors based on the

actions of SKF 10,047 must be viewed with caution, because this

compound acts at several different sites, i.e. PCP receptors, opiate

receptors and sigma receptors. Although much effort has recently

been devoted to developing more selective compounds (47-49, and

numerous articles by De Costa and colleagues, reviewed in 50),

most of them have not been well characterized enough,

pharmacologically and physiologically, to be recognized as

selective ligands for sigma receptors.

Despite these difficulties, sigma receptors have been linked to phenomena in the central nervous system and in peripheral tissues. (+)-Pentazocine has been shown to increase local cerebral glucose utilization in brain areas that are rich in sigma receptors (51). Steinfels et al., found that (+)-pentazocine produced a sigma-selective discriminative stimulus cue in rats, enabling them to discriminate non-sigma drugs from the training drug ,

(+)-pentazocine (52). Sigma receptors have also been implicated in the etiology of schizophrenia. Both autoradiographic and homogenate radioligand binding studies showed dramatic reductions of sigma receptors in several brain regions from autopsy

samples of schizophrenic patients (53, 54). Other studies have

provided evidence that sigma receptors may be also involved in

certain movement disorders such as dystonia (55). In the periphery,

Campbell and colleagues showed that sigma ligands blocked

electrically or 5-HT-induced contractions of the isolated guinea pig

ileum/myenteric plexus preparation (56). Also, electrically induced

twitches of isolated guinea-pig vas deferens were potentiated by

sigma agonists (42). Sigma agonists blocked a tonic, outward

potassium current in NCB-20 neuroblastoma cells (57). In an in

vitro assay system using bovine adrenal chromaffin cells, sigma

ligands were shown to noncompetitively inhibit nicotine-stimulated

catecholamine release (58).

Role of Sigma Receptors in the Immune System

Sigma receptors may be of importance for the immune system

as well. Fudenberg, Whitten and Khansari found that high

concentrations of PCP (10'5M) could inhibit functional activities of

human peripheral blood leukocytes (HPBL) in vitro, and they

detected specific binding sites for [3H]PCP on HPBL (59, 60). In

1987, Dornand and colleagues reported that PCP and PCP analogs

suppressed mitogen-induced proliferation of murine splenocytes

(61). They found that, at high doses (10‘6M to 10"5M), PCP analogs caused a reduction in the resting potential of splenocytes, and

reduced the ability of these cells to depolarize and elevate their intracellular calcium levels in response to stimulation with 8

concanavalin A (ConA). However, in previous studies, Wolfe and

associates could not detect high affinity PCP receptors (PCP-i) in

rat spleen or HPBL, while they did find strikingly high

concentrations of sigma receptors in both of these tissues (45, 46).

This suggests that the effect of PCP on immune cells may be

mediated through sigma receptors. Recently, Carr et al., reported

that the sigma agonists 1 ,3-di(2-tolyl)guanidine (DTG), haloperidol,

and (+)-pentazocine suppressed ConA-induced proliferation of

murine splenocytes. In contrast, bacterial lipopolysaccharide

(LPS)-induced proliferation was enhanced by these drugs (62).

They also reported that pokeweed mitogen (PWM)-induced and

LPS-induced immunoglobulin production was either enhanced or

suppressed by sigma agonists, depending on the mitogen and

concentration of the drug. In addition, they have shown that sigma

agonists may also modulate murine splenic natural killer activity

and forskolin-induced cAMP production in splenocytes (63). More

recently, Garza et al., found that high concentrations (10_6M to

10_5M) of haloperidol and the novel sigma ligand

(+)-azidophenazocine (64) suppressed ConA-induced interferon

production. However, in their hands, the sigma agonists DTG,

(+)-pentazocine, and (+)-1-propyl-3-(3-hydroxyphenyl)piperidine

((+)-3-PPP) did not significantly affect interferon production (65),

which suggests that this may not have been a sigma receptor-

mediated effect. More recently, drug binding and modulation of

several in vivo and in vitro functional assays have been reported for a novel sigma ligand, SR 31747 (66, 67). SR 31747 inhibited 9

mitogen induced lymphocyte proliferation, and prevented both

graft-versus-host disease and delayed-type hypersensitivity granuloma formation in vivo. The same group also reported that

SR 31747 given in vivo caused an indirect inhibition of

lipopolysaccharide-induced production of interleukin (IL)-1, IL-6

and tumor necrosis factor-a, due to an elevation of the level of

corticosteroids (68).

Collectively, these data suggest that sigma receptors are

present on lymphocytes, and that these receptors may regulate both T cell and B cell functions. However, there are several major problems with the studies mentioned above:

1. In lymphocyte proliferation assays, mitogens that could activate multiple types of cells were used on mixed cell populations, such as lipopolysacharride (LPS) which could act on both B cells and macrophages, and pokeweed mitogen (PWM) which could act on both B cells and T cells. This makes it hard to interpret which subpopulation of cells were responding to sigma agonists.

2. Sigma agonists do not act exclusively at sigma receptors, but can also be quite potent at a multitude of other sites, including ion channels, dopamine receptors, serotonin receptors, opiate receptors, phencyclidine receptors, ai adrenergic receptors, and uptake sites for serotonin and dopamine (reviewed in 69). Many of these sites have been identified on immune cells, among them adrenergic receptors (70-72), dopamine receptors (73-81), serotonin receptors (82, 83), and opioid receptors (reviewed in 84). Therefore, 10

immune modulation by putative sigma agonists might actually be

the result of drug actions at a multitude of receptors on immune

cells. In most of the existing studies, attempts to discriminate

between the effects through sigma receptors and other receptors

were not made. In addition, the subtypes of sigma receptors,

sigma-1 and sigma-2, which have overlapping pharmacology, were

not distinguished from one another.

3. SR 31747, a recently described putative sigma ligand, has

a very different pharmacological profile from either sigma-1 or

sigma-2 (67). The authors argued that it might bind to an allosteric

modulation site of sigma receptors. Although SR 31747 has been

shown to have immunomodulatory effects (66, 68), many questions

still remain as to how this is related to sigma receptors.

In order to demonstrate that sigma receptors modulate a

biological function, the receptors must be identified, quantitated,

and characterized kinetically and pharmacologically in the cells

and tissues of interest. Sigma agonists must modulate functional

activity in those cells or tissues, and the ability of drugs to

modulate function must correlate with drug binding potency at the

receptor. (It is to be expected that some compounds may fall

outside the correlation due to additional actions at other receptors.)

This has been accomplished in other systems in which sigma

receptors have been reported to act (55, 56, 58, 85), but these criteria

have not yet been met in the immune system. In the immune

system, several laboratories have demonstrated drug binding sites and/or have shown modulation of functional activities by sigma 11

agonists (45, 46, 59-63, 65-67), but a mathematical correlation

between drug binding at the receptor and the pharmacology of the

biological responses has not been demonstrated.

Hypothesis and Significance

A significant recent development in immunology is a growing

appreciation that the immune system functions as part of a

complex, interactive network consisting of the immune, endocrine,

and nervous systems. Together, these systems deal with external

and internal stimuli, and maintain homeostasis under dynamic

conditions. There is a considerable body of evidence that the

immune system is modulated by endocrine and nervous inputs, and that it in turn causes changes in endocrine status and central

nervous system activity (reviewed in 86). The mechanisms of this communication, -neurotransmitters, hormones, “immune" cytokines and their receptors, - are shared among these three systems (87).

Although characterization of endogenous sigma ligand(s) is still in preliminary stages, the underlying premise of our investigations is that the endogenous sigma ligand(s) is one of these mediators. The presence of sigma receptors in high density in the central nervous system, the endocrine system and the immune system suggests that endogenous sigma ligand(s) may have a special role in coordinating the responses of these three systems to environmental stimuli.

The present study was undertaken to determine whether sigma agonists have direct effects on T cell and B cell function. 12

The objective was to identify sigma receptors on immune cells pharmacologically, using radioligand binding assays. As mentioned before, due to the cross reactivity of sigma ligands, a series of sigma ligands were employed to characterize sigma receptors on immune cells. The same group of drugs were also tested in in vitro assays, for their effects on a commonly used measure of T and B cell responsiveness: mitogen-induced proliferation. If sigma agonists do influence T or B cell proliferation, we would try to correlate their ability to modulate function with their binding potency at sigma receptors. We would be able to conclude whether sigma receptors could mediate immunomodulatory effects, based on the outcome of the correlation. We believe this study is the first definitive study on whether sigma receptors in the immune system have functional significance. CHAPTER II MATERIALS AND METHODS

Animals

Male Sprague-Dawley rats (Harlan Sprague-Dawley Inc.,

Indianapolis, IN), 150 - 300 gram, were used for radioligand binding assays and ConA-induced splenocyte proliferation assays.

Animals were housed three to four per cage.

Male C57BL/6J mice (Jackson Laboratories, Bar Harbor, ME),

8-12 weeks old, were used in splenic B cell proliferation assays.

Animals were housed ten per cage.

All animals had access to food and water ad libitum, and were maintained on a 12 hour light/dark cycle at a room temperature of 22°C. All animals were treated in accordance with

National Institute of Health animal care guidelines.

Cell Cultures

The mouse B lymphoma cell line, A20, was obtained from

American Type Culture Collection (Rockville, MD). Cells were

13 14

maintained in suspension in RPMI 1640 medium (GIBCO BRL Life

Technologies, Inc., Gaithersburg, MD) containing 10% fetal bovine

serum (FBS, defined grade, HyClone Laboratories, Inc., Logan, UT), 50 |a.g/ml gentamycin (GIBCO), and 5 x 10_5M

2-mercaptoethanol (BIO-RAD Laboratories, Richmond, CA). Cells

were seeded at 2 x 105 viable cells/ml in 150 cm2 tissue culture

flasks (Corning Inc., Corning, NY), and harvested and passed every

three days. Harvested cells were washed once in Ca2+/Mg2+-free

Hank’s Balanced Salt Solution (HBSS, GIBCO), and counted. Cell

pellets were stored at -80°C until use in binding assays.

Drugs and Reagents

Haloperidol, haloperidol metabolite I (4-(4-chlorophenyl)-4-

hydroxypiperidine), haloperidol metabolite II ((±)-4-(4-chlorophenyl)-a-(4-fluorophenyl)-4-hydroxy-1 - piperidinebutanol), (+) and (-)-pentazocine, (+) and (-)-butaclamol,

(+) and (-)-3-PPP hydrochloride (3-(3-hydroxyphenyl)-N- propylpiperidine hydrocloride), (+) and (-)-SKF 10,047

(N-allynormetazocine hydrochloride), DTG (1,3-di(2-tolyI) guanidine), rimcazole dihydrochloride, TCP hydrochloride

( hydrochloride, 1 -[1 -(2-thienyl)cyclohexyl]piperidine hydrochloride), PCP hydrochloride (phencyclidine hydrochloride,

1-(1-phenylcyclohexyl)piperidine hydrochloride), spiperone hydrochloride, (-)-sulpiride, pindobind-5HT1 A, (+)-SCH-23390 hydrochloride, , LY-53,857 maleate, WB-4101 hydrochloride, 5-methyl-, dopamine hydrochloride, 15

(-)-norepinephrine bitartrate, and serotonin hydrochloride were

purchased from Research Biochemicals, Inc. (Natick, MA).

Phentolamine hydrochloride, , and

concanavalin A (ConA) were purchased from Sigma Chemical Co.

(S t. Louis, MO). Affinity-purified F(ab ’)2 fragment goat anti-mouse IgM (anti-p) antibodies were purchased from Jackson

ImmunoResearch Laboratories, Inc. (West Grove, PA).

Preparation of Rat Spienocytes

Spleens from C02-ki I led Sprague-Dawley rats were removed

and teased into single-cell suspensions by pressing them through a

stainless steel mesh. Spienocytes were suspended in Ca2+ and

Mg2+ free HBSS (GIBCO), layered over 70% Percoll (Pharmacia,

LKB Biotechnology, Uppsala, Sweden), and centrifuged at 1000 x g

for 15 min., at 4°C. Spienocytes were collected from the

HBSS-Percoll interface, washed twice by centrifugation in HBSS,

and suspended in RPMI 1640 medium (BioWhittaker, Inc.,

Walkersville, MD) containing 10% fetal bovine serum (FBS, defined grade, HyClone), 50 pg/ml gentamycin (GIBCO), and 5 x 10‘5 M

2-mercaptoethanol (BIO-RAD).

B Cell Preparation

Spleens from C02-ki I led C57BL/6J mice were removed and teased into single-cell suspensions by pressing them through a

stainless steel mesh. Spienocytes were suspended in

Ca2+/Mg2+-free HBSS (GIBCO), layered over a discontinuous 16

55%/70% Percoll gradient, and centrifuged at 1000 x g for 15 min.

at 4°C. Cells were collected from the 55% /7 0% Percoll interface,

washed twice by centrifugation in HBSS, and suspended at 2 x 107

cells/ml in Dulbecco’s phosphate-buffered saline (D-PBS, GIBCO)

containing 0.1% bovine serum albumin (BSA, Sigma). The cells were incubated for 30 min. at 4°C with 20 pg/ml biotinylated anti-

Thy1.2 (CD90) and biotinylated anti-Mac-1, ccm chain (CD11 b)

monoclonal antibodies (PharMingen, San Diego, CA), and washed

twice by centrifugation through D-PBS/0.1 % BSA. They were then

incubated with streptavidin-conjugated magnetic beads

(Dynabeads M-280 streptavidin, Dynal, Inc., Lake Success, NY)

(1 x 1 07 cells in 0.5 ml, at a celhbead ratio of 1:5). After incubation

at 4°C for 15 min., 3 ml of D-PBS/0.1 % BSA was added to the

mixture of cells and beads, and bead-coupled Thy1.2+ and Mac-1 +

cells were removed by placing the tubes on a magnetic particle

concentrator (Dynal) for 3 min. Supernatant medium and

non-labeled cells were decanted, and the magnetically entrapped cells and beads were resuspended three times with 3 ml of

D-PBS/0.1% BSA, and re-separated using the magnet.

