Original research
The Study of Carbamoyl Phosphate Synthetase 1 Deficiency Sheds Light on the Mechanism for Switching On/Off the Urea Cycle

https://doi.org/10.1016/j.jgg.2015.03.009Get rights and content

Abstract

Carbamoyl phosphate synthetase 1 (CPS1) deficiency (CPS1D) is an inborn error of the urea cycle having autosomal (2q34) recessive inheritance that can cause hyperammonemia and neonatal death or mental retardation. We analyzed the effects on CPS1 activity, kinetic parameters and enzyme stability of missense mutations reported in patients with CPS1 deficiency that map in the 20-kDa C-terminal domain of the enzyme. This domain turns on or off the enzyme depending on whether the essential allosteric activator of CPS1, N-acetyl-L-glutamate (NAG), is bound or is not bound to it. To carry out the present studies, we exploited a novel system that allows the expression in vitro and the purification of human CPS1, thus permitting site-directed mutagenesis. These studies have clarified disease causation by individual mutations, identifying functionally important residues, and revealing that a number of mutations decrease the affinity of the enzyme for NAG. Patients with NAG affinity-decreasing mutations might benefit from NAG site saturation therapy with N-carbamyl-L-glutamate (a registered drug, the analog of NAG). Our results, together with additional present and prior site-directed mutagenesis data for other residues mapping in this domain, suggest an NAG-triggered conformational change in the β4-α4 loop of the C-terminal domain of this enzyme. This change might be an early event in the NAG activation process. Molecular dynamics simulations that were restrained according to the observed effects of the mutations are consistent with this hypothesis, providing further backing for this structurally plausible signaling mechanism by which NAG could trigger urea cycle activation via CPS1.

Introduction

Carbamoyl phosphate synthetase 1 (CPS1) deficiency (CPS1D, OMIM #237300), a recessively inherited autosomal (2q34) (McReynolds et al., 1981) inborn error of the urea cycle (Freeman et al., 1964, Gelehrter and Snodgrass, 1974), has an estimated incidence of 1/50000 to 1/300000 (Uchino et al., 1998, Summar et al., 2013). CPS1 is the entry point of ammonia, the nitrogenous waste product of protein catabolism, into the urea cycle (Fig. 1A). Therefore, CPS1D causes pure hyperammonemia (Häberle and Rubio, 2014), leading to encephalopathy and even death (Brusilow and Horwich, 2001), and to depletion of downstream urea cycle intermediates, particularly of citrulline (Häberle and Rubio, 2014).

A large repertory of mutations affecting the CPS1 gene (OMIM #608307; 201,425 nucleotides; start/end chromosome 2 coordinates, 211,342,405/211,543,830, plus strand; http://www.genecards.org/cgi-bin/carddisp.pl?gene=CPS1) has been compiled from patients with CPS1D (Häberle et al., 2011). Over 50% of these mutations are missense changes spreading over the entire 1462-residue mature CPS1 polypeptide (Nyunoya et al., 1985, Haraguchi et al., 1991). The CPS1 gene encompasses 4500 coding nucleotides over 38 exons (Funghini et al., 2003, Häberle et al., 2003, Summar et al., 2003). It may be difficult to ascertain the responsibility of a given CPS1 missense mutation in causing CPS1D, particularly for mutations mapping outside the two catalytic domains of the enzyme (the two phosphorylation domains, Fig. 1B) which bind the substrates and catalyze the three-step CPS1 reaction (Alonso et al., 1992, Alonso and Rubio, 1995).

Our present work dealt with the analysis of the effects of CPS1D-associated mutations (called here clinical mutations) that affect a non-catalytic domain of human CPS1, the C-terminal domain of 20 kDa (Häberle et al., 2011). This domain is called the allosteric domain (abbreviated ASD) (Fig. 1B) because it binds N-acetyl-L-glutamate (NAG) (Rodriguez-Aparicio et al., 1989, Pekkala et al., 2009), the essential allosteric activator of CPS1. Without NAG, CPS1 is inactive (Rubio et al., 1981, Rubio et al., 1983), possibly reflecting the need to stop catalysis by the enzyme (Shigesada et al., 1978, Stewart and Walser, 1980) before the ammonia level is too low (Fig. 1A). Too much decrease in the ammonia level would lead to depletion of ammonia-derived amino acids such as glycine, glutamate and glutamine (Bender, 2012), and possibly to protein catabolism. NAG is a proper effector of the CPS1 switch because its level reflects the nitrogen burden manifested in the glutamate level (Fig. 1A). This is so because NAG has a short half-life (Morita et al., 1982) and it is made from glutamate by an enzyme (NAG synthase) exhibiting a high Km for glutamate (Sonoda and Tatibana, 1983).

