Elsevier

Pharmacology & Therapeutics

Volume 88, Issue 3, December 2000, Pages 281-309
Pharmacology & Therapeutics

Associate editor: D.R. Sibley
α1-Adrenergic receptor regulation: basic science and clinical implications

https://doi.org/10.1016/S0163-7258(00)00092-9Get rights and content

Abstract

Adrenergic receptors (ARs) are members of the G-protein-coupled receptor family, which includes α1ARs, α2ARs, β1ARs, β2ARs, β3ARs, adenosine, muscarinic, angiotensin, endothelin receptors, and many others that are responsible for a large variety of physiologic effects through G-protein coupling. This review focuses on α1ARs and their regulation at both the mRNA and protein levels. Currently, three α1AR subtypes have been characterized both pharmacologically and at the gene level: α1aAR, α1bAR, and α1dAR. These are expressed in a species- and tissue-dependent manner. Mutagenesis approaches have been extremely valuable in the identification of key residues that govern α1AR ligand binding and signaling. These studies reveal that α1ARs have evolved an exquisitely sensitive regulation of their activity in which any disruption of the native structure has profound effects on subsequent function and effector coupling. Significant advances have also been made in the elucidation of signaling pathway components, resulting in the identification of novel pathways that can lead to pathologic conditions. Specific topics include mitogen-activated protein kinase , phosphatidylinositol 3-kinase , and G-protein-coupled receptor cross-talk pathways. Within this context, recent studies identifying underlying transcriptional mechanisms involved in the regulation of the α1AR subtypes is also discussed. Finally, given the potentially important role of α1ARs in the vasculature, as well as in the pathology of many diseases, such as myocardial hypertrophy and benign prostatic hyperplasia, the clinical relevance of α1AR distribution, pharmacology, and therapeutic intervention is reviewed.

Introduction

Drugs targeting adrenergic receptors (ARs) are some of the most widely utilized therapeutic agents in clinical medicine. Activation of ARs provides enhanced inotropy, chronotropy, bronchodilation, vasconstriction, sedation, and analgesia. Inhibition of ARs results in vasodilation, decreased heart rate, inotropy, and relaxation of prostate smooth muscle. Each of these actions is important in treating clinical diseases, such as congestive heart failure, angina, hypertension, benign prostatic hyperplasia (BPH), acute/chronic pain, anesthetic responses, and asthma (reviewed in Smiley et al., 1998). ARs were first divided into α and β by Ahlquist in his landmark manuscript in the Journal of Physiology in 1948. By 1967, Lands had subdivided βARs into β1 and β2 subtypes, and by the mid-1970s, αARs had been subdivided into α1 and α2. These four subtypes are often referred to as “classic” AR subtypes (Schwinn et al., 1991). By the late 1980s and early 1990s, at least 9 AR subtypes (3 α1ARs, α1a, α1b, α1d; 3 α2ARs, α2a, α2b, α2c; and 3 βARs, β1, β2, β3) had been discovered, cloned, expressed stably in cells, and characterized pharmacologically (Fig. 1) (Mizobe et al., 1996, Schwinn et al., 1991a, Schwinn et al., 1991b; for reviews, see Graham et al., 1996, Guarino et al., 1996, Piascik et al., 1996).