Supernatants were pooled and separated again using the magnet.

The composition of the negatively-selected B cell populations was then assessed by flow cytometry. The enriched B cell populations were found to be contaminated less than 5% by Thy1.2+ and

Mac-1 + cells. 17

ConA-induced Proliferation of Rat Spienocytes

Rat spienocytes (2.5 x 1 05 cells/well) were added to

flat-bottom 96-well microtiter plates (Becton Dickinson, Lincoln

Park, NJ) and incubated at 37°C for two hours with varying

concentrations of drugs, after which ConA (final concentration, 0.5

|ig/ml) was added to all wells. The cells were then incubated at

37°C in 5% C02/95% air for 48 hours, and [3H]thymidine (0.5

pCi/well, Amersham Corp., Arlington Heights, IL) was added during the last 4-8 hours of culture. Cultures were harvested on Whatman

GF/B glass fiber filters in a cell harvester (Biomedical Research and Development Laboratories, Inc., Gaithersburg, MD), washed with distilled water. Filter discs were collected in 6 ml liquid scintillation vials (Wheaton Scientific, Millville, NJ), and 3 ml liquid scintillation cocktail (Formula-989, Biotechnology systems, NEN

Research Products, Boston, MA) was added to each vial. Liquid scintillation vials were then counted in a scintillation counter

(Beckman LS6000IC) to measure [3H]thymidine incorporation as an index of cell proliferation. Drugs used in these assays were: haloperidol, haloperidol metabolite I, haloperidol metabolite II,

(+) and (-)-pentazocine, (+) and (-)-butaclamol, (+) and (-)-3-PPP,

(+) and (-)-SKF 10,047, DTG, rimcazole, TCP, and PCP.

B cell proliferation assay

Enriched B cells were suspended in RPMI 1640 medium

(GIBCO) containing 10% fetal bovine serum (FBS, defined grade, HyClone), 50 pg/ml gentamycin (GIBCO), and 5 x 10~5 M 18

2-mercaptoethanol (BIO-RAD). B cells (1 x 105/well) were

incubated in flat-bottom 96-well microtiter plates (Corning Glass

Works, Corning, New York) at 37°C for 1 hour with varying concentrations of drugs, after which anti-p antibody (final concentration, 0.5 pg/ml) was added to all wells. The cells were then incubated at 37°C in 5% C02/95% air for 48 hours, after which [3H]thymidine (0.5 (iCi/well, Amersham) was added, and cells were incubated for another 16-18 hours. Using a cell harvester

(Biomedical Research and Development Laboratories), cultures were harvested and washed with distilled water by filtration on glass fiber filters (Whatman GF/B, Biomedical Research and

Development Laboratories). Filter discs were collected in 6 ml liquid scintillation vials (Wheaton Scientific, Millville, NJ), and 3 ml liquid scintillation cocktail (Formula-989, Biotechnology systems,

NEN Research Products, Boston, MA) was added to each vial.

Liquid scintillation vials were then counted in a scintillation counter

(Beckman LS6000IC) to measure [3H]thymidine incorporation as an index of cell proliferation. Drugs used in these assays were: haloperidol, spiperone, DTG, (+) and (-)-pentazocine, (+) and

(-)-butaclamol, (+)-SCH-23390, (-)-sulpiride, ritanserin, LY-53,857, pindobind-5HT1 A, WB-4101, 5-methyl-urapidil, , methoxamine, clonidine, serotonin, dopamine, and

(-)-norepinephrine. 19

Preparation of Rat Spleen, Rat Spienocytes, and A20 cell membranes

Rats were killed with CO2 and their spleens were dissected.

For radioligand binding assays in which whole spleens were used,

spleens were frozen rapidly on dry ice, then stored at -80°C. For

assay, frozen spleens were disrupted with a Polytron tissue

homogenizer (Brinkman Instruments, Westbury, NY) in 25-50

volumes of ice-cold 50 mM Tris-HCI buffer (pH 7.7 at 4°C).

Membranes were pelleted by centrifugation at 40,000 x g for 10

min. at 4°C, and washed twice by resuspension in the same buffer

and recentrifugation. After the second wash they were

resuspended in 50 mM Tris-HCI buffer (pH 7.7 at 22°C) and kept on

ice until placed in binding assays. To prepare washed splenocyte

membranes for binding assays, cells were teased from freshly dissected spleens and isolated on Percoll gradients as described for ConA-induced proliferation assays, above. Isolated spienocytes were then frozen on dry ice and stored at -80°C. For assay, frozen spienocytes and A20 cells (prepared as described previously), were thawed, subjected to homogenization, centrifuged, washed and resuspended in a manner identical to the procedure described for whole spleen. Protein content of the membrane suspensions was determined using commercial protein assay kit (Sigma, Cat.

No. P5656). 20

Characterization of [3H]Haloperidol Binding to Sigma-1 Sites in Rat Spleen, Rat Spienocytes, and A20 Cells

For saturation binding assays (16, 45, 88), membrane

preparations equivalent to 5.0 mg wet weight of rat spleen/tube,

2 x 107 rat splenocytes/tube, or 1 x 107 A20 cells/tube, were

incubated at room temperature in total volume of 2.5 ml 50 mM

Tris-HCI buffer (pH 7.7), with increasing concentrations (0.1-10 nM)

of [3H]haloperidol (Specific Activity, 50 Ci/mmol, New England

Nuclear, Boston, MA). Spiperone (50 times the [3H]haloperidol

concentration) was included in all incubations in order to block

[3H]haloperidol binding to D 2 dopamine and 5 -HT2 serotonin

receptors (89, 90). Non-specific binding was defined by the

presence of 10 pM DTG. After a 90 min. incubation,

membrane-bound [3H]haloperidol was separated from free

radioligand in a cell harvester (Biomedical Research and

Development Laboratories) by rapid filtration through Whatman

GF/B glass fiber filters (Biomedical Research and Development

Laboratories) that had been pretreated with 0.001% or 0.5% polyethylenimine (PEI) to reduce non-specific binding. The filters and entrapped membranes were washed at room temperature for

10 seconds with 5 mM Tris-HCI buffer (pH 8.0). Filter discs were collected in 6 ml liquid scintillation vials (Wheaton Scientific,

Millville, NJ), and 3 ml liquid scintillation cocktail (Formula-989,

Biotechnology systems, NEN Research Products, Boston, MA) was added to each vial. Liquid scintillation vials were then counted in a scintillation counter (Beckman LS6000IC). For competition binding assays, rat spleen membranes (5 mg

wet weight/tube) were incubated under the conditions described

above with 0.7 nM [3H]haloperidol in the presence of 35 nM

spiperone plus increasing concentrations of competing drugs.

Non-specific binding again was defined as that occurring in the presence of 10 jo.M DTG. The following drugs were used in

competition binding assays: haloperidol metabolite I, haloperidol

metabolite II, (+) and (-)-pentazocine, (+) and (-)-butaclamol,

(+) and (-)-3-PPP, (+) and ( - ) - S K F 10,047, DTG, rimcazole, TCP,

and PCP.

Data Analysis

Saturation binding and drug competition data were analyzed

using the MacLIGAND nonlinear curve fitting program (91) in order to obtain Kd (ligand concentration at which 50% of binding sites are

occupied at steady state), K j and Bmax (number of binding sites per

cell or unit protein) values for drugs. A least squares fit to a

logarithm-probit analysis was used to calculate EC 50 (drug concentration at which 50% of maximum effect is achieved) values for drugs in suppression of ConA-induced proliferation assays.

Analysis of the relationship between drug potencies in binding and proliferation experiments was performed using PROC REG in the

SAS computer package. Because we expected relationships to be relative, rather than absolute, data were examined in a logarithmic format, comparing geometric, rather than arithmetic means.

Although a total of 15 compounds were tested in proliferation and competition assays, seven were used to test whether sigma receptors modulate cell function. These reference compounds were chosen before this series of experiments was performed.

They were haloperidol, haloperidol metabolite II, DTG,

(+)-pentazocine, (+)-SKF 10,047, (+)-3-PPP, and (-)-butaclamol.

These seven reference compounds were examined with a linear regression analysis. The leverage of individual points was examined to determine their influence on the regression. Cook’s distance statistic was used as a test for outliers. The remaining drugs were then compared to the regression in order to determine whether they fell within the 95% prediction interval of the reference compounds. Unpaired t-te s t was performed using StatView 4.0 for

B cell proliferation assays. CHAPTER III RESULTS

A. Sigma Receptors and T Cell Function

Sigma Agonists Modulate ConA-induced Splenocyte Proliferation

A series of fifteen compounds were tested as described in

Materials and Methods for their ability to influence ConA-induced rat splenocyte proliferation in vitro, which is a commonly used measure of T cell activation. A stimulus of 0.5 pg/ml ConA was used for all experiments. This concentration was in the linear range of the ConA dose response, and induced somewhat less than

50% of the maximum proliferative response that could be obtained using higher lectin concentrations (Figure 1). In this system, all fifteen drugs suppressed mitogenic responses in a dose-dependent manner. Representative experiments with the prototypic sigma agonists haloperidol , (+)-pentazocine and DTG are shown in

Figure 2. Controls were performed for all diluents required to solublize drugs, which were the combination of ethanol, 100 mM

23 Figure 1. ConA-induced rat splenocyte proliferation dose response curve. The concentration of ConA used in subsequent experiments, was 0.5 |ig/ml, which was in the linear range of the curve and induced somewhat less than 50% of the maximum proliferative response. The figure is a representative experiment, which was replicated five times.

24 100000

8 0 0 0 0

6 0 0 0 0 E Q. O 4 0 0 0 0

experimental 20000 condition 0.5 ]ig/ml

.01 100

Figure 1

NJ t n Figure 2. Three classical sigma agonists, — haloperidol, (+)- pentazocine and DTG, — suppress Con A-induced rat splenocyte proliferation in a dose-dependent manner. Conditions were as specified in Materials and Methods. Data shown are from a representative experiment. Each point represents average values of three replicates, and each titration curve was repeated three times. The other twelve drugs were tested in an identical manner.

26 %Control [3H]Thymidine 0 0 1 120 0 8 0 4 0 6 20 0 10 -10

a □ (+)- (+)- Pentazocine □ Haloperidol • DTG

10 -9

10 -8

10 Figure 2 rg (M) Drug

-7

10 -6

10

-5

10 -4 28 acetic acid or deionized water. The cpm values of experimental group were then compared to those of control group which contained vehicle(s) only and was considered as 100% proliferation. The EC 50 values of the drugs ranged from 2 x10'7 M to 6 x1 O'5 M. The rank order of drug potencies (from most to least potent) was: haloperidol metabolite II > haloperidol > rimcazole >

(+)-butaclamol > TCP > (-)-butaclamol > PCP > (+)-pentazocine =

DTG > haloperidol metabolite I > (+)-3-PPP > (-)-pentazocine >

(+)-SKF 10,047 > (-)-SKF 10,047 > (-)-3-PPP (Table 1).

Haloperidol, one of the most potent drugs at suppressing

ConA-induced proliferation of rat spienocytes, also is a potent antipsychotic drug which a c ts as a D2 dopamine (13). Since haloperidol binds to sigma and D 2 receptors with similar affinity, some of the immunosuppressive effects of haloperidol might have been mediated through its actions at D 2 receptors. This does not appear to be the case, as (-)-sulpiride, a potent D 2 antagonist, had no effect on ConA-induced proliferation

(Figure 3).