To explore the effects of ASD-mapping missense clinical mutations (abbreviated ASD clinical mutations), we utilized a novel system for production and mutagenesis of recombinant CPS1 that uses baculovirus and insect cells (Díez-Fernández et al., 2013). We had already applied this system to analyze the effects of some ASD missense mutations (Pekkala et al., 2010, Díez-Fernández et al., 2013). We now extend this analysis to all the reported ASD clinical mutations, as well as to some mutations designed to test the role of the ASD (Fig. 1B, vertical lines, and Table 1, Table 2). Analysis of the effects of these mutations has helped assess disease causality, opening the way to improved genetic counselling and even to individualized therapy. Furthermore, we now shed some light on the as yet unclarified NAG activation process, and define better the NAG site. We had previously localized the NAG site (Fig. 1C) (Pekkala et al., 2009) by photoaffinity labeling and by in silico docking in the deposited (but unpublished) crystal structure of the isolated human ASD free from NAG [Protein Databank (PDB; www.rcsb.org) file 2YVQ; Xie et al., 2007]. We now have found that some ASD mutations have effects that are inconsistent with the previously proposed NAG site. With the help of molecular dynamics (MD), applying restraints based on the site-directed mutagenesis results, we now propose a refined NAG site structure where a conformational change in the β4-α4 loop could be the initial signal in NAG activation.

Section snippets

Clinical ASD domain mutations

The twelve reported (Häberle et al., 2011) CPS1D-associated missense mutations mapping in the ASD are listed in Table 1. All of them affect residues that are invariant or highly conserved in NAG-sensitive CPSs. Except for two mutations (R1371L and Y1491H), they were given unanimous predictions of being likely to have negative effects by two widely used pathogenicity prediction servers, Polyphen-2 and MutPred (Table 1) (Li et al., 2009, Adzhubei et al., 2010).

The clinical mutations highlighted

Discussion

The present and earlier experimental results (Pekkala et al., 2010, Díez-Fernández et al., 2013, Díez-Fernández et al., 2014) corroborate the value of the baculovirus/insect cell system used here for assessing the effects of CPS1D missense mutations. In the case of the ASD, only for one of the twelve reported clinical mutations (Häberle et al., 2011), P1411L, the expression studies could not ascertain the disease causality of the mutation (Table 1) (Pekkala et al., 2010). Three ASD clinical

Human CPS1 production

Pure recombinant human mature liver CPS1 with an N-terminal His6-tag, and the desired site-directed mutants of this enzyme were produced from the pFastBac-CPS1 vector as previously described (Díez-Fernández et al., 2013). The same purification procedure (including cell centrifugation, lysis, centrifugal clarification, Ni-affinity chromatography and centrifugal ultrafiltrative concentration) proved appropriate for wild type and mutant enzyme forms. Purity was monitored by SDS-PAGE (8%

Acknowledgments

This work was supported by grants from the Alicia Koplowitz Foundation, the Valencian government (No. PrometeoII/2014/029 to V.R.), the Spanish government (Nos. BFU2011-30407 to V.R., BFU2012-30770 to J.G. and a FPU fellowship to C.D.-F.), and the Swiss National Science Foundation (No. 310030_127184 to J.H.). We thank Belén Barcelona (IBV-CSIC, Valencia, Spain) for help with the structural model of the complete enzyme and Fig. 4D. The original mutation analysis of some CPS1D patients was kindly

References (46)

  • V. Rubio et al.

    Carbamoyl phosphate synthetase I of human liver. Purification, some properties and immunological cross-reactivity with the rat liver enzyme

    Biochim. Biophys. Acta

    (1981)
  • T. Sonoda et al.

    Purification of N-acetyl-L-glutamate synthetase from rat liver mitochondria and substrate and activator specificity of the enzyme

    J. Biol. Chem.

    (1983)
  • P.M. Stewart et al.

    Short term regulation of ureagenesis

    J. Biol. Chem.

    (1980)
  • M.L. Summar et al.

    Characterization of genomic structure and polymorphisms in the human carbamyl phosphate synthetase I gene

    Gene

    (2003)
  • M.L. Summar et al.

    The incidence of urea cycle disorders

    Mol. Genet. Metab.

    (2013)
  • I.A. Adzhubei et al.

    A method and server for predicting damaging missense mutations

    Nat. Methods

    (2010)
  • E. Alonso et al.

    Affinity cleavage of carbamoyl-phosphate synthetase I localizes regions of the enzyme interacting with the molecule of ATP that phosphorylates carbamate

    Eur. J. Biochem.