ARs can be described by their pharmacology (Fig. 2) and physiology (Fig. 3). In terms of pharmacology, it is important to note that the endogenous catecholamines norepinephrine (NE) and epinephrine are agonists for all AR subtypes, although with varying affinity. βAR subtypes have different agonist potency series, enabling discrimination of subtypes. αARs have an identical agonist potency series; hence, it was not until selective antagonists (prazosin for α1ARs and idazoxan/yohimbine for α2ARs) were discovered that these two subtypes could be discriminated pharmacologically. βARs generally couple via Gs to stimulate adenylyl cyclase, with resultant increases in cyclic AMP (cAMP) in the cell. βARs also couple to K+ channels, possibly via Go, and the β2AR has been shown to be capable of coupling to Gi in cell culture and animal models Kompa et al., 1999, Xiao et al., 1999. Of note, β3ARs couple via Gs in fat cells, but couple via Gi in human heart (Gauthier et al., 1996). α2ARs, when located presynaptically (autoreceptors), inhibit NE release, and when located postsynaptically, mediate vaso- and venoconstriction. α2ARs are postsynaptic on platelets, the spinal cord, and in multiple sites in the CNS. α2ARs couple to Gi, which inhibits the enzyme adenylyl cyclase, causing decreases in cAMP in the cell (Smiley et al., 1998). α1ARs couple predominantly through Gq, resulting in hydrolysis of membrane phospholipids via phospholipase (PL)Cβ, to yield the second messengers inositol triphosphate (IP3) and diacylglycerol, leading to muscle contraction through mobilization of intracellular Ca2+ (Graham et al., 1996) and activation of protein kinase C (PKC). In addition to coupling to Gq, α1ARs have also been reported to activate pertussis-sensitive G-proteins. In myocytes, this leads to increased inotropy, [Na+-K+]- ATPase activation, modulation of intracellular Ca2+ levels, and cell shortening Mizobe et al., 1996, Steinberg et al., 1985, Terzic et al., 1993. Stimulation of α1ARs has been implicated in the pathogenesis of myocardial hypertrophy Simpson et al., 1982, Simpson & Savion, 1982 and BPH Hieble & Ruffolo, 1996, Schwinn & Price, 1999. Finally, recent evidence suggests that another G-protein, Gh [also known as transglutaminase type II (TGII)], has been shown to mediate α1AR stimulation of PLCδ1, increasing inositol phosphate turnover in TGII-transfected cells Chen et al., 1996, Feng et al., 1999, Feng et al., 1996, Park et al., 1998. TGII is a unique bifunctional protein that can (1) act as a transglutaminase that catalyzes Ca2+-dependent post-translational modification of proteins through formation of isopeptide bonds between glutamine and lysine residues (Folk, 1980) and (2) bind guanine nucleotides in a 1:1 ratio and hydrolyze GTP (Lee et al., 1989). Its role as a G-protein remains controversial, however.

ARs are members of the much larger family of G-protein-coupled receptors (GPCRs), which include, for example, muscarinic cholinergic receptors, serotonin receptors, dopamine receptors, neurokinin receptors, as well as the photoreceptor rhodopsin. Overall, membrane topology of ARs (and GPCRs) includes seven hydrophobic α-helical membranes called transmembrane regions (TM1–7), an extracellular amino-terminus, three extracellular loops, an intracellular carboxyl (COOH)-terminus, three main intracellular loops (IC1–3), plus a fourth IC loop created by palmitoylation of a COOH-terminal cysteine residue Mizobe et al., 1996, Schwinn et al., 1991a, Schwinn et al., 1991b. TM amino acid (aa) identity between AR subtype families (e.g., α vs. β) is ≈45%; this rises to 75% between subtypes within a family (e.g., α1a vs. α1b). Species homologues have even higher TM aa homology, often >90%, although key mutations can have functional consequences (e.g., human and rat α2a have different pharmacology due to one critical aa difference). Genomic organization is conserved within the different subtypes, suggesting that they evolved through gene duplication events Graham et al., 1996, Guarino et al., 1996.

Knockout and transgenic mice have provided researchers with valuable information on the effect of a single gene when deleted or overexpressed, respectively. However, a few caveats need to be kept in mind when discussing results from knockout/transgenic animal experiments. First, since other genes may compensate for the targeted gene, the resultant knockout phenotype may not represent the physiology of the gene of interest. Sometimes, no change in phenotype is seen in transgenic animals by knocking out individual genes due to the functional redundancy of closely related family members. Second, any variance in genetic background of animals used for transgenic/knockout experiments can significantly confound interpretation of the resulting phenotypes. Third, in potentially lethal deletions, surviving animals may represent a phenotype (not directly related to the knockout) that facilitated their survival. In spite of these drawbacks, genetically engineered mice have proved to be powerful tools, occasionally resulting in identification of previously unrecognized functions of specific gene products. Transgenic and knockout mice have been made for most AR subtypes Rohrer et al., 1998, Rohrer & Kobilka, 1998. Results from these studies are summarized in Table 1.