Spienocytes Have [3H]Haloperidol-labelable Sigma-1 Receptors

Although sigma-1 receptors have been previously identified and characterized on human peripheral blood leukocytes and in rat spleen (45, 46), their presence had not been demonstrated on the isolated rat spienocytes used for the present functional assays. As may be seen in Figure 4A and B, washed membranes of spienocytes isolated on 70% Percoll gradients bound 29

Table 1. Pharmacological specificity of [ 3H]haloperidol-labeled sigma-1 receptors in rat spleen ( K j ) and potency of drugs suppressing ConA-induced proliferation of rat spienocytes (EC 50)

Drug EC 50 (a) (|iM) K|(a) (nM)

1. haloperidol metabolite II .197 ± .081 ( 4 ) 5.67 ± 0.16 ( 3 ) 2 . haloperidol .260 ± .053 ( 3 ) .654 ±.012 (b)(3) 3. rimcazole .290 ± .036 ( 3 ) 448 ± 117 ( 3 ) 4. (+)-butaclamol 4.00 ± 1.21 ( 3 ) 1420 ± 180 ( 3 ) 5. TCP 5.48 ± 1.70 ( 4 ) 2530 ± 280 ( 3 ) 6 . (-)-butaclamol 5.90 ± 1.83 ( 3 ) 31.0 ± 4.0 (3) 7. PCP 9.82 ± 2.44 (5) 5980 ± 1690 (3) 8 . (+)-pentazocine 12.8 ± 4.8 (3) 20.3 ± 3.6 (3) 9. DTG 12.9 ± 2.9 (3) 69.6 ± 12.5 ( 3 ) 10. haloperidol metabolite I 20.4 ± 6.4 (3) 389 ± 23 (3) 11. (+)-3-PPP 26.5 ± 17.5 (3) 68.9 ± 23.8 ( 3 ) 12. (-)-pentazocine 30.7 ± 10.5 (3) 28.5 ± 3.8 ( 3 ) 13. (+)-SKF 10,047 31.2 ± 9.9 (3) 233 ± 20 (3) 14. (-)-SKF 10,047 55.6 ± 36.9 (3) 2110 ± 310 ( 3 ) 15. (-J-3-PPP 57.7 ± 26.7 (3) 207 ± 41 ( 3 )

(a) Values represent arithmetic means ± SEM of (number of independent experiments). (b) This is a Kd value. Figure 3. The classical dopamine D 2 receptor antagonist,

(-)-sulpiride, has no effect on ConA-induced rat splenocyte proliferation. Conditions were as specified in Materials and

Methods. Each point represents average values of three replicates.

The figure is a representative experiment, which was replicated twice.

30 200

CD 1 5 0 C

E >, sz h; 100 roX

c o 5 0 CJNO 0s

0 i______-7 -6 -5 10 10 10 10 ' 4 10 ~3

(-)-Sulpiride (M)

Figure 3

03 Figure 4. Specific binding sites for [3H]haloperidol are present on membranes of rat spleen (A) and rat spienocytes (B). As described in Materials and Methods, membrane preparations were incubated with 0.078-10 nM [3H]haloperidol in the presence of 50-fold excess spiperone (3.9-500 nM) to block dopamine and serotonin receptors, and nonspecific binding was defined using 10 pM DTG. The figures shows the pooled data from three experiments for each tissue.

32 Bound [3H]Haloperidol (fmol/mg protein) 2000 Cn O Cn O O O O O O

r \ j

co n: :c n Q)_ <5‘ o c T3 “T (D O

oo

o

ee Bound [3HJHaloperidol (fmol/mg protein) 0 0 0 3 2000 1000 0 i 0 . SpienocytesB. 3H]Haloperidol [3 Figure 4BFigure 6 4 (nM) A A ■A— 8 10 35

[3H]haloperidol in a manner comparable to membranes derived

from whole spleen. In both tissues, binding data gave linear

Rosenthal plots (Figure 5A and B ) , indicating the presence of a

single class of sites, with Kd values (mean + S.E.M.) of 0.65 ± 0.12

nM and 1.07 + 0.27 nM in whole rat spleen and rat spienocytes,

respectively. As may be seen from the B m a x values (the number of

binding sites/unit tissue) in Table 2 , there was an enrichment of

binding sites in spienocytes relative to whole spleen, indicating

that the cells we isolated, rather than structural elements of the

spleen, were a major source of sigma-1 receptors in this tissue.

Regulation of ConA-induced Rat Splenocyte Proliferation by

Sigma Agonists Correlates with the Pharmacology of Splenic

Sigma-1 Receptors

The ability of our te s t compounds to compete with

[3H]haloperidol for binding at sigma-1 receptors was determined in competition binding assays, as illustrated in Figure 6. K j values at

[3H]haloperidol-labeled sigma-1 receptors (mathematically equal to the drugs’ Kd values) were determined for each compound tested in the ConA proliferation assays. In accordance with previous studies

(46), the [3H]haloperidol-labeled sites in rat spleen displayed a pattern of drug potency typical of the pharmacology of sigma-1 receptors. Haloperidol was the most potent compound, while PCP and TCP were least potent. Furthermore, there was selectivity for

(+)-stereoenantiomers of pentazocine, SKF 10,047 and 3-PPP, and for (-)-butaclamol. The rank order of potency in the binding assay Figure 5. Rosenthal plots of the same data as in Figure 4, which show straight lines, indicative of a single class of drug binding site.

Isolated rat spienocytes (A) had higher densities (Bmax) of

[3H]haloperidol binding sites than the rat spleens (B) from which they were obtained. Data from three experiments are shown, each point being the average of three replicates within one experiment.

36 3000

2 5 0 0 -

CD CD

"O C 3 O CO 1 5 0 0 o

§_ 1 0 0 0 c o

5 0 0 '

1000 2000 3 0 0 0

Specific Bound

Figure 5A

CO-'si 3000

2 5 0 0 B

2000

1 5 0 0

A 1000

A 5 0 0

■------1------1_____ i_\ ■ 2 0 0 0 3 0 0 0 4 0 0 0

Specific Bound Figure 5B

CO 00 39

Table 2. Splenic [3H]haloperidol binding sites are present on

isolated splenocytes

Tissue Kd

______loM ) (fmol/mg protein)

Spleen 0.65 + 0.12 1377 ±183

Splenocytes 1.07 + 0.27 2560 ± 502 (b)

(a) Values represent simultaneous analysis ± SEM of three independent

experiments.

(b) Splenocyte Bmax is equivalent to 2582 ± 506 sites/cell. Figure 6. Using competition binding assays, splenic binding sites for 0.7 nM [3H]haloperidol (in the presence of 35 nM spiperone) were demonstrated to have the pharmacology of sigma-1 receptors.

Representative curves using the prototypic sigma ligands haloperidol, (+)-pentazocine and DTG are shown in the figure. The other twelve drugs were tested in an identical manner.

Experiments were carried out as described in Materials and

Methods. Each point represents triplicate measurements, and all experiments were repeated three times.

40 120

100 - T3 C 3 O “ 80 - o ■g ‘i—

I 4 0 ' • Hal oper idol < £ ° (+)Pentazoci ne

20 - A DTG

0 -14 -12 -10 -8 -6 -4 10 10 10 10 10 10 Drug (M)

Figure 6

-p>. 42

(from most to least) was: haloperidol > haloperidol metabolite II >

(+)-pentazocine > (-)-pentazocine >_(-)-butaclamol > (+)-3-PPP >

DTG > (-)-3-PPP > (+)-SKF 10,047 > haloperidol metabolite I >

rimcazole > (+)-butaclamol > (-)-SKF 10,047 > TCP > PCP (Table

1 )- Competition binding assays were performed using membrane

preparations from homogenized spleens, whereas proliferation

assays utilized splenocytes teased from spleen and isolated on

Percoll gradients. Since saturation binding studies demonstrated

the existence of only a single class of binding sites in spleen, and

this site also occurred on isolated splenocytes derived from whole

spleen (Figure 4 and 5, Table 2), we did not deem it necessary to

repeat the pharmacological characterization of drug binding sites

using isolated splenocytes as a tissue source.

Of the fifteen compounds tested, the seven most sigma-1

selective agonists, haloperidol, haloperidol metabolite II, DTG,

(+)-pentazocine, (+)-SKF 10,047, (+)-3-PPP, and (-)-butaclamol,

were chosen as reference compounds. Five drugs, (-)-pentazocine,

(-)-SKF 10,047, (-)-3-PPP, (+)-butaclamol, and haloperidol

metabolite I, were chosen to serve as contrasting, less potent

stereoenantiomers or metabolites of the reference compounds.

Although it is a weak sigma agonist and is considerably more

potent at modulating several ion channels in cells (92, 93), PCP was

also te s te d because of its historical association with sigma

receptors (94) and immune modulation (59-61). TCP (95) and 43

rimcazole (14, 96, 97) were selected as a PCP analog and newly

described sigma antagonist, respectively.

To determine whether the biological effect was due to actions

at sigma-1 receptors, drugs’ EC 50 values in the proliferative assay

were compared on a logarithmic scale with their binding assay K j

values at sigma-1 receptors. These values are listed in Table 1.

Because of the actions of sigma agonists at multiple receptors

(discussed in Introduction), only the seven most sigma-1 selective te s t compounds, haloperidol, haloperidol metabolite II, DTG,

(+)-pentazocine, (+)-SKF 10,047, (+)-3-PPP, and (-)-butaclamol, were used to establish this correlation, which was quite strong

(adjusted r = 0.86). The slope of the correlation line (0.98) had a value approaching unity (Figure 7).

As previously stated, PCP was initially included in the reference group because of its historical role in the field of sigma receptors (see Introduction). By this analysis, a weak, but marginally significant regression (r = 0.63, P approx. 0.05) was obtained. However, PCP seemed to fall outside the pattern of the other compounds. The average leverage of the points was 0.25, but the leverage of PCP was 0.6, indicating that PCP exerted an unusually large influence on the position of the line. A statistical te s t to determine whether PCP could be declared an outlier was inconclusive (P = 0.148, F ^ j) statistic forthe size of Cook’s distance = 2.67). Because in non-immune cells PCP is known to have a variety of actions in addition to those at sigma receptors (92,

93), we felt it should be removed from the reference group if there 44

was any question as to its actions in this system. As discussed

above, this resulted in a strong correlation (adjusted r = 0.86)

between binding and function for the remaining seven compounds

(Figure 7)

When the compounds not used to make the correlation were

considered, four, (-)-pentazocine, (-)-SKF 10,047, (-)-3-PPP and

haloperidol metabolite I, fell within the 95% prediction interval of

the seven reference standards. Three, (+)-butaclamol, PCP and

TCP, fell at the borders of the prediction interval, and rimcazole (a

putative antagonist) was clearly an outlier (Figure 8).

B. Sigma Receptors and B Cell Function

Mouse B lymphoma Cell Line A20 Has [3H]Haloperidol-labelable Sigma-1 Receptors

A20, a commonly used B lymphoma cell line, was used as a

source of pure B lymphocytes. Saturation binding assays using the

membrane preparations from these cells were performed under the same conditions used for rat spleen and rat splenocytes. Specific and saturable binding was obtained, and nonlinear curve fitting data analysis indicated the presence of a single class of sites, with

Kd value (mean + S.E.M.) of 1.09 ± 0 .3 5 nM and Bmax value of 8 8 .7

± 21.5 f mo l/mg protein or 17826± 4323 sites/cell (Figures 9 and

10). Although A20 is a mouse B cell line, its binding of

[3H]Haloperidol was comparable to that of rat spleen and rat Figure 7. The ability of seven reference sigma agonists, haloperidol metabolite II (1), haloperidol (2), (-)-butaclamol (6),

(+)-pentazocine (8), DTG (9), (+)-3-PPP (11) and ( + ) - S K F 1 0 , 0 4 7

(13), to suppress Con A-induced proliferation (EC 50 values) correlated highly (r = 0.86) with their potency in binding

[3H]haloperidol-labeled splenic sigma-1 receptors ( K j values). The

9 5 % prediction interval of the regression is indicated by the dashed lines. Numbers refer to drug’s potency rank in suppressing

ConA-induced proliferation (Table 1 ) .

45 o>

oLD CJ LU

Slope = 0.98 r = 0.86

-10 9 8 7 6 5 Ki (log M)

Figure 7

CD Figure 8. Haloperidol metabolite I (10), (-)-pentazocine (12),

(-)-SKF 10,047 (14) and (-)-3-PPP (15) fell within the 95% prediction interval (dashed lines) of the regression line of the seven reference sigma agonists (circled numbers, see Fig. 7).

Numbers identify drugs by their relative potencies to suppress Con

A-induced splenocyte proliferation (Table 1). PCP (7), TCP (5) and

(+)-butaclamol (4), fell at the limits of the prediction interval. The putative sigma antagonist rimcazole (3) was, as expected, an outlier in this relationship.

47 -3

- 4

m - 5

O) =2o -6 o LO o LU - 7

-8 / /

- 9 -J ------1------1------1______i______L. -10 - 9 - 8 - 7 -6 Ki (log M)

Figure 8 Figure 9. Specific binding sites for [3H]haloperidol are present on membranes of mouse B lymphoma cell line, A20 cells. As described in Materials and Methods, membrane preparations were incubated with 0.078-10 nM [3H]haloperidol in the presence of

50-fold excess spiperone (3.9-500 nM) to block dopamine and serotonin receptors, and nonspecific binding was defined using 10 pM DTG. The figure is a representative experiment, which was replicated three times.

49 250

200

o

S - C 1 5 0 a'® J2 o ro k X D) 100 ro E

13 M— E o w / CD 5 0

0 *f -----■------L. -1 ______I______I______I 0 4 6 . 8

[3H]Haloperidol (nM)

Figure 9 Figure 10. Rosenthal plots of the same data as in Figure 9, which

show straight lines, indicative of a single class of drug binding site.

Data from a representative experiment are shown, each point being the average of three replicates within one experiment.

51 —I______1______1______I_ 50 TOO 150 200

Specific Bound

Figure 10 cro n 53 splenocytes. Their Kd values are very close, all around 1 nM. A20 cells had more binding sites per cell, but less binding sites per mg of protein compared to rat splenocytes. This is probably due to the size of A20 cells. Because A20 cells are constantly dividing cells, they appear under the microscope to be much larger than splenocytes. Larger cell size would cause greater membrane surface, and space to have more binding sites per cell, but their higher cell membrane mass per cell resulted in less binding sites per unit protein.