    (1995)
  • D.A. Bender

    Amino Acid Metabolism

    (2012)
  • S.W. Brusilow et al.

    Urea cycle enzymes

  • D.A. Case et al.

    The Amber biomolecular simulation programs

    J. Computat. Chem.

    (2005)
  • C. Díez-Fernández et al.

    Molecular characterization of carbamoyl-phosphate synthetase (CPS1) deficiency using human recombinant CPS1 as a key tool

    Hum. Mutat.

    (2013)
  • S. Funghini et al.

    Structural organization of the human carbamyl phosphate synthetase I gene (CPS1) and identification of two novel genetic lesions

    Hum. Mutat.

    (2003)
  • T.D. Gelehrter et al.

    Lethal neonatal deficiency of carbamyl phosphate synthetase

    N. Engl. J. Med.

    (1974)
  • Cited by (24)

    • Unraveling the therapeutic potential of carbamoyl phosphate synthetase 1 (CPS1) in human diseases

      2023, Bioorganic Chemistry
      Citation Excerpt :

      NAG (N-acetyl-L-glutamate), the essential allosteric activator of CPS1, when binding to the allosteric domain of CPS1, triggers conformational changes of it. NAG binding leads to a dramatic remodeling that stabilizes the catalytically competent conformation, and the building of the ∼ 35 Å-long tunnel that allows migration of the intermediate from one active site of formation to another active site[12–14]. Furthermore, insight into the structure of the NAG binding site revealed that an αβα open sandwich, whose central parallel sheet of five strands covered by three α-helices on one side and two α-helices on the other side, together forming a Rossmann fold (Fig. 2B)[3,15].

    • Clinical and structural insights into potential dominant negative triggers of proximal urea cycle disorders

      2021, Biochimie
      Citation Excerpt :

      Protein was concentrated and stored in 10 mM HEPES pH 7.5, 500 mM NaCl, and 5% glycerol at −80 °C. Overexpression and purification of rhCPS1 was carried out as previously described [55–57]. Briefly, protein expression and production were mediated by the CPS1 cDNA-carrying bacmid, using Bac-to-Bac® Baculovirus Expression System (Invitrogen) in Sf9 insect cells.

    • CPS1 T1405N polymorphism, HDL cholesterol, homocysteine and renal function are risk factors of VPA induced hyperammonemia among epilepsy patients

      2019, Epilepsy Research
      Citation Excerpt :

      HA is a metabolic disturbance featured by an excess of ammonia in the peripheral plasma, which may lead to the injury of brain and even death (Hernandez-Rabaza et al., 2016; Jayakumar and Norenberg, 2018). The hepatic urea cycle plays a significant role in removing ammonia from the body, in which carbamoyl-phosphate synthase 1 (CPS1) catalyzes the very first step of converting the ammonia and bicarbonate into carbamoyl phosphate (Dimski, 1994; Diez-Fernandez et al., 2015). Genetic mutations in CPS1 gene have been reported to cause CPS1 deficiency (CPS1D, OMIM #237300) which featured HA and “valproate sensitivity”.

    • Farnesoid X receptor: A “homeostat” for hepatic nutrient metabolism

      2018, Biochimica et Biophysica Acta - Molecular Basis of Disease
      Citation Excerpt :

      The urea cycle relies on the mitochondrial enzymes carbamoyl-phosphate synthase I (CPS1), ornithine transcarbamylase (OTC), and the cytoplasmic enzymes argininosuccinate synthase (ASS1), argininosuccinate lyase (ASL), and arginase (ARG1). Urea cycle progression requires synthesis of N-acetylglutamate (NAG) by NAG synthase (NAGS) and allosteric binding of NAG to CPS1 [153]. Other than allosteric cofactors, regulation of urea cycle enzymes involves substrate availability, hormonal and transcriptional regulation, and PTMs [154,155].

    • Precision medicine in rare disease: Mechanisms of disparate effects of N-carbamyl-L-glutamate on mutant CPS1 enzymes

      2017, Molecular Genetics and Metabolism
      Citation Excerpt :

      The newly developed baculovirus/insect cell expression system, which can produce a large amount of the recombinant CPS1 with desired mutations, provides a useful tool to evaluate the residual activity, protein yield and stability of CPS1 mutants. This tool has been successfully used to study the disease-causing roles and pathogenic mechanism of various mutations in CPS1 [36,38,51–54]. Even though the concentration of NCG in liver mitochondria of the subjects receiving 100 mg/kg/d of the drug is unknown, there seemed sufficient NCG to activate CPS1 and restore ureagenesis in subjects with negligible NAG.

    View all citing articles on Scopus
    View full text