The cDNAs encoding three α1AR subtypes (α1a, α1b, and α1d) have been cloned and pharmacologically characterized Cotecchia et al., 1988, Lomasney et al., 1991, Perez et al., 1991, Schwinn et al., 1995, Schwinn et al., 1990. [Of note, α1AR nomenclature has changed within the last 5 years, with the International Union of Pharmacology adopting α1a, α1b, and α1d in 1995 (Hieble et al., 1995); the α1a used to be called α1c, the α1b name has not changed, and the α1d used to be called α1a, α1d, or α1a/d.] Although a fourth α1AR subtype (termed α1LAR, due to its comparatively lower affinity for the α1AR antagonist prazosin) has been described, recent evidence suggests that this subtype represents the low-affinity state of the α1aAR, and not a distinct receptor (Ford et al., 1997). At first glance, the physiologic role for individual AR subtypes appears to have been eloquently dissected using transgenic/knockout mice (Table 1). However, in the process of cloning cDNAs encoding each of the nine AR subtypes, tissue distribution studies in several animal species resulted in the surprising finding that tissue distribution of specific AR subtypes is species dependent. One of the first sets of experiments that noted this was performed in our laboratory, and they demonstrated that the α1AR subtype present in the liver depends on the species examined (the α1bAR is the only α1AR subtype in rat liver, while the α1aAR predominates in human liver; Price et al., 1993, Price et al., 1994a, Price et al., 1994b). Our laboratory extended these studies to demonstrate extensive species differences in all AR subfamilies (α1, α2, βARs), with αARs>βARs Berkowitz et al., 1994, Berkowitz et al., 1995, Price et al., 1993, Price et al., 1994a, Price et al., 1994b. Table 2 gives examples of α1AR subtype expression in a few rat and human tissues Malloy et al., 1998, Price et al., 1993, Price et al., 1994a, Price et al., 1994b. While many tissue-expression studies initially utilized RNA approaches (RNase protection assays, in situ hybridization), most of these findings have since been confirmed at a protein level (ligand binding and contraction studies), with generally good correlation between RNA and protein Malloy et al., 1998, Rudner et al., 1999.

In many tissues, quantification and localization of α1ARs has been difficult to accurately determine, due to limitations in reagents and the high degree of conservation of the α1AR subtypes. Recent methodologic advances seem to be providing more efficient and accurate means for characterizing receptor expression and pharmacology. Daly et al. (1998) recently described the use of a fluorescent quinazoline antagonist with a high affinity for α1ARs in a simplified model system of rat-1 fibroblasts stably expressing the α1dAR. This hopefully will lead to subtype-specific markers with distinct fluorescent tags to enable discrimination of α1AR subtypes within the same cell population. Subtype-specific antibodies are also being used to quantify cell-surface receptor expression, particularly in the cardiovascular system, where α1ARs are known to mediate many sympathetic responses, such as smooth muscle contraction Hrometz et al., 1999, Villalobos-Molina & Ibarra, 1996, cardiac contractility (Graham et al., 1996), cardiac hypertrophy Simpson et al., 1982, Simpson & Savion, 1982, and hypertension (Veelken & Schmieder, 1996). Hrometz et al. (1999) utilized an antibody approach to examine the correlation between function and α1AR subtype distribution in rat vasculature. Results showed that although all three subtypes were expressed in rat vasculature, it appears that only α1dAR is able to mediate smooth muscle contraction in the femoral artery, whereas α1aAR mediates renal artery contraction. It is important to note that results generated with α1aAR- and α1dAR-selective antibodies in cultured smooth muscle cells (SMCs) were not as clear as the α1bAR results, and further work is necessary to provide unequivocal data using antibody-based approaches. In a human model system, Ricci et al. (1999) used a combination of subtype-selective antibodies and reverse transcriptase-polymerase chain reaction (RT-PCR) (to analyze message levels) to demonstrate that all three subtypes are expressed in peripheral blood lymphocytes, with α1bAR levels predominating. As new reagents become more specific and reliable, we will be able to better ascertain the correlation between α1AR subtype expression and functionality in a given tissue. At a minimum, however, these studies should provide valuable insight into how α1AR subtype expression may be linked to pathologies and, thus, may provide useful markers for cardiovascular disorders.