Haloperidol and Spiperone Modulate Splenic B Cell Proliferation, But Not Through Sigma Receptors

A series of sigma agonists, haloperidol, DTG, (+) and

(-)-pentazocine, and (-)-butaclamol were tested as described in Materials and Methods for their ability to influence anti-p induced murine splenic B cell proliferation in vitro. A suboptimal concentration of 0.5 pg/ml anti-p antibody was used in these experiments. Unlike the case with ConA-induced T cell proliferation, which was suppressed by sigma agonists, proliferation of B cells was enhanced by haloperidol in a dose dependent manner (Figure 11). However, the other sigma agonists mentioned above did not have any effect, indicating that haloperidol did not exert its effect through sigma receptors in this case. Later, we found spiperone was another drug that had the same effect as haloperidol did on anti-p induced B cell proliferation Figure 11. Haloperidol and spiperone enhance anti-p induced B

cell proliferation in a dose-dependent manner. Conditions were as

specified in Materials and Methods. * p<0.02, ** p<0.0005, ***

p<0.0001 as determined by unpaired t-test comparing control and

drug treated cells. The figure is a representative experiment, which was replicated five times. Each point in the figure represents average values of three replicates. If error bar is not visible, it is smaller than the symbol.

54 20000 * * *

1 6 0 0 0

spiperone

12000 * *

8 0 0 0 haloperidol

4 0 0 0

0 10'7 10'6 TO'5

Drug (M) Figure 11 56

(Figure 11). In the absence of anti-p antibody, neither haloperidol

nor spiperone triggered B cell proliferation.

Haloperidol and spiperone were then tested for their ability to influence the threshold concentration of anti-p antibody needed to

trigger proliferation. Haloperidol or spiperone (10 pM) were

incubated at 37°C with cells in microtiter plates for one hour, after which doubling concentrations of anti:p antibodies (0.078 - 1.25

pg/ml) were added to the wells. A constant amount of vehicle used

to solublize drugs was present in all experimental and control

groups. As shown in Figures 12A and 12B, in the absence of

drugs, 0.31 2 pg/ml of antibody was required to produce significant

[3H]thymidine uptake, while in the presence of 10 pM haloperidol or

spiperone this threshold was reduced to 0.156 pg/ml. Again,

neither haloperidol nor spiperone caused significant changes in [3H]thymidine incorporation in the absence of anti-p antibody

(Figures 12A and 12B).

Dopamine, Norepinephrine and Serotonin Modulate Anti-p

Induced B Cell Proliferation

Haloperidol and spiperone have similar pharmacological

profiles. They are potent neuroleptics that bind to multiple

receptors. They can act as D2 , 5HT2, and ai antagonists.

Therefore we determined whether the endogenous ligands of D 2

receptors - dopamine, ai receptors - norepinephrine, and 5 HT2

receptors - serotonin, modulated anti-p induced B cell proliferation.

Dopamine inhibited anti-p induced B cell proliferation in a Figure 12. Haloperidol and spiperone lower the threshold of anti-p needed to trigger B cell proliferation from 0.312 pg/ml to 0.156 pg/ml. (A): 10 pM haloperidol; (B): 1 0 pM spiperone. Conditions were as specified in Materials and Methods. * p<0.003, ** p<0.001,

*** p<0.0001 as determined by unpaired t-test comparing control and drug treated cells. Representatives of two replicate experiments are shown. The bars in the figures represent average values of three replicates.

57 2 4 0 0 0

□ control 20000 ^ haloperidol Q. O a) 1 6 0 0 0 c * * * E >> 12000 JZ r—■M i x CO 8 0 0 0 I! * * 8 ? u a.CD 4 0 0 0 co 0 1------1______I______1___ 0 0 . 0 7 8 0.156 0.312 0.625 1.25

Anti-|i((xg/ml) Figure 12A

Ol 00 24000

20000 □ control £ B Q. spiperone a g 16000 ig I 12000 x on 8000 y h — ‘o a ) Q. 4000 C/3

o

0 0.078 0.156 0.312 0.625 1.25

Anti-|i(|ig/ml)

Figure 12B

Ul CD dose-dependent manner (Figure 13), as did norepinephrine (Figure

14). In contrast, serotonin had no effect at 10‘8 M through 1 0-5 M

concentrations, while at 10'4 M, serotonin caused a slight increase

in [3H]thymidine incorporation (Figure 15). This unusual

dose-response could be accounted for by the presence of serotonin

in the fetal bovine serum used for our culture medium. Using a

commercially available serotonin assay kit (Immunotech Inc.,

Westerbrook, ME), serotonin levels in the culture medium were found to be 3.4 - 3.7 pM.

These results were consistent with haloperidol and spiperone

modulating B cell proliferation by acting as antagonists at dopamine or norepinephrine receptors, but were not consistent with

modulation of proliferation via antagonist actions at serotonin receptors.

Spiperone Antagonizes the inhibitory Effect of Dopamine and Norepinephrine on Anti-p Induced B cell Proliferation

To determine whether spiperone could antagonize the effects of dopamine and norepinephrine, B cells were co-incubated with varying concentrations of spiperone and these two endogenous agonists. Spiperone was used in these assays because it had a stronger modulating effect than haloperidol on B cell proliferation.

As shown in Figures 1 6 and 1 7, proliferation at all concentrations of spiperone was reduced by the presence of either dopamine or norepinephrine. Conversely, the [3H]thymidine incorporation at all Figure 13. Dopamine inhibits anti-p induced B cell proliferation.

Conditions were as specified in Materials and Methods. Figure is a representative experiment, which was replicated four times. Each point in the figure represents average values of three replicates.

61 6000

5000 £ Q. a 4000 a? c '-O £ 3000 -C 4—1 X 2000 co y 4— ' o 1000

-//- 0 10'7 10‘6

Dopamine (M)

Figure 13

O) co Figure 14. Norepinephrine inhibits anti-p induced B cell

proliferation. Conditions were as specified in Materials and

Methods. Figure is a representative experiment, which was

replicated four times. Each point in the figure represents average values of three replicates.

63 6000

5000 -

4000 -

3000 -

2000 -

1000 -

0 -

Norepinephrine (M)

Figure 14 Figure 15. Serotonin enhances anti-p induced B cell proliferation at 10'4 M. Conditions were as specified in Materials and Methods.

Figure is a representative experiment, which was replicated three times. Each point in the figure represents average values of three replicates.

65 6000

5000

4000

3000

2000

1000

0 -// _i______i 0 10'7 10'6 10 ' 5 10 ' 4

5-HT (M) Figure 15 o> 0 5 Figure 16. Spiperone and dopamine antagonize each other’s effects on anti-p induced B cell proliferation. (A) 1 pM dopamine reduces [3H]thymidine incorporation in the presence of various concentrations of spiperone; (B) 5 pM spiperone increases

[3H]thymidine incorporation in the presence of various concentrations of dopamine. Conditions were as specified in

Materials and Methods. * p<0.02, ** p<0.003, *** p<0.0002 as determined by unpaired t-test comparing control (one-drug-treated) and two-drug-treated cells. Each figure is a representative experiment, which was replicated twice. Each point in the figure represents average values of three replicates.

67 Specific [3H]thymidine (cpm) 0 0 0 0 2000 0 0 0 4 0 0 0 6 0 0 0 8 0 Spiperone(M) Figure 16A Figure control 1|iM dopamine 1 4 0 0 0

* * * * * 5(iM spiperone 12000 ------// ■ i * * * Q. O 10000 ! JZ +-> S’ 6 0 0 0 *** m y control u— 4 0 0 0 'o CD Q. to 2000

Dopamine (M)

Figure 16B

CO CD Figure 17. Spiperone and norepinephrine antagonize each other’s effects on anti-p induced B cell proliferation. (A) 1 pM norepinephrine reduces [3H]thymidine incorporation in the presence of various concentrations of spiperone; (B) 5 pM spiperone increases [3H]thymidine incorporation in the presence of various concentrations of norepinephrine. * p<0.03, ** p<0.01, *** p<0.0002 as determined by unpaired t-test comparing control (one- drug-treated) and two-drug-treated cells. Conditions were as specified in Materials and Methods. Each figure is a representative experiment, which was replicated twice. Each point in the figure represents average values of three replicates.

70 Specific [3H]thymidine (cpm) 12000 10000 0 0 0 4 0 0 0 8 2000 0 0 0 6 Spiperone(M) Figure 17A Figure * * * 1|iM norepinephrine 1|iM control ** £ 14000

12000

10000

^ *** 8000

6000

4000 control £

2000

0 /A 10 ' 7 10 ' 6 10 ' 5

Norepinephrine (M)

Figure 17B 73 concentrations of dopamine and norepinephrine was increased in the presence of spiperone (Figures 16 and 17).

D2 antagonists, 5HT2 antagonists, andai Antagonists Do Not Enhance Anti-p Induced B Cell Proliferation

In an attempt to mimic the actions of spiperone and haloperidol, we then tested a variety of dopamine, serotonin (5HT), and ai antagonists. These compounds included D-|, D 2 , D3 , D5 antagonists [(+)-SCH 23390, (-)-sulpiride,

(+)-butaclamol], 5 HT2 and 5HT-|a antagonists [ritanserin, LY-53,857, pindobind-5HT1 A], ai antagonists [phentolamine,

WB-4101 and 5-methyl-urapidil], and a agonists [methoxamine and clonidine]. Over a concentration range of 10'7 M to 1 0*5 M, none of these drugs enhanced anti-p induced B cell proliferation (Table 3). 74

Table 3. Drugs that do not enhance anti-jj. induced murine splenic B cell proliferation*

sigma agonists DTG (+)-pentazocine (-)-pentazocine (-)-butaclamol Di antagonist (+)-SCH-23390 D2 antagonists (-)-sulpiride (+)-butaclamol D3 antagonist (-)-sulpiride D5 antagonist (+)-SCH-23390 5 HT2 antagonists ritanserin LY-53,857 5HTi a antagonist pindobind-5HT1 A ai antagonists WB-4101 5-methyl-urapidil phentolamine a agonists methoxamine clonidine

* Enriched splenic B cells were cultured with drugs, stimulated with anti-p, pulsed with [3 H]thymidine, and harvested as described in Materials and Methods in a manner identical to that of the experiments shown in Figurel 1. Compounds were tested at 10'7, 10 '6 and 10 _5 M concentrations. Within each experiment, triplicate determinations were made for each drug concentration, and all experiments were replicated at least twice. CHAPTER VI DISCUSSION

The present study is composed of two major parts. The first

part is about the influence of sigma receptors on a commonly used

measure of T cell responsiveness - ConA-induced proliferation.

The second part is about the influence of sigma receptors on B cell

proliferation.

Sigma-1 Receptors Modulate T Cell Proliferation

Our study on the effect of sigma receptors on T cell function

resulted in three main observations: (i) sigma-1 receptors are

present in rat spleen and on isolated splenocytes; (ii) a series of 1 5 sigma agonists and related compounds suppressed ConA-induced proliferation in a dose-dependent manner; (iii) drug suppression of

ConA-induced proliferation correlated highly with the pharmacology of sigma-1 receptors in spleen. The regression of

EC 50 values in bioassay versus K j values in binding assays had a slope approaching unity, which is suggestive of a one-for-one

75 relationship between sigma-1 receptor binding events and

biological responses. The reliability of measurements was high for

both Kj and EC 50 values. On a logarithmic scale, the standard

errors of Kj and EC 50 values were 0.07 and 0.19 orders of

magnitude, respectively. Therefore, the correlation was not

significantly biased by errors in measurement. These results

strongly support the hypothesis that drugs can modulate T cell

functional activities via actions at sigma-1 receptors.

It has been reported for more than a decade that PCP and

PCP analogs can depress lymphocyte proliferation and alter

secretion of antibody, IL-1 and IL-2 in vitro (60, 61), and that PCP

can act at sigma receptors (94, 98). At the time of these initial observations, the presumed site of action of these drugs was

referred to as the sigma/PCP receptor (98). Subsequently, two distinct receptors were recognized that could bind PCP (94). These are now termed sigma and PCP receptors. Although certain drugs can bind to both of these, the affinities of binding and the rank order of potency of compounds are distinctive for each (94). With the advent of more selective labels for PCP and sigma receptors,

Wolfe and associates were able to examine the immune system and to determine that sigma receptors were present in abundance in both rat spleen and human circulating leukocytes, while PCP receptors were absent or below the threshold of detectability in these tissues (45,46). However, PCP and TCP have additional actions at a variety of other non-sigma sites, including several potassium channels (92, 93). Since these drugs fell right at the 77

confidence limit of our correlation, we can neither hypothesize, nor

rule out, a contribution by other, non-sigma-1 receptors to the

inhibition of proliferation caused by PCP and TCP.

Because most sigma agonists also act at other receptors, it

would be unrealistic to expect a perfect correlation between

modulation of cell proliferation and drug binding potency at

sigma-1 receptors. Nonetheless, all but three of the fourteen tested

agonists clearly fell within the 95% confidence limits of the

correlation, and gave a collective r value of 0.84, inclusive with the

reference group (r value of the reference group alone was 0.86).

