Structural features and regulatory properties of the brain glutamate decarboxylases

https://doi.org/10.1016/S0197-0186(00)00014-0Get rights and content

Abstract

It is widely recognized that the two major forms of GAD present in adult vertebrate brains are each composed of two major sequence domains that differ in size and degree of similarity. The amino-terminal domain is smaller and shows little sequence identity between the two forms. This domain is thought to mediate the subcellular targeting of the two GADs. Substantial parts of the amino-terminal domain appear to be exposed and flexible, as shown by proteolysis experiments and the locations of posttranslational modifications. The carboxyl-terminal sequence domain contains the catalytic site and shows substantial sequence similarity between the forms. The interaction of GAD with its cofactor, pyridoxal-5′ phosphate (pyridoxal-P), plays a key role in the regulation of GAD activity. Although GAD65 and GAD67 interact differently with pyridoxal-P, their cofactor-binding sites contain the same set of nine putative cofactor-binding residues and have the same basic structural fold. Thus the cofactor-binding differences cannot be attributed to fundamental structural differences between the GADs but must result from subtle modifications of the basic cofactor-binding fold. The presence of another conserved motif suggests that the carboxyl-terminal domain is composed of two functional domains: the cofactor-binding domain and a small domain that closes when the substrate binds. Finally, GAD is a dimeric enzyme and conserved features of GADs superfamily of pyridoxal-P proteins indicate the dimer-forming interactions are mediated mainly by the carboxyl-terminal domain.

Introduction

GABA and glutamate decarboxylase (GAD) were discovered in brain almost 50 years ago (Awapara et al., 1950, Roberts and Frankel, 1950a, Roberts and Frankel, 1950b, Udenfriend, 1950, Wingo and Awapara, 1950), at a time when the prolonged debate over the very existence of chemical synaptic transmission had just concluded (Bacq, 1975). At that time, of course, the function of brain GABA was yet to be established. GABA is now known best as the predominant inhibitory transmitter in brain, but very substantial fractions of GABA and GAD are found far from synaptic endings in cell bodies and dendrites, and the idea that GABA has important nonsynaptic functions cannot be dismissed (reviewed by Martin and Rimvall, 1993, Soghomonian and Martin, 1998). The discovery that vertebrates have two genes for GAD and express two major forms of the enzyme (Bu et al., 1992, Erlander et al., 1991) opened the possibility that each form is specialized to synthesize a specific pool of GABA that serves a specific function (Martin and Barke, 1998). In this paper we will summarize the functional differences and similarities between the two forms and examine aspects of the structures of the two forms that help us to understand their functions.

The two forms of GAD, which are called GAD65 and GAD67, are the products of two independently regulated genes (Bu et al., 1992, Erlander et al., 1991). The two forms are widely but not equally expressed among cell types in different brain regions (Esclapez et al., 1993, 1994; Feldblum et al., 1993, Mercugliano et al., 1992). GAD65 and GAD67 are almost equally abundant in several regions of mouse brain but GAD65 is much more abundant than GAD67 in most regions of rat brain (Sheikh et al., 1999).

Immunocytochemical and subcellular fractionation studies have shown that the two GADs are distributed differently within neurons (Erlander et al., 1991, Esclapez et al., 1994, Hendrickson et al., 1994, Henry and Tappaz, 1991, Reetz et al., 1991, Rimvall and Martin, 1994, Wood et al., 1976). GAD65 is more concentrated within terminals and appears to be associated with synaptic vesicles while GAD67 is more uniformly distributed throughout the neuron, being easily detected in cell bodies and dendrites. In brain, GAD65 and GAD67 are present both as soluble enzymes and associated with membranes. In addition, brain contains GAD65–GAD67 heteromers (Sheikh and Martin, 1996). The difference in subcellular targeting of the two enzymes, the membrane association of the enzymes and the formation of heteromers are important matters for which clues may be sought in the structures of the enzymes.

Both forms of GAD require the cofactor pyridoxal-P for activity, and the interaction of GAD with pyridoxal-P plays a central role in the regulation of GAD activity. At least half of brain GAD is present as apoenzyme (GAD without bound cofactor) which provides a reserve of inactive enzyme that can be activated when additional GABA synthesis is required (Itoh and Uchimura, 1981, Miller et al., 1977). Active-site labeling studies indicate that the high proportion of apoGAD in brain is unusual, if not unique, among pyridoxal-P dependent enzymes (Martin et al., 1991). The reserve function of apoGAD appears to operate in tissues, as depolarizing synaptosomes leads to the conversion of apoGAD to holoGAD and stimulates GABA synthesis (Gold et al., 1978, Gold and Roth, 1979, Miller et al., 1980). Finally, the interconversion of apo- and holoGAD is a complex, regulated process that involves a large conformational change in the enzyme (Chen et al., 1998, Martin and Rimvall, 1993).