Given the fact that catecholamine levels increase with age and the putative involvement of α1ARs in age-related pathologies, such as myocardial hypertrophy and BPH, a number of studies have focused on α1AR expression and signaling with age. Our laboratory recently examined human vascular α1AR subtype expression with age, and demonstrated artery-specific increases in α1AR density (Rudner et al., 1999). Xu et al. (1997) recently showed an age-dependent decrease in mRNA for all α1AR subtypes in rat aorta, but no change in mesenteric or pulmonary arteries. Thus, α1AR mRNA levels in the vasculature appear to vary with age in a vessel- and species-specific manner. In the heart, studies suggest that α1AR signaling decreases with age (del Balzo et al., 1990). Thus, a number of groups have sought to determine the molecular basis for signal dampening. Gascon et al. (1993) have shown through ligand-binding studies with the α1AR antagonist prazosin an overall reduction of α1ARs in the rat heart with age, although the equilibrium between high- and low-affinity sites is maintained. Miller et al. (1996) subsequently examined steady-state α1AR mRNA levels in the four chambers of the rat heart, and reported no decrease in α1AR mRNA levels with age, suggesting that dampening of α1AR signaling is post-transcriptional. This is in contrast to a previous study noting decreases in both α1AR mRNA and protein levels with age in the heart (Kimball et al., 1991). Taken together, the above studies suggest that age-associated dampening of α1AR signaling in the heart is mediated by nontranscriptional mechanisms, such as receptor synthesis or degradation, internalization, or effector coupling. Conversely, Simpson and colleagues demonstrated a significant increase in all three α1AR subtypes in adult (10 month) isolated rat cardiomyocytes versus 1- to 2-day-old neonatal isolated rat cardiomyocytes (Stewart et al., 1994). In this same study, cardiac fibroblasts were found to be devoid of expression of all three subtypes, demonstrating potentially important developmental as well as tissue-specific expression of the α1ARs in the heart. The potential ramifications of developmental increases in α1AR cardiomyocyte expression levels remain to be explored.