Most significantly, none of the compounds tested were less potent

in suppressing proliferation than was predicted by their relative

binding potencies at sigma-1 sites. We interpret this as very

strong, if not definitive, evidence that sigma-1 receptors on rat

splenocytes are physiologically active.

In our hands, the immunosuppressive effect of sigma agonists

required drug concentrations several orders of magnitude greater than their K d or K j values in binding assays. This is common in

other bioassays of sigma agonists as well (58, 60-62). Five possible

explanations for this are:

1) The buffer used for binding assays is different from tissue

culture medium in many respects, such as pH, ionic strength,

the presence of specific ions, and the presence of exogenous

proteins in serum. In the tissue culture medium in which the

cells encountered drugs, the binding affinities of sigma 78

agonists may be lower than the Kd and Kj values obtained

under optimized binding conditions.

2) Over the course of two-day cultures with live cells, sigma

agonists may be metabolized into inactive forms. Therefore,

higher initial drug concentrations may be required in order to

maintain effective drug levels for the duration of the cultures.

3) Binding sites for sigma ligands have been reported to reside

within cells, as well as on the external surfaces of the plasma

membranes (99). Proliferation assays used intact cells, while

binding assays were done with disrupted cytoplasmic

membranes plus internal membranes and nuclei. If the

biologically active sigma receptors reside within cells, and if

the cytoplasmic membrane is not completely permeable to the

drugs, elevated drug concentrations in the medium may be

required to reach effective levels at the site of the active

receptors within cells.

4) Since we observed an inhibition of activity, it might be a

nonspecific toxic, rather than a receptor-mediated effect.

Sigma agonists have been reported to be toxic to glioma cells

over periods of about one week in culture (100). W e saw no

short-term drug toxicity evidenced by trypan blue vital dye

exclusion. However, in the absence of stimulation,

lymphocytes tend to gradually die off in culture, and it is

difficult to distinguish between slow toxic effects and a

drug-mediated blockade of the mitogenic stimulus (Con A).

Recently, Casellas and associates (66) reported that a novel 79

sigma ligand, SR 31747, could suppress proliferation of

human lymphocytes, and argued that this suppression was

not due to toxicity, because after 1 20 hours of drug-induced

suppression the cells could be restored to full mitogenic

responsiveness by removal of the drug. Regardless, the

correlation of biologic effect with the pharmacology of

sigma-1 receptors in our hands indicates that if this is a toxic

effect, it is mediated through drug actions at sigma-1

receptors.

5) Suppression of ConA-induced proliferation may not occur

unless a very high percentage of sigma-1 receptors are

occupied.

Scenario #1, above, seems improbable because we have

performed binding studies in RPMI-1640 culture medium with the

identical additives as were used for cell cultures, and we observed

no significant changes in affinities of drug binding. At this time, we

cannot distinguish among the remaining four possibilities.

Sigma-1 receptors are selective for the (+)-stereoenantiomers of 3-PPP and the psychotomimetic benzomorphans (pentazocine and SKF 10,047), while sigma-2 receptors display minimal, or the

reverse, selectivity (16). In the proliferation assay, (+)-pentazocine,

(+)-SKF 10,047 and (+)-3-PPP were invariably more biologically potent than their (-)-stereoenantiomers in all experiments, although the differences were less pronounced than expected by their relative binding affinities (Table 1 and Figure 8). It should be emphasized that all stereoisomers of these compounds fell within the 95% prediction interval of the correlation. Therefore, no

actions by these compounds at non-sigma-1 sites are predicted or

supported by our data. However, it is tempting to suggest possible

explanations for this reduction of stereoselectivity. One of these is

that the stereoselectivity that we observed in binding assays is an

artifact of the buffer conditions, and that the receptor is less

stereoselective in physiological buffers. However, in our hands,

competition binding revealed no diminution of stereoselectivity in

tissue culture medium. Therefore, we consider it more probable

that other receptors may have made minor contributions to our

proliferative assays. Receptors of particular interest in this regard

are sigma-2, dopamine, and opiate receptors.

There is evidence that both sigma-1 and sigma-2 receptors

are biologically active. Sigma-1 sites appear to mediate the

inhibition of electrically- and 5-HT-induced contractions of guinea

pig ileum (56), the modulation of muscarinic acetylcholine

receptor-triggered phosphoinositide turnover in brain (29, 30), and

nicotinic receptor function in adrenal chromaffin cells (58). On the other hand, it is most likely that the dystonia from microinjection of sigma agonists into the rat red nucleus (55), and modulation of potassium channels (16, 85) are mediated through sigma-2 sites.

Interestingly, sigma-1 and sigma-2 receptors can coexist. For example, rat liver contains approximately 25% sigma-1 sites and

75% sigma-2 sites (17); rat brain also contains both sites (101). Our preliminary experiments support the notion that sigma-2 receptors are also present along with sigma-1 receptors in immune tissues. Using [3H]DTG under sigma-2 receptor-selective conditions (16), we

have labeled specific binding sites in splenic homogenates,

isolated splenocytes, and T and B cell lines (Wolfe et al.,

unpublished). However, the data we present here do not correlate

with the pharmacological profile of sigma-2 receptors (18).

Therefore, while a minor contribution by these receptors cannot be

excluded, the overall pattern that we observed cannot be

accounted for on the basis of drug actions at sigma-2 sites.

There is overlapping drug specificity between dopamine,

5 -HT2 (serotonin) receptors and sigma receptors. Haloperidol is

most commonly known for its actions as a dopamine receptor

antagonist (reviewed in 13). However, the carbonyl-reduced

metabolite of haloperidol (haloperidol metabolite II) has almost the

same affinity to sigma receptors as haloperidol, but an 85-fold

lower affinity to dopamine receptors; the chlorophenyl-hydroxy-

piperidine metabolite of haloperidol (haloperidol metabolite I) lacks affinity for dopamine receptors, but has a moderate affinity for sigma receptors (102). Both of these haloperidol metabolites were immunosuppressive (Table 1), and fell within our correlation

(Figure 8). Furthermore, the D 2 dopamine antagonist (-)-sulpiride failed to affect proliferation. Thus, it seems unlikely that D 2

receptors had a significant contribution to the immunosuppressive effect we observed. At this time, we cannot rule out a minor contribution by D 5 dopamine receptors (80), but it should be emphasized that our over-all correlation of suppression of proliferation with sigma-1 receptor binding indicates that sigma-1 82

sites play the predominant role in the present study. We obtained

highly selective labeling of sigma-1 receptors with [3H]haloperidol

by blocking dopamine and 5 -HT2 (serotonin) receptors with excess

spiperone in all radioligand binding assays. We demonstrated the

sigma-1 pharmacology of the sites labeled in this manner (Table 1),

and showed a correlation between these drug binding sites and

regulation of cell proliferation. Therefore, contributions by

dopamine and serotonin receptors to the present findings must be

minimal.

Opiate receptors are known to be present on immune cells,

and to affect a variety of functional activities, but the literature in

this field is inconsistent. Our data are consistent with drug actions

via sigma-1 receptors. However, we consider it probable that

interactions of opiate compounds with sigma-1 receptors, such as

was the case with (-)-pentazocine and (-)-SKF 10,047 in this study,

may be responsible for some of the confusion in the opiate

literature.

Rimcazole is a putative sigma antagonist (14,96,97). As such,

it should have had no effect in this system unless endogenous

sigma agonists were present in the culture medium, or were generated by the cells in culture. Although, as expected, rimcazole was the only clear outlier, it was not expected to affect the system

in the same direction as the agonists. We must assume that this

relatively uncharacterized compound has additional, non-sigma actions that have yet to be described. 83

The mechanism by which sigma-1 receptors modulate T cell

proliferation is unknown, though one might speculate an inhibitory

action on T cell phosphoinositide turnover in splenocytes. Such an

action would be similar to that reported for rat synaptoneurosomes

(29) rat cortical brain slices (103) and pheochromocytoma cells (101).

There have been reports of sigma-1 receptor coupling to guanine

regulatory proteins (25, 26), but no evidence for this was found in

other studies. Dornand and associates (61), found that PCP

prevented resting splenocytes from hyperpolerizing, and reduced

subsequent mitogen-induced depolarization and calcium

mobilization. However, our data indicate that PCP and TCP may

have additional, non-sigma-1 actions on splenocytes. Thus, it is

not clear which receptors are responsible for the phenomena

reported by Dornand and associates.

It is clear that a major obstacle to investigations into the

physiologic role of sigma receptors is the broad cross-reactivity of

many or most sigma agonists and antagonists with other, non­

sigma receptors. Much effort has recently been devoted to developing and characterizing more sigma-selective compounds

(47-50). The intent of the present study was to correlate function with the “classical” pharmacology which has been reported for sigma binding sites. Based on this approach, we did find that sigma-1 receptors modulate ConA-induced T cell proliferation.

Future studies in the immune system, in particular examination of cell signaling and early gene activation, are planned which will utilize several of these newly-developed compounds. 84

Sigma Receptors Do Not Modulate B Ceil Proliferation

The second part of our investigation is about the influence of

sigma receptors on B cell proliferation. This was carried out with

enriched murine B cells from spleen. What we found was: (i) sigma

-1 receptors are present in mouse whole spleen (Whitlock and

Wolfe, unpublished data); (ii) sigma-1 receptors are present in a

mouse B lymphoma cell line, A20; (iii) sigma receptors did not modulate anti-p induced murine splenic B cell proliferation; (iv) the

neuroleptics haloperidol and spiperone potentiated anti-p induced

murine B lymphocyte proliferation in vitro, and lowered the

threshold of antibody needed to trigger proliferation. However,

their actions are not due to actions at known dopamine, serotonin, a1 adrenergic, or sigma receptors.

As mentioned above we have identified sigma-1 receptors in

rat splenocytes, and demonstrated that sigma agonists could

modulate T cell function through these receptors. The second part

of our study was undertaken to determine whether sigma receptors

have comparable physiologic effects on B cells. Our laboratory has shown that sigma-1 receptors are also present in mouse whole spleen membrane preparations (Whitlock and Wolfe, unpublished data). Because of the great difficulty obtaining a large number of purified B cells from mouse spleen, we used a mouse B lymphoma cell line, A20, as a pure source of mouse B cells, for radioligand binding assays. The reason that we switched to a mouse system was that two previous studies on sigma receptors and B cell 85

functions were done with mouse spleen B cells (62, 63). We felt that

we would be able to compare our finding with their results better if

we use mouse instead of rat.

Sigma receptors were reported by Carr and associates (62, 63)

to regulate proliferation and immunoglobulin production by B cells

in vitro. These investigators found that the sigma agonists DTG,

(+)-pentazocine, (-)-pentazocine, and haloperidol could enhance

LPS-induced mouse splenocyte proliferation. However, these

studies were performed with unfractionated splenocyte populations.

Because LPS is not only a B cell mitogen, but also activates

monocytes and macrophages, it is highly likely that cell cross-talk

via cytokine secretion occurred in these cultures. Therefore it is

difficult to conclude from these studies that the reported effects

were due solely to direct actions of drugs at sigma receptors on B

cells. Furthermore, using either all splenocytes or adherent cell-

depleted splenocytes stimulated by LPS, we could not repeat the

results reported by Carr and associates (62). The report by Carr

and associates on modulation of antibody production by sigma

receptors is also questionable. Their dose-response curves were

uninterpretable to us, and while some sigma agonists enhanced

LPS- or PWM- induced immunoglobulin production, others did not

have any effect or even caused inhibition (63). Therefore we

consider the effect of sigma receptors on B cell functions to be

unsettled.

In the present study, we have sought to minimize the effects of other cells on the measured B cell responses by: 1) removing T 86

cells, macrophages and monocytes, and using enriched B cell populations; and 2) using anti-p antibody, which provides a more

selective B cell proliferative stimulus than does LPS. Although we

have identified sigma receptors on mouse spleen and the A20 B

lymphoma cell line, in our studies the classical sigma ligands DTG,

(+)-pentazocine, (-)-pentazocine, and (-)-butaclamol had no effect on anti-p triggered B cell proliferation. We conclude that either: 1)

the modulation of proliferation reported by Carr and associates (62)

was due to drug actions at sigma receptors on contaminating non-B

cells in their cultures; or 2) since LPS and anti-p activate different

intracellular signals, it might be the case that LPS-stimulated B cell triggering events are susceptible to modulation by sigma receptors, while those initiated by crosslinking of cell surface IgM with anti-p antibody are not. In addition, 3) it may be that differentiation and

maturation of B cells to antibody production (63) is sigma receptor- sensitive, while proliferation of B cells is not. Future experiments are planned to distinguish among these three possibilities.

Haloperidol is a potent ligand at sigma receptors (88), at which spiperone may also bind (104). Haloperidol-displaceable

[3H]spiperone binding sites have been observed on lymphocytes

(73), primarily on B cells (74). In the present study, we found that haloperidol and spiperone could modulate B cell proliferation in vitro. Unlike the case with T cells, which are suppressed by sigma agonists, proliferation of B cells was enhanced in a dose-dependent manner by these two compounds. However, other sigma agonists had no effect on anti-p induced B cell proliferation, 87

indicating that haloperidol and spiperone did not exert their effects

through sigma receptors in this case.