GAD65 and GAD67 appear to differ in their interactions with pyridoxal-P, although these differences have not yet been studied in detail. Experiments with the individual forms of recombinant enzyme and with GAD65 and GAD67, prepared by immunoprecipitation from cerebellum, indicate that GAD65 is more responsive to added pyridoxal-P than is GAD67 (Erlander et al., 1991, Kaufman et al., 1991). In addition, GAD65 accounts for most of the apoGAD in synaptosomes and several regions of rat brain (Martin et al., 1991). At first glance the latter result seems consistent with the idea that GAD65 is more highly regulated by the cofactor than is GAD67. However, GAD65 is much more abundant than GAD67 in synaptosomes and most rat-brain regions, and the greater abundance of apoGAD65 appears to result in large part from the greater abundance of GAD65 (Rimvall and Martin, 1994, Sheikh et al., 1999).

The unusual and complex interaction of GAD with pyridoxal-P and the related differences between the two GADs suggest that these enzymes have interesting and unusual structural features.

Section snippets

Materials

Chemicals were obtained from Sigma (St Louis, MO, USA). Sequencing grade trypsin and endoproteinase Asp-N were obtained from Sigma and Boehringer Mannheim (Indianapolis, IN). Q-Sepharose was from Pharmacia Biotech (Piscataway, NJ, USA). Immobilon-P was from Millipore (Bedford, MA, USA). AntiGAD sera W887, W883, and W624 were prepared in this laboratory (Sheikh and Martin, 1996). W887 was prepared against a peptide corresponding to residues 570–585 of GAD65 and 578–593 of GAD67. This shared

Sequence domains of the GADs

GAD65 and GAD67 are each composed of two major sequence domains as revealed by pairwise sequence alignment (Erlander et al., 1991). One of these domains is composed of residues 1–95 in human GAD65 and 1–101 in GAD67 and is strikingly different in the two GADs, as it has only 23% sequence identity. This domain, called the N-terminal domain appears, to be responsible for subcellular targeting and the formation of GAD65–GAD67 heteromers (Dirkx et al., 1995, Sheikh and Martin, 1996, Shi et al., 1994

Acknowledgements

Fluorimaging, mass spectrometry and protein sequencing were conducted in the Wadsworth Center’s Molecular Immunology, Biological Mass Spectroscopy, and Protein Chemistry Core Facilities, respectively. We thank Charles R. Hauer for his invaluable advice and assistance with mass spectrometry. This work was supported by grant MH35664 from the National Institute of Mental Health, DHHS.

References (41)

  • E. Roberts et al.

    γ-Aminobutyric acid in brain: its formation from glutamic acid

    J. Biol. Chem.

    (1950)
  • J.-J. Soghomonian et al.

    Two isoforms of glutamate decarboxylase: why?

    Trends Pharmacol. Sci.

    (1998)
  • S. Udenfriend

    Identification of γ-aminobutyric acid in brain by the isotope derivative method

    J. Biol. Chem.

    (1950)
  • W.J. Wingo et al.

    Decarboxylation of l-glutamic acid by brain

    J. Biol. Chem.

    (1950)
  • Z.M. Bacq
  • D.-F. Bu et al.

    Two human glutamate decarboxylases, 65-kDa GAD and 67-kDa GAD, are each encoded by a single gene

    Proc. Natl. Acad. Sci. USA

    (1992)
  • S. Christgau et al.

    Membrane anchoring of the autoantigen GAD65 to microvesicles in pancreatic β-cells by palmitoylation in the NH2-terminal domain

    J. Cell Biol.

    (1992)
  • M. Esclapez et al.

    Comparative localization of two forms of glutamic acid decarboxylase and their mRNAs in rat brain supports the concept of functional differences between the forms

    J. Neurosci.

    (1994)
  • M. Esclapez et al.

    Comparative localization of mRNAs encoding two forms of glutamic acid decarboxylase with nonradioactive in situ hybridization methods

    J. Comp. Neurol.

    (1993)
  • S. Feldblum et al.

    Different distributions of GAD65 and GAD67 mRNAs suggest that the two glutamate decarboxylases play different functional roles

    J. Neurosci. Res.

    (1993)
  • Cited by (31)

    • The blood supply of the cat's visual cortex and its postnatal development

      2004, Brain Research
      Citation Excerpt :

      Barone et al. [3] have noted that alkaline phosphatase is highly expressed in regions of neuropil where glutamic acid decarboxylase (GAD, the rate-limiting enzyme in the synthesis of GABA) is also highly expressed, for example, layer 4 of visual cortex [4,26]. They argue that alkaline phosphatase plays a critical role at GABAergic terminals, dephosphorylating pyridoxal phosphate, thereby allowing pyridoxal to be transported into the terminal, where it is rephosphorylated and acts as a co-factor for GAD65 [61]. This suggestion is consistent with earlier observations that alkaline phosphatase knockout mice exhibit seizures [65] that resemble those in mice fed diets deficient in vitamin B-6 and can be mitigated by administering pyridoxal HCL [66].

    View all citing articles on Scopus
    View full text