The direct role of α1ARs in mediating cardiac hypertrophy has been elegantly demonstrated in transgenic mouse models (the reader is referred to a recent review of GPCRs in the heart (Brodde & Michel, 1999). Lefkowitz and colleagues introduced a constitutively active mutant (CAM) of α1bAR into mouse hearts, resulting in myocardial hypertrophy in the absence of an increase in blood pressure (Milano et al., 1994). This provided conclusive evidence that α1AR stimulation can result in myocardial hypertrophy in the absence of increased afterload. However, it is important to note that transgenic hearts expressing the wild-type α1bAR at levels 40-fold higher than normal did not develop hypertrophy, despite enhanced Gq signaling [measured by an 8-fold increase in atrial natriuretic factor (ANF) ventricular mRNA] (Akhter et al., 1997; reviewed in Dorn & Brown, 1999). Probing potential mechanisms for altered myocardial function revealed that dual coupling of the α1bAR exists, as pertussis-toxin-sensitive G-protein-mediated attenuation of adenylyl cyclase was seen in myocardial membranes purified from wild-type α1bAR-expressing animals (Akhter et al., 1997). Additionally, it is not clear if the effect by CAM is mediated directly by α1bARs or indirectly by PLC stimulation, leading to, among other things, induction of α1aAR in the heart. Consistent with the latter hypothesis, responsiveness of cardiac cells to α1aAR stimulation persists, whereas desensitization and down-regulation of most other GPCRs is observed (including the α1b and α1dAR subtypes). This is an intriguing example of differential regulation between receptor family subtypes, with important implications regarding the role of α1aARs in chronically stimulated muscular hypertrophy. Both the Simpson and Woodcock laboratories showed that rat neonatal cardiomyocytes chronically stimulated with either NE or the α1aAR-selective agonist A-61603 resulted in induction of α1aAR message levels, with concurrent α1bAR and α1dAR down-regulation Autelitano & Woodcock, 1998, Rokosh et al., 1996. Taken together with studies performed in our laboratory showing that α1aAR predominates in the human heart, whereas α1aAR and α1bAR subtypes predominate in the rat heart Price et al., 1994a, Price et al., 1994b, these data suggest a key role for α1ARs (particularly α1aARs) in mediating myocardial hypertrophy. Great care must be taken, however, when applying results from different species to humans, as receptor-expression levels are species-dependent, with rat cardiac α1-adrenoceptors roughly 8 times the level in human heart (reviewed in Brodde & Michel, 1999).

Section snippets

Ligand binding

α1ARs, like other members of the GPCR protein family, are single polypeptide chains, ranging from 355 to 511 as that fold into a highly conserved structure of 7 predicted membrane-spanning regions, TM1–7. The seven TM regions fold into a conformation that creates a hydrophilic ligand-binding pocket surrounded by a hydrophobic core for the physiologic agonists epinephrine and NE (reviewed in Piascik et al., 1996). Although there is little aa homology among the different GPCRs, the predicted

Regulation of receptor signaling

While the precise signaling mechanisms through which α1AR-stimulated transcriptional activation occurs are not known, multiple pathways have been implicated. Recent experiments involving transgenic animals overexpressing a Gq inhibitor (thus, a functional Gq knockout) demonstrated a 60–70% reduction in myocardial hypertrophy, suggesting that 30–40% of hypertrophic pathways are PLC/PKC independent (Akhter et al., 1998). From recent studies aimed at elucidating the subcellular mechanisms of α1AR

Clinical aspects of α1-adrenergic receptor pharmacology

α1ARs have wide-ranging physiologic roles in health and disease (Table 4). Although present in diverse organs and tissues throughout the body, α1ARs are of particular importance in the cardiovascular system. In the myocardium, α1ARs mediate myocardial inotropy and hypertrophy, as well as play a role in atrial/ventricular arrhythmias and ischemic preconditioning Anyukhovsky et al., 1994, Anyukhovsky et al., 1997, Li et al., 1997, Tomai et al., 1999. Since α1ARs mediate smooth muscle contraction,

Summary

Recent advances in our understanding of the molecular mechanisms governing α1AR function have been described in this review. A tremendous acceleration in the amount of information regarding the structure and function of α1AR has been realized and will provide the framework for future drug design and potential gene therapy approaches. Elucidation of the precise structure of the ligand-binding pocket for each α1AR subtype will aid in the development and characterization of α1AR subtype-specific

Acknowledgements

The authors would like to recognize support from the following National Institutes of Health grants: AG-00745, AG-13853, HL-49103, and GCRC759 (to D.A.S.).

References (242)

  • S. Chen et al.

    Alpha1-adrenergic receptor signaling via Gh is subtype specific and independent of its transglutaminase activity

    J Biol Chem

    (1996)
  • M.M. Christensen et al.

    Clinical manifestations of benign prostatic hyperplasia and indications for therapeutic intervention

    Urol Clin North Am

    (1990)
  • P.G. De Benedetti et al.