Haloperidol and spiperone are largely used as antagonists of

dopaminergic-D 2 and -5HT2 receptors in the central

nervous system (CNS). Because of its strong antagonism at CNS

D2 receptors, haloperidol is a commonly prescribed neuroleptic

used in the treatment of psychotic disorders. Haloperidol and

spiperone also have moderate affinity at ai adrenergic receptors.

The present findings indicate that haloperidol and spiperone act

not only in the CNS, but also directly on B lymphocytes. Therefore,

our next working hypothesis was that haloperidol and spiperone act

on B cells through one of the major non-sigma receptors, at which they are known to act in non-immune cells: dopamine (D 2 ),

serotonin ( 5 HT2 ), or adrenergic (ai) receptors (12, 105).

Dopaminergic, serotonergic and adrenergic receptors are

known to be present on immune cells. Although the existence of dopamine D 2 receptors on immune cells is still a matter of debate

(76, 79), D 3 and D 5 receptors have been identified on human peripheral blood lymphocytes by both molecular biology and radioligand binding techniques (77, 78, 80, 81). Serotonin has been shown to regulate the functions of T cells, pre-B cells, macrophages, and NK cells (82, 83, 106-108), and serotonin 5HTia receptors have been identified on activated human T cells (83).

(3-Adrenergic receptors have been demonstrated on lymphocytes, macrophages and granulocytes (70, 72,109). Although evidence for the presence of a-adrenergic receptors on immune cells is mostly 88

indirect (71,110-112), they may be present on low frequency cell

subpopulations.

Consistent with the notion that haloperidol and spiperone

might be acting as dopaminergic or adrenergic antagonists in our

system, we found that the endogenous agonists for these receptors

(i.e., dopamine and norepinephrine) down-regulated B cell

proliferation (Figures 13 and 14). Spiperone reversed the effects of

both dopamine and norepinephrine in that it increased

[3H]thymidine incorporation (Figures 16 and 17). However these

experiments did not produce definitive shifts to the right in

dose-response curves as expected for receptor antagonism. These

experiments (Figures 16 and 17) are consistent with two

interpretations, either: 1) the measured responses are simply the

additive effects of actions at two or more entirely different receptor

systems, or 2) agonist (dopamine or norepinephrine) is present in the culture medium, or is released in an autocrine manner by the cells in culture. To test this, other adrenergic and dopaminergic antagonists were screened in order to determine whether they could mimic the effects of haloperidol and spiperone.

Li and associates (113) demonstrated that norepinephrine inhibits anti-jo. induced B cell proliferation by acting on

P-adrenergic receptors. However, since haloperidol and spiperone have little or no activity at p-receptors (12,105), they most likely act through a different mechanism in our system. Therefore we tested the a-adrenergic antagonists phentolamine, WB-4101 and

5-methyl-urapidil. These compounds had no effect on B cell 89

proliferation; nor did the a-adrenergic agonists methoxamine and

clonidne (Table 1). Therefore, it is likely that haloperidol's and spiperone's effect is not mediated by a or p-adrenergic receptors.

Haloperidol and spiperone are antagonists with high affinity

at dopamine D 2 receptors, and moderate to low affinity at D 3 and

D5 sites. However, because the D2 antagonist (+)-butaclamol, the

D2 /D3 antagonist (-)-sulpiride, and the D 1/D5 antagonist

(+)-SCH-23390 had no effect (Table 3), it is most likely that

haloperidol and spiperone did not potentiate B cell proliferation in

our cultures through antagonist actions at D-|, D 2 D3 or D 5

receptors.

This does not indicate, however, that dopamine receptors are

absent from or do not modulate the functions of murine B cells, or

that haloperidol and spiperone could not act at those receptors to

modulate function if dopamine were present in our cultures. To the

contrary, exogenously added dopamine was found to suppress

proliferation (Figure 13), and spiperone opposed this effect

(Figures 1 6 A and 16B). Interestingly, it may be that dopamine

inhibits proliferation indirectly, by suppressing autocrine

production of prolactin and thyrotropin by B cells in culture.

Prolactin is an immunopermissive hormone (114-116), and

thyrotropin can enhance lymphocyte proliferation (117). These

hormones are made by lymphoid cells, and dopamine has been

reported to cause B lymphoma cell death by inhibiting prolactin and thyrotropin release (118). Regardless, our results suggest that

dopamine was not present in our culture medium, and was not 90

secreted by the cells in an autocrine manner. Thus, in the absence

of agonist, no alteration of functional activity would be expected by

dopamine antagonists.

Exogenous serotonin (5-hydroxytryptamine, 5HT) did not

suppress proliferation. It had no effect on B cell proliferation at 10 (iM or lower concentrations, while at 100 pM a slightenhan cement

was observed. This dose-response was probably due to the

presence of serotonin in the fetal bovine serum used for cultures,

which we determined to be at a final concentration in the medium of 3.4 - 3.7 pM. Regardless, because the 5HT2 antagonists ritanserin

and LY-53,857 failed to mimic the actions of haloperidol and

spiperone, and because exogenous serotonin influenced the

proliferative response in the same direction as did haloperidol and

spiperone, it is unlikely that serotonin receptor antagonism is the

basis of haloperidol's and spiperone's actions in this system.

It is possible that haloperidol and spiperone act on B cells through separate mechanisms. Although these two compounds

bind almost the same receptors, there are exceptions. Most

importantly, spiperone is a potent 5 HTia receptor antagonist, while

haloperidol has low affinity at this site (12, 105). To this end, the

5HT-ia receptor antagonist pindobind-5HT1 A was tested, and found

not to enhance B cell proliferation. Therefore, we do not believe the present phenomenon was mediated by 5 HTia receptors on B cells.

There have been previous reports of effects of haloperidol or spiperone on immune activities. As mentioned above, our studies have shown that haloperidol can act at sigma-1 receptors to inhibit

ConA-induced rat splenocyte (presumably, T cell) proliferation.

Others have found that haloperidol could inhibit natural killer

activity or Epstein-Barr virus infection of B lymphocytes (119, 120).

Sharpe and associates (121), found that systemic or topical

spiperone could inhibit cutaneous contact hypersensitivity in mice,

and that its mechanism of action was apparently independent of

either dopamine or serotonin receptors. In the present

investigation, haloperidol and spiperone modulated B cell

proliferation by an unkown mechanism.

Summary And Conclusions

In summary, we have found that suppression of ConA-induced

proliferation correlates with the pharmacology of sigma-1 receptors

on splenocytes. Most other physiological assays involving sigma

receptors have utilized preparations containing neuronal elements,

and elicited responses with electrical stimulation or by means of

neurotransmitters such as norepinephrine or 5-HT (29, 56, 58,

122-124). To the best of our knowledge, the splenocyte proliferative

response to the mitogen, Con A, is the first demonstration of sigma

receptors modulating a function that does not involve neuronal

elements or induction of responses by neurotransmitters or

neuroactive drugs. This implies that sigma agonists have a direct effect on immune cells. Since sigma receptors have been identified in the nervous, endocrine and immune systems (40,45,46), it is 92 reasonable to suggest that endogenous sigma ligands may play a important role in neuroimmune modulation.

Also, we have shown that sigma receptors have no effect on anti-p induced B cell proliferation. This finding is in contrast to other reports (62, 63) that sigma receptors modulate proliferation and immunoglobulin production by heterogeneous splenocyte populations stimulated with LPS. In addition, the present study reports that the neuroleptics haloperidol and spiperone enhance mouse splenic B cell proliferation triggered by anti-p antibody.

Among the endogenous agonists haloperidol and spiperone are known to antagonize, dopamine and norepinephrine inhibited and serotonin enhanced B cell proliferation in our system. Although in our hands spiperone could oppose the suppression of proliferation by dopamine and norepinephrine, the inability of other antagonists to enhance proliferation suggests it is unlikely that the effects of haloperidol and spiperone in this system are due to actions at Di,

□ 2 . D3 . D5 , o ra i, receptors. By the same logic, it is also unlikely that the enhancement of proliferation was caused by actions at

5 HT2 , 5HTi a or sigma receptors. The specific mechanism by which haloperidol and spiperone modulate anti-p induced mouse splenic B cell proliferation remains to be determined.

There is significant exposure of humans to compounds that can act at sigma receptors. Among these are the prescribed drugs haloperidol (haldol), (thorazine), perphenizine, , pentazocine and antiinflammatory steroids, and the abused drugs PCP, , and possibly anabolic steroids. Except for steroids, these drugs are used primarily for their effects at other receptors in the brain. However, the widespread distribution of sigma receptors in the central nervous, endocrine and digestive systems, and our demonstration that sigma-1 sites are present and functional in the immune system, indicates that these sigma agonists may have direct actions on multiple target organs that result in more global physiological effects.

Specifically, alteration of mental, endocrine, and immune status by prescribed and abused sigma agonists may be directly relevant to human health. LIST OF REFERENCES

1. Martin, W.R., C.G. Eades, J.A. Thompson, R.E. Hoppler, and P.E. Gilbert. 1976. The effects of morphine- and nalorphine- like drugs in the nondependent and morphine-dependent chronic spinal dog. J. Pharmacol. Exp. Ther. 197: 517-532.

2. Keats, A.S. and J. Telford. 1964. Narcotic antagonists as analgesics. Adv. Chem. Ser. 45: 170-176.

3. Vaupel, D.B. 1983. Naltrexone fails to antagonize the effects of PCP and SKF 10,047 in the dog. J. Pharmacol. 92: 269-274.

4. Su, T.-P. 1981. Psychotomimetic opioid binding: specific binding of [3H]SKF-10047 to etorphine-inaccessible sites in guinea pig brain. Eur. J. Pharmacol. 75: 81-82.

5. Su, T. 1982. Evidence for sigma opioid receptor: binding of [3H]SKF-1 0047 to etorphine-inaccessible sites in guinea-pig brain. J. Pharmacol. Exp. Ther. 223: 284-290.

6. Mendelsohn, L.G., V. Kalra, B.G. Johnson, and G.A. Kerchner. 1985. Sigma opioid receptor: characterization and co-identity with the phencyclidine receptor. J. Pharmacol. Exp. Ther. 233: 597-602.

7. Zukin, S.R., K.T. Brady, B.L. Silfer, and R.L. Balster. 1984. Behavioral and biochemical stereoselectivity of sigma opiate/PCP receptors. Brain Res. 294: 174-177.

8. Sonders, M.S., J.F.W. Keana, and E. Weber. 1988. Phencyclidine and psychotomimetic sigma opiates: recent insights into their biochemical and physiological sites of action. TINS 11: 37-40.

9 4 95

9. Weber, E., M. Sonders, M. Quarum, S. McLean, S. Pou, and J.F.W. Keana. 1 986. 1,3-di(2-[5-3H]tolyl)guanine: a selective ligand that labels a-type receptors for psychotomimetic opiates and antipsychotic drugs. Proc. Natl. Acad. Sci. USA 83: 8784-8788.

10. Largent, B.L., H. Wikstrom, A.L. Gundlach, and S.H. Snyder. 1 987. Structural determinants of a receptor affinity. Mol. Pharmacol. 32: 772-784.

11. Arnt, J., K.P. Bogeso, A.V. Christensen, J. Hyttel, J. Larsen, and O. Svendesen. 1983. Dopamine receptor agonistic and antagonistic effects of 3-PPP enantiomers. Psychopharmacology 81: 199-207.

12. Leysen, J.E., P.M.F. Janssen, A. Schotte, W.H.M.L. Luyten, and A.A.H.P. Megens. 1993. Interaction of antipsychotic drugs with neurotransmitter receptor sites in vitro and in vivo in relation to pharmacological and clinical effects: role of 5HT2 receptors. Psychopharmacol. 112: S40-S54.

13. Seeman, P. 1981. Brain dopamine receptors. Pharmacol. Rev. 32: 229-313.

14. Ferris, R.M., F.L.M. Tang, K. Chang, and A. Russell. 1986. Evidence that the potential antipsychotic agent rimcazole (BW234U) is a specific, competitive antagonist of sigma sites in brain. Life Sci. 38: 2329-2337.

15. Hellewell, S.B. and W.D. Bowen. 1990. A sigma-like binding site in rat pheochromocytoma (PC12) cells: decreased affinity for (+)-benzomorphans and lower molecular weight suggest a different sigma receptor form from that of guinea pig brain. Brain Res. 527: 244-253.

16. Quirion, R., W.D. Bowen, Y. Itzhak, J.L. Junien, J.M. Mussachio, R.B. Rothman, T.-P. Su, S.W. Tam, and D.P. Taylor. 1992. A proposal for the classification of sigma binding sites. TIPS 13: 85-86.

17. Hellewell, S.B., A.E. Bruce, and W.D. Bowen. 1990. Characterization of "sigma-like" binding sites in rat liver membranes: further evidence for sigma-1 and sigma-2 sites. In New Leads in Opioid Research: Proceedings of the International Narcotics Reserach Conference, International Congress Series #914. J.M. van Ree, A.H. Mulder, V.M. 96

Wiegant, andT.B. van Wimersma Greidanus, eds. Excerpta Medica - Elsevier, Amsterdam, The Netherlands, p. 270-271.

18. Vilner, B.J. and W.D. Bowen. 1992. Characterization of sigma-like binding sites of NB41 A3, S-20Y, and N1 E-11 5 neuroblastomas, C6 glioma, and NG108-15 neuroblastoma- glioma hybrid cells: further evidence for sigma-2 receptors. In Multiple Sigma and PCP Receptor Ligands: Mechanisms for Neuromodulation and Neuroprotection? J.-M. Kamenka and E.F. Domino, eds. NPP Books, Ann Arbor, Ml, p. 341-353.

19. Craviso, G.L. and J.M. Musacchio. 1 983. High -affinity binding sites in guinea pig brain. I. Initial characterization. Mol. Pharmacol. 23: 619-628.

20. Kavanaugh, M.P., B.C. Tester, M.W. Scherz, J.F.W. Keana, and E. Weber. 1 988. Identification of the binding subunitofthe sigma-type opiate receptor by photoaffinity labeling with 1 - (4-azido-2-methyl{6-3H}phenyl)-3-(2-mehtyl{4,6,-3H}phenyl) guanidine. Proc. Natl. Acad. Sci. USA 85: 2844-2848.

21. Kavanaugh, M.P., J. Parker, D.H. Bobker, J.F. Keana, and E. Weber. 1989. Solubilization and characterization of sigma-receptors from guinea pig brain membranes. Journal of Neurochemistry 53: 1575-80.

22. Adams, J.T., J.F.W. Keana, and E. Weber. 1987. Labeling of the haloperidol-sensitive sigma receptor in NCB-20 hybrid neuroblstoma cells with [3H]-(1,3)-di-ortho-tolyl-guanine and identification of the binding subunit by photoaffinity labeling with [3H]-m-azido-di-ortho-tolyl-guanidine. Soc. Neurosci. Abst. 13: 471.7.

23. McCann, D.J. and T. Su. 1991. Solubilization and characterization of haloperidol-sensitive (+)-[3H]SKF-1 0,047 binding sites (sigma sites) from rat liver membranes. J. Pharmacol. Exp. Ther. 257: 547-554.

24. Schuster, D.I., G.K. Ehrlich, and R.B. Murphy. 1994. Purification and partial amino acid sequence of a 28 kDa cyclophilin-like component of the rat liver sigma receptor. Life Sci. 55: PL151.

25. Itzhak, Y. and M. Khouri. 1988. Regulation of the binding of a- and phencyclidine (PCP)-receptor ligands in rat brain membranes by quanine nucleotides and ions. Neurosci. Lett. 85: 147-152. 97

26. Itzhak, Y. 1 989. Multiple affinity binding states of the a receptor: Effect of GTP-binding protein-modifying agents. Mol. Pharmacol. 36:512-517.

2.1. Basile, A.S., I.A. Paul, A. Mirchevich, G. Kuijpers, and C.B. De. 1992. Modulation of (+)-[3H]pentazocine binding to guinea pig cerebellum by divalent cations. Molecular Pharmacology 42: 882-9.

28. Itzhak, Y. and I. Stein. 1 991. Regulation of a receptors and responsiveness to guanine nucleotides following repeated exposure of rats to haloperidol: further evidence for multiple a binding sites. Brain Res. 566: 166-172.

29. Bowen, W.D., B.N. Kirschner, A.H. Newman, and K.C. Rice. 1988. a Receptors negatively modulate agonist-stimulated phosphoinositide metabolism in rat brain. Eur. J. Pharm. 149: 399-400.

30. Brog, J.S. and M.C. Beinfeld. 1990. Inhibition of carbachol-induced inositol phosphate accumulation by phencyclidine, phencyclidine-like ligands and sigma agonists involves blockade of the muscarinic cholinergic receptor: a novel dioxadrol-preferring interaction. J. Pharmacol. Exp. Ther. 254: 952-956.

31. Su, T.-P., A.D. Weissman, and S.-Y. Yeh. 1986. Endogenous ligands for sigma opioid receptors in the brain ("sigmaphin"): evidence from binding assays. Life Sciences 38: 2199-2210.

32. Contreras, P.C., D.A. Dimaggio, and T.L. O'Donohue. 1987. An endogenous ligand for the sigma opioid binding site. Synapse 1: 57-61.

33. Nagornaia, L.V., N.N. Samovilvo, N.V. Korobov, and V.A. Vinogradov. 1988. Partial purification of endogenous inhibitors of (+)-[3H]SKF-1 0047 binding with sigma opioid receptors of the liver. Biull. Eksp. Biol. Med. 106: 314-317.

34. Patterson, T.A., M. Connor, and C. Chavkin. 1994. Recent evidence for endogenous substance(s) for sigma receptors. In Sigma receptors. Y. Itzhak, eds. Academic Press, San Diego, p. 171-189. 98

35. Roman, F.J., X. Pascaus, O. Duffy, D. Vauche, B. Martin, and J.L.L. Junien. 1989. Neuropeptide Y and peptide YY interact with rat brain o and PCP binding sites. Eur. J. Pharmocol. 174: 310-302.

36. Tam, S.W. and K.N. Mitchell. 1991. Neuropeptide Y and peptide YY do not bind to brain sigma and phencyclidine binding sites. Eur. J. Pharmacol. 193: 121-122.

37. Su, T., E.D. London, and J.H. Jaffe. 1 988. Steroid binding at a receptors suggests a link between endocrine, nervous, and immune systems. Science 240: 219-221.

38. Connor, M.A. and C. Chavkin. 1992. Ionic zinc may function as an endogenous ligand for the haloperidol sensitive sigma 2 receptor in rat brain. Mol. Pharmacol. 42: 471-479.

39. Gundlach, A.L., B.L. Largent, and S.H. Snyder. 1986. Autoradiographic localization of o-receptor binding sites in guinea pig and rat central nernous system with (+)-3H-3-(3-hydroxyphenyl)-N-(1 -propyl)-piperidine. J. Neurosci. 6: 1757-1770.

40. Wolfe Jr., S.A., S.G. Culp, and E.B. De Souza. 1989. a-Receptors in endocrine organs: identification, characterization, and autoradiographic localization in rat pituitary, adrenal, testis, and ovary. Endocrinology 124: 1160-1172.

41. Jansen, K.L., M. Dragunow, and R.L. Faull. 1990. Sigma receptors are highly concentrated in the rat pineal gland. Brain Res. 507: 158-160.

42. Su, T.P. and X.Z. Wu. 1990. Guinea pig vas deferens contains sigma but not phencyclidine receptors. Neurosci. Lett. 108: 341-345.

43. Roman, F., X. Pascaud, D. Vauche, and J.L. Junien. 1988. Evidence for a non-opioid sigma binding site in the guinea pig myenteric plexus. Life Sci. 42: 2217-2222.

44. Samovilova, N.N., L.V. Nagornaya, and V.A. Vinogradov. 1988. (+)-[3H]SKF 10,047 binding sites in rat liver. Eur. J. Pharmacol. 147: 259-264. 99

45. Wolfe Jr., S.A., C. Kulsakdinun, G. Battaglia, J.J. Jaffe, and E.B. De Souza. 1988. Initial identification and characterization of sigma receptors on human peripheral blood leukocytes. J. Pharmacol. Exp. Ther. 247: 1114-1119.

46. Wolfe Jr., S.A. and E.B. De Souza. 1992. Sigma receptors in the brain-endocrine-immune axis. In Multiple Sigma and PCP Receptor Ligands: Mechanisms for Neuromodulation and Neuroprotection? J.M. Kamenka and E.F. Domino, eds. NPP Books, Ann Arbor, p. 927-958.

47. Su, T.P., X.Z. Wu, E.J. Cone, K. Shukla, T.M. Gund, A.L. Dodge, and D.W. Parish. 1991. Sigma compounds derived from phencyclidine: identification of PRE-084, a new, selective sigma ligand. Journal of Pharmacology & Experimental Therapeutics 259: 543-50.

48. Bonhaus, D.W., D.N. Loury, L.B. Jakeman, Z. To, A. Desouza, R.M. Eglen, and E. Wong. 1993. [3H]BIMU-1, a 5- Hydroxytryptamine(3) Receptor Ligand in NG-108 Cells, Selectively Labels Sigma-2 Binding Sites in Guinea Pig Hippocampus. J Pharmacol Exp Ther 267: 961 -970.

49. Bonhaus, D.W., D.N. Loury, L.B. Jakeman, S.A.O. Hsu, Z.P. To, E. Leung, K.D. Zeitung, R.M. Eglen, and E.H.F. Wong. 1994. [3H]RS-23597-190, a potent 5-hydroxytryptamine4 antagonist labels sigma-1 but not sigma-2 binding sites in guinea pig brain. J. Pharmacol. Exp. Ther. 271: 484-493.

50. De Costa, B.R. and X.S. He. 1994. Structure-activity relationships and evolution of sigma receptor ligands (1976-present). In Sigma receptors. Y. Itzhak, eds. Academic Press, San Diego, p. 45-111.

51. Puppa, A.D. and E.D. London. 1989. Cerebral metabolic effects of a ligands in the rat brain. Brain Res. 505: 283-290.

52. Steinfels, G.F., G.P. Alberici, S.W. Tam, and L. Cook. 1988. Biochemical, behavioral, and electrophysiologic actions of the selective sigma receptor ligand (+)pentazocine. Neuropsychopharmacology 1: 321-327.

53. Weissman, A.D., M.F. Casanova, J.E. Kleinman, E.D. London, and E.B. De Souza. 1 991. Selective loss of cerebral cortical sigma, but not PCP binding sites in schizophrenia. Biol. Psychiat. 29:41-54. 100

54. Simpson, M., P. Slater, M. Royston, and J. Deakin. 1992. Alterations in phencyclodine and sigma binding sites in schizophrenic brains; effects of disease process and neuroleptic . Schizophrenia Res. 6 :4 1 -4 8 .

55. Matsumoto, R.R., M.K. Hemstreet, N.L. Lai, A. Thurkauf, B. de Costa, K.C. Rice, S.B. Hellewell, W.D. Bowen, and J.M. Walker. 1990. Drug specificity of pharmacological dystonia. Pharmacol. Biochem. Behav. 36: 151-155.

56. Campbell, B.G., M.W. Scherz, J.F.W. Keana, and E. Weber. 1989. Sigma receptors regulate contractions of the guinea pig ileum longitudinal muscle/myenteric plexus preparation elicited by both electrical stimulation and exogenous serotonin. J. Neurosci. 9:3380-3391.

57. Wu, X.Z., J.A. Bell, C.E. Spivak, E.D. London, and T.P. Su. 1991. Electrophysiological and binding studies on intact NCB-20 cells suggest presence of a low affinity sigma receptor. J. Pharmacol. Exp. Ther. 257: 351 -359.

58. Paul, I.A., A.S. Basile, E. Rojas, M.B.H. Youdim, B. de Costa, P. Skolnick, H.B. Pollard, and G.A.J. Kuijpers. 1 993. Sigma receptors modulate nicotinic receptor function in adrenal chromaffin cells. FASEB J. 7: 1171-1178.

59. Fudenberg, H.H., H.D. Whitten, Y.K. Chou, P. Arnaud, A.A. Shums, and N.K. Khansari. 1984. Sigma receptors and autoimmune mechanisms in schizophrenia: preliminary findings and hypotheses. Biomedicine and Pharmacother. 38: 285-290.

60. Khansari, N., H.D. Whitten, and H.H. Fudenberg. 1984. Phencyclidine-induced immunodepression. Science 225: 76-78.

61. Dorn and, J., J.-M. Kamenka, A. Bartegi, and J.-C. Mani. 1 987. PCP and analogs prevent the proliferative response of T lymphocytes by lowering IL2 production: an effect related to the blockade of mitogen-triggered enhancement of free cytosolic calcium concentration. Biochem. Pharmacol. 36: 3929-3936. 101

62. Carr, D.J.J., B.R. de Costa, L. Radesca, and J.E. Blalock. 1991. Functional assessment and partial characterization of [3H](+)-pentazocine binding sites on cells of the immune system. J. Neuroimmunol. 35: 153-166.

63. Carr, D.J.J., S. Mayo, T.W. Woolley, and B.R. de Costa. 1992. Immunoregulatory properties of (+)-pentazocine and sigma ligands. Immunology 77:527.

64. De Costa, B.R. and W.D. Bowen. 1991. Synthesis and characterization of optically pure [3H](+)-azidophenazocine ([3H](+)-AZPH), a photoaffinity label for sigma receptors. J. Labelled Compd. Radiopharm. 29: 443.

65. Garza Jr., H.H., S. Mayo, W.D. Bowen, B. de Costa, and D.J.J. Carr. 1993. Characterization of a (+)-azidophenazocine- sensitive sigma receptor on splenic lymphocytes. J. Immunol. 151: 4672-4680.

66. Casellas, P., B. Bourrie, X. Canat, P. Carayon, I. Buisson, R. Paul, J.-C. Breliere, and G. Le Fur. 1994. Immunopharmacological profile of SR 31 747: in vitro and in vivo studies on humoral and cellular responses. J. Neuroimmunol. 52: 193-203.

67. Paul, I.A., S. Lavastre, D. Floutard, R. Floutard, X. Canat, P. Casellas, G. Le Fur, and J.C. Breliere. 1994. Allosteric modulation of peripheral sigma binding sites by a new selective ligand: SR 31747. J. Neuroimmunol. 52: 183-192.