    Alpha 1-adrenoceptor subtype selectivity: molecular modelling and theoretical quantitative structure–affinity relationships

    Bioorg Med Chem

    (1997)
  • N.L. Diehl et al.

    Identification of the α1D-adrenoceptor sybtype in rabbit arteries and the human saphenous vein using the polymerase chain reaction

    Eur J Pharmacol

    (1994)
  • D. Diviani et al.

    Effect of different G protein-coupled receptor kinases on phosphorylation and desensitization of the alpha1B-adrenergic receptor

    J Biol Chem

    (1996)
  • D. Diviani et al.

    Characterization of the phosphorylation sites involved in G protein-coupled receptor kinase- and protein kinase C-mediated desensitization of the alpha1B-adrenergic receptor

    J Biol Chem

    (1997)
  • G.W. Dorn et al.

    Gq signaling in cardiac adaptation and maladaptation

    Trends Cardiovasc Med

    (1999)
  • J.F. Feng et al.

    Evidence that phospholipase delta1 is the effector in the Gh (transglutaminase II)-mediated signaling

    J Biol Chem

    (1996)
  • S.S. Ferguson et al.

    Molecular mechanisms of G protein-coupled receptor desensitization and resensitization

    Life Sci

    (1998)
  • B. Gao et al.

    Isolation and characterization of the gene encoding the rat alpha 1B adrenergic receptor

    Gene

    (1993)
  • B. Gao et al.

    Transcription of the rat alpha 1B adrenergic receptor gene in liver is controlled by three promoters

    J Biol Chem

    (1994)
  • B. Gao et al.

    Cell type-specific transcriptional activation and suppression of the alpha1B adrenergic receptor gene middle promoter by nuclear factor 1

    J Biol Chem

    (1998)
  • B. Gao et al.

    The rat alpha 1B adrenergic receptor gene middle promoter contains multiple binding sites for sequence-specific proteins including a novel ubiquitous transcription factor

    J Biol Chem

    (1995)
  • R. Gaspar et al.

    Evidence of non-synaptic regulation of postpartum uterine contractility in the rat

    Life Sci

    (1998)
  • R.D. Guarino et al.

    Recent advances in the molecular pharmacology of the alpha 1-adrenergic receptors

    Cell Signal

    (1996)
  • H. Gurdal et al.

    α1-Adrenoceptor responsiveness in the aging aorta

    Eur J Pharmacol

    (1995)
  • A.N. Harris et al.

    A novel A/T-rich element mediates ANF gene expression during cardiac myocyte hypertrophy

    J Mol Cell Cardiol

    (1997)
  • B.E. Hawes et al.

    Distinct pathways of Gi- and Gq-mediated mitogen-activated protein kinase activation

    J Biol Chem

    (1995)
  • J.P. Hieble et al.

    The use of alpha-adrenoceptor antagonists in the pharmacological management of benign prostatic hypertrophy: an overview

    Pharmacol Res

    (1996)
  • A. Hirasawa et al.

    Cloning, functional expression and tissue distribution of human alpha 1c-adrenoceptor splice variants

    FEBS Lett

    (1995)
  • M. Hoshijima et al.

    The low molecular weight GTPase Rho regulates myofibril formation and organization in neonatal rat ventricular myocytes. Involvement of Rho kinase

    J Biol Chem

    (1998)
  • Z.W. Hu et al.

    Alpha 1 adrenergic receptors activate phosphatidylinositol 3-kinase in human vascular smooth muscle cells. Role in mitogenesis

    J Biol Chem

    (1996)
  • J. Hwa et al.

    Identification of critical determinants of alpha 1-adrenergic receptor subtype selective agonist binding

    J Biol Chem

    (1995)
  • R. Janknecht et al.

    Activation of the Sap-1a transcription factor by the c-Jun N-terminal kinase (JNK) mitogen-activated protein kinase

    J Biol Chem

    (1997)
  • P. Abrams

    Benign prostatic hyperplasia poorly correlated with symptoms

    Br Med J

    (1993)
  • J.W. Adams et al.