68. Derocq, J.-M., B. Bourrie, M. Segui, G. Le Fur, and P. Casellas. 1995. In vivo inhibition of endotoxin-induced pro-inflammatory cytokines production by the sigma ligand SR31747. J. Phar. Expt. Ther. 272: 224-230.

69. Walker, J.M., W.D. Bowen, F.O. Walker, R.R. Matsumoto, B. de Costa, and K.C. Rice. 1990. Sigma receptors: biology and function. Pharmacol. Rev. 42: 355-402.

70. Brodde, O.E., G. Engel, E. Hoyer, K.D. Block, and F. Weber. 1981. The beta-adrenergic receptor in human lymphocytes - subclassificaition by the use of a new radio-ligand (±)[125lodo]cyanopidolol. Life Sci. 29: 2189-2198.

71. Titinchi, S. and B. Clark. 1984. Alpha2-adrenoceptors in human lymphocytes: Direct characterization by [3H] binding. Biochem. Biophys. Res. Commun. 121: 1-7. 102

72. Abrass, C.K., S.W. O'Connor, P.J. Scarpace, and I.B. Abrass. 1 985. Characterization of the beta-adrenergic receptor of the rat peritoneal macrophage. J. Immunol. 135: 1338-1341.

73. Le Fur, G., T. Phan, and A. Uzan. 1980. Identification of stereospecific [3H]spiroperidol binding sites in mammalian lymphocytes. Life Sci. 26: 1139-1148.

74. Uzan, A., T. Phan, and G. Le Fur. 1 981. Selective labelling of murine B lymphocytes by [3H]spiroperidol. J. Pharm. Pharmacol. 33: 102-103.

75. Shaskan, E.G., M. Ballow, M. Lederman, S.L. Margoles, and R. Melchreit. 1984. Spiroperidol binding sites on mouse lymphoid cells. J. Neuroimmunol. 6: 59-66.

76. Ovadia, H. and O. Abramsky. 1987. Dopamine receptors on isolated membranes of rat lymphocytes. J. Neurosci. Res. 18: 70-74.

77. Takahashi, N., Y. Nagai, S. Ueno, Y. Saeki, and T. Yanagihara. 1992. Human peripheral blood lymphocytes express D5 dopamine receptor gene and transcribe the two pseudo genes. FEBS Lett. 314:23-25.

78. Nagai, Y., S. Ueno, Y. Saeki, F. Soga, and T. Yanagihara. 1993. Expression of the D-3 dopamine receptor gene and novel variant transcript generated by alternative splicing in human peripheral blood lymphocytes. Biochem. Biophys. Res. Commun. 194:368-374.

79. Santambrogio, L., M. Lipartiti, A. Bruni, and R.D. Toso. 1993. Dopamine receptors on human T- and B-lymphocytes. J. Neuroimmunol. 45:113-120.

80. Ricci, A. and F. Amenta. 1994. Dopamine D5 receptors in human peripheral blood lymphocytes: A radioligand binding study. J. Neuroimmunol. 53: 1-7.

81. Ricci, A., F. Veglio, and F. Amenta. 1995. Radioligand binding characterization of putative dopamine D3 receptor in human peripheral blood lymphocytes with [3H]7-OH-DPAT. J. Neuroimmunol. 58: 139-144. 103

82. Aune, T.M., K.A. Kelley, G.E. Ranges, and M.P. Bombara. 1990. Serotonin-activated signal transduction via serotonin receptors on Jurkat cells. J. Immunol. 145: 1826-1831.

83. Aune, T.M., K.M. McGrath, T. Sarr, M.P. Bombara, and K.A. Kelley. 1993. Expression of 5HT1 a receptors on activated human T cells. J. Immunol. 151: 1175-1183.

84. Sibinga, N.E.S. and A. Goldstein. 1988. Opioid peptides and opioid receptors in cells of the immune system. Ann. Rev. Immunol. 6:219-249.

85. Jeanjean, A.P., M. Mestre, J.-M. Maloteaux, and P.M. Laduron. 1993. Is the o2 receptor in rat brain related to the K+ channel of class III antiarrhythmic drugs? Eur. J. Pharmacol. 241: 111-116.

86. Ader, R., D.L. Felten, and N. Cohen, ed. Psychoneuroimmunology, 2nd Ed. 1 991, Academic Press: San Diego.

87. Ader, R., D. Felten, and N. Cohen. 1990. Interactions between the brain and the immune system. Annu. Rev. Pharmacol. Toxicol. 30:561-602.

88. Tam, S.W. and L. Cook. 1984. a opiates and certain antipsychotic drugs mutually inhibit (+)-[3H]SKF 10,047 and [3H]haloperidol binding in guinea pig brain membranes. Proc. Natl. Acad. Sci. USA 81: 5618-5621.

89. Creese, I., K. Stewart, and S.H. Snyder. 1979. Species variations in dopamine receptor binding. Eur. J. Pharmacol. 60: 55-66.

90. Peroutka, S.J. and S.H. Snyder. 1979. Multiple serotonin receptors: differential binding of 3H-5-hydroxytryptamine, 3H-lysergic acid diethylamide and 3H-spiroperidol. Mol. Pharmacol. 16:687-699.

91. Munson, P.J. and D. Rodbard. 1980. Ligand: A versatile approach for characterization of ligand-binding systems. Anal. Biochem. 107:220-239. 104

92. Albuquerque, E.X., L.G. Aguayo, J.E. Warnick, H. Weinstein, S.D. Glick, S. Maayani, R.K. Ickowicz, and M.P. Blaustein. 1981. The behavioral effects of may be due to their blockade of potassium channels. Proc. Natl. Acad. Sci. USA 78: 7792-2296.

93. Blaustein, M.P. and R.K. Ickowicz. 1983. Phencyclidine in nanomolar concentrations binds to synaptosomes and blocks certain potassium channels. Proc. Natl. Acad. Sci. USA 80: 3855-3859.

94. Quirion, R., R. Chicheportiche, K.M. Johnson, D. Lodge, S.W. Tam, J.H. Woods, and S.R. Zukin. 1987. Classification and nomenclature of phencyclidine and sigma receptor sites. Trends Neurosci. 10: 444-446.

95. Vignon, J., R. Chicheportiche, M. Chicheportiche, J.-M. Kamenka, P. Geneste, and M. Lazdunski. 1983. [3H]TCP: a new tool with high affinity for the PCP receptor in rat brain. Brain Res. 280: 194-197.

96. Kennedy, C., G.E. Jarvis, and G. Henderson. 1990. The ligand rimcazole antagonizes (+)SKF 10,047, but not (+)3-PPP, in the mouse isolated vas deferens. Eur. J. Pharmacol. 181: 315-318.

97. Matsuno, K., T. Senda, and S. Mita. 1993. Correlation between potentiation of neurogenic twitch contraction and benzomorphan sigma-receptor binding potency in the mouse vas-deferens. Eur. J. Pharmacol. 231: 451 -457.

98. Quirion, R., R.P. Hammer Jr., M. Herkenham, and C.B. Pert. 1 981. Phencyclidine (angel dust)/o "opiate" receptor: visualization by tritium-sensitive film. Proc. Natl. Acad. Sci. USA 78: 5881-5885.

99. McCann, D.J. and T.P. Su. 1990. Haloperidol-sensitive (+)[3H]SKF-10,047 binding sites ( sigma sites) exhibit a unique distribution in rat brain subcellular fractions. Eur. J. Pharmacol. 188: 211-218.

100. Vilner, B.J. and W.D. Bowen. 1 993. Sigma-receptor-active neuroleptics are cytotoxic to C6 glioma cells in culture. Eur. J. Pharmacol. Molec. Pharm. 244: 199-201. 105

101. Bowen, W.D., P.J. Tolentino, K.K. Hsu, J.M. Cutts, and S.S. Naidu. 1992. Inhibition of the cholinergic phosphoinositide response by sigma ligands: distinguishing a sigma receptor- mediated mechanism from a mechanism involving direct cholinergic receptor antagonism. In Multiple Sigma and PCP Receptor Ligands. J.-M. Kamenka and E.F. Domino, eds. NPP Books, Ann Arbor, Ml, p. 155-167.

102. Bowen, W.D., E.L. Moses, P.J. Tolentino, and J.M. Walker. 1990. Metabolites of haloperidol display preferential activity at sigma receptors compared to dopamine D-2 receptors. Eur. J. Pharmacol. 177:111-118.

103. Candura, S.M., T. Coccini, L. Manzo, and L.G. Costa. 1990. Interaction of sigma-compounds with receptor-stimulated phosphoinositide metabolism in the rat brain. J. Neurochem. 55: 1741-1748.

104. Coccini, T., L. Manzo, and L.G. Costa. 1991. 3H-spiperone labels sigma receptors, not dopamine D2 receptors, in rat and human lymphocytes. Immunopharmacol. 22: 93-106.

105. Goffinet, A.M., J. Leysen, and D. Labar. 1990. In vitro pharmacological profile of 3-N-(2-fluoroethyl)spiperone. J. Careb. Blood Flow Metab. 10: 140-142.

106. Choquet, D. and H. Korn. 1988. Dual effects of serotonin on a voltage-gated conductance in lymphocytes. Proc. Natl. Acad. Sci. USA 85: 4557-4561.

107. Aune, T.M., H.W. Golden, and K.M. McGrath. 1994. Inhibitors of serotonin synthesis and antagonists of serotonin 1A receptors inhibit T lymphocyte function in vitro and cell-mediated immunity in vivo. J. Immunol. 153: 489-498.

108. Young, M.R.I. and J.P. Matthews. 1995. Serotonin regulation of T-cell subpopulations and of macrophage accessory function. Immunol. 84: 148-152.

109. Landmann, R., E. Burgisser, M. West, and F.R. Buhler. 1985. Beta adrenergic receptors are different in subpopulations of human circulating lymphocytes. J. Recept. Res. 4:37-50.

110. Sanders, V.M. and A.E. Munson. 1985. Role of alpha adrenoceptor activation in modulating the murine primary antibody response in vitro. J. Pharmac. Exp. Ther. 232: 395- 400. 106

111. Spengler, R.N., R.M. Allen, D.G. Remick, R.M. Strieter, and S.L. Kunkel. 1 990. Stimulation of alpha-adrenergic receptor augments the production of macrophage-derived tumor necrosis factor. J. Immunol. 145: 1430-1434.

112. Felsner, P., D. Hofer, I. Rinner, S. Porta, W. Korsatko, and K. Schauenstein. 1995. Adrenergic suppression of peripheral blood T cell reactivity in the rat is due to activation of peripheral a 2-receptors. J. Neuroimmunol. 57:27-34.

113. Li, Y.S., E. Kouassi, and J.P. Revillard. 1 990. Differential regulation of mouse B-cell activation by p-adrenoceptor stimulation depending on type of mitogens. Immunology 69: 367-372.

114. Nagy, E., I. Berczi, G.E. Wren, S.A. Asa, and K. Kovacs. 1 983. Immunomodulation by . Immunopharmacol. 67: 231-243.

115. Bernton, E.W., M.T. Meltzer, and J.W. Holaday. 1988. Suppression of macrophage activation and T-lymphocyte function in hypoprolactinemic mice. Science 239: 401-404.

116. Hartman, D., J.W. Holaday, and E.W. Bernton. 1989. Inhibition of lymphocyte proliferation by antibodies to PRL. FASEB J. 3: 2194-2201.

117. Provinciali, M., G. Di Stefano, and N. Fabris. 1992. Improvement in the proliferative capacity and natural killer cell activity of murine spleen lymphocytes by thyrotropin. Int. J. Immunopharmacol. 14:865-870.

118. Braesch-Andersen, S., S. Paulie, and I. Stamenkovic. 1992. Dopamine-induced lymphoma cell death by inhibition of hormone release. Scand. J. Immunol. 36: 547-553.

119. Nemerow, G.R. and N.R. Cooper. 1 984. Infection of B lymphocytes by a human herpesvirus, Epstein-Barr virus, is blocked by calmodulin antagonists. Proc. Natl. Acad. Sci. USA 81: 4955-4959.

120. Solovera, J.J., M. Alvarez-Mon, J. Casas, J. Carballido, and A. Durantez. 1987. Inhibition of human natural killer (NK) activity by calcium channel modulators and a calmodulin antagonist. J. Immunol. 139:876-880. 107

121. Sharp, R.J., A. Chandrasekar, K.A. Arndt, Z.-S. Wang, and S.J. Galli. 1 992. Inhibition of cutaneous contact hypersensitivity in the mouse with systemic or topical spiperone: topical application of spiperone produces local immunosuppression without inducing systemic neuroleptic effects. J. Invest. Dermatol. 99: 594-600.

122. Massamiri, T. and S.P. Duckies. 1990. Multiple vascular effects of sigma and PCP ligands: inhibition of amine uptake and contractile responses. J. Pharmacol. Exp. Ther. 253: 124-129.

123. Massamiri, T. and S.P. Duckies. 1990. Multiple sites of action of (+)-3-(3-hydroxyphenyl)-N-(1 -propyl)piperidine ((+)-3-PPP) in blood vessels. Eur. J. Pharmacol. 190: 295-303.

124. Massamiri, T. and S.P. Duckies. 1 991. Interactions of sigma and phencyclidine receptor ligands with the norepinephrine uptake carrier in both rat brain and rat tail artery. J. Pharmacol. Exp. Ther. 256: 519-524.