    Enhanced Gaq signaling: a common pathway mediates cardiac hypertrophy and apoptotic heart failure

    Proc Natl Acad Sci USA

    (1998)
  • R. Ahlquist

    Study of adrenotropic receptors

    Am J Physiol

    (1948)
  • S.A. Akhter et al.

    Targeting the receptor–Gq interface to inhibit in vivo pressure overload myocardial hypertrophy

    Science

    (1998)
  • J.T. Andersen et al.

    Prostatism: the correlation between symptoms, cystometric, and urodynamic findings

    Scand J Urol Nephrol

    (1979)
  • E.P. Anyukhovsky et al.

    Receptor-effector coupling pathway for alpha 1-adrenergic modulation of abnormal automaticity in ‘ischemic’ canine Purkinje fibers

    Circ Res

    (1994)
  • E.P. Anyukhovsky et al.

    Responses to norepinephrine of normal and ‘ischemic’ canine Purkinje fibers are consistent with activation of different alpha 1-receptor subtypes

    J Cardiovasc Electrophysiol

    (1997)
  • A. Ardati et al.

    A nuclear pathway for α1-adrenergic receptor signaling in cardiac cells

    EMBO J

    (1993)
  • D. Berkowitz et al.

    Localization of mRNA for three distinct α2-adrenergic receptor subtypes in human tissues: evidence for species heterogeneity and implications for human pharmacology

    Anesthesiology

    (1994)
  • G.E. Billman

    Effect of alpha 1-adrenergic receptor antagonists on susceptibility to malignant arrhythmias: protection from ventricular fibrillation

    J Cardiovasc Pharmacol

    (1994)
  • M.O. Boluyt et al.

    Rapamycin inhibits alpha 1-adrenergic receptor-stimulated cardiac myocyte hypertrophy but not activation of hypertrophy-associated genes. Evidence for involvement of p70 S6 kinase

    Circ Res

    (1997)
  • J. Booth et al.

    Acute depression of myocardial β-adrenergic receptor signaling during cardiopulmonary bypass: impairment of the adenylyl cyclase moiety

    Anesthesiology

    (1998)
  • S.E. Borst et al.

    Alpha 1-adrenergic and arginine vasopressin stimulation of inositide hydrolysis in rat hepatocytes is unaltered in senescence

    J Gerontol

    (1994)
  • M.R. Bristow

    Changes in myocardial and vascular receptors in heart failure

    J Am Coll Cardiol

    (1993)
  • M.R. Bristow et al.

    Changes in the receptor-G protein-adenylyl cyclase system in heart failure from various types of heart muscle disease

    Basic Res Cardiol

    (1992)
  • O.E. Brodde et al.

    Adrenergic and muscarinic receptors in the human heart

    Pharmacol Rev

    (1999)
  • Cited by (216)

    • Peripheral Alpha Blockers

      2023, Hypertension: A Companion to Braunwald's Heart Disease
    • Chiral analogues of (+)-cyclazosin as potent α<inf>1B</inf>-adrenoceptor selective antagonist

      2018, Bioorganic and Medicinal Chemistry
      Citation Excerpt :

      Beyond vascular contractile activity,1 which is more important in peripheral vasculature of older individuals,6 studies with α1B transgenic mice suggested that the α1B-adrenoceptor is involved in cardiac hypertrophy and neurodegeneration.7,8 Also the α1D-adrenoceptor is found in the vascular tissues,2 but it is predominantly expressed in bladder detrusor, where its stimulation causes bladder instability and irritability.3,9 However, the functional role of all three α1-adrenoceptors subtypes is not completely defined because of the lack or shortage of highly selective ligands, useful to elucidate their physiological activities and also to allow the treatment of the diseases in which they are involved.

    • Use of in vitro models in drug development and issue resolution

      2018, Advanced Issue Resolution in Safety Pharmacology
    View all citing articles on Scopus
    View full text