• No results found

Isovaleric acidemia: an integrated approach toward predictive laboratory medicine - Chapter 6: A novel UPLC-MS/MS based method to determine the activity of N-acetylglutamate synthase in liver tissue

N/A
N/A
Protected

Academic year: 2021

Share "Isovaleric acidemia: an integrated approach toward predictive laboratory medicine - Chapter 6: A novel UPLC-MS/MS based method to determine the activity of N-acetylglutamate synthase in liver tissue"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

Isovaleric acidemia: an integrated approach toward predictive laboratory

medicine

Dercksen, M.

Publication date

2014

Link to publication

Citation for published version (APA):

Dercksen, M. (2014). Isovaleric acidemia: an integrated approach toward predictive laboratory

medicine.

General rights

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s)

and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open

content license (like Creative Commons).

Disclaimer/Complaints regulations

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please

let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material

inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter

to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You

will be contacted as soon as possible.

(2)

Chapter 6

A novel UPLC-MS/MS based method to determine

the activity of N-acetylglutamate

synthase in liver tissue

Marli Dercksen 1,2

; Marinus Duran1

; Lodewyk IJlst1

; Wim Kulik1

;J.P. Ruiter 1

Arno van Cruchten1 and Ronald J.A Wanders1

Laboratory Genetic Metabolic Diseases, Departments of Pediatrics and Laboratory Medicine, Academic Medical Center, University of Amsterdam, The Netherlands1; Centre for Human Metabonomics, North-West University (Potchefstroom Campus), South Africa 2 In preparation for submission

(3)

Abstract

N-acetylglutamate synthase (NAGS) plays a key role in the removal of ammonia via the urea cycle by catalyzing the synthesis of N-acetylglutamate (NAG), the obligatory cofactor in the carbamyl phosphate synthetase reaction. The enzyme assay of NAGS in homogenates has remained problematic, which prompted us to develop a novel method. The assay conditions were optimized using purified human NAGS and the method was applied to mouse liver homogenates. UPLC-MS/MS in synergy with a stable isotope (N-acetylglutamic-2,3,3,4,4-d5 acid) for the quantitative detection of NAG was applied. Initially we observed a low signal-to-noise ratio in liver tissue samples which obstructed quantitative analyses. This was partly due to the low activity of NAGS in total liver homogenates, but more importantly due to non-enzymatic formation of NAG. Therefore we adapted the assay by quenching acetyl-CoA immediately after the incubation. With this novel approach we successfully determined NAGS activity in mouse liver homogenates and were able to differentiate between NAGS activity in wild type and knockout NAGS mice. The improved method can be used for the diagnosis of human inherited NAGS deficiency and may also be useful in the study of secondary hyperammonemia present in various inborn errors of metabolism as well as drug treatment.

(4)

Introduction

N-acetylglutamate synthase (NAGS, E.C:2.3.1.1) utilizes acetyl-CoA and L-glutamate for the formation of N-acetylglutamate (NAG), which is an obligatory allosteric activator of carbamyl phosphate synthetase 1 (CPS) and is therefore indispensable for proper urea formation. Inherited NAGS deficiency (OMIM: 237310) is an autosomal recessive disorder which is primarily associated with hyperammonemia (Bachmann et al., 1982b). Secondary hyperammonemia may occur in the so-called organic acidemias and upon intake of valproate, which both result in inhibition of NAGS and subsequently the dysfunction of CPS (Coudé et al., 1979; Coudé et al., 1982). Primary as well as induced deficiency of NAGS is associated with various symptoms including lethargy, vomiting, hypotonia, seizures, and eventually coma or death (Bachmann et al., 1982a). Primary NAGS deficiency is currently confirmed by mutational analyses of the NAGS gene. Confirmatory studies at the enzyme level have thus far been problematic due to the requirement of a relatively large amount of liver tissue as well as insufficient sensitivity of the measurement of the product, i.e. NAG (Morizono et al., 2004). It has been shown that patients with NAGS deficiency, as confirmed by pathogenic mutations in the NAGS gene, do not always exhibit a consistently deficient NAGS activity, upon in vitro enzyme analysis in liver biopsies, for reasons yet unknown (Morizono et al., 2004; Caldovic et al., 2003). Consequently, this disease may remain underdiagnosed due to the lack of specific biomarkers other than ammonia and the limited availability of centers able to measure NAGS and CPS activity.

Various assays for hepatic and purified NAGS have been developed, which include the use of radioactive substrates, and either HPLC, GC-MS, or LC-MS/MS for substrate and product analysis (Coudé et al., 1979; Lund and Wiggens 1984; Alonso and Rubio 1985; Tuchman and Holtzknecht 1990a; Tuchman and Holtzknecht 1990b). Initial methods used acetyl-CoA and [14C]-L-glutamate as substrate for NAGS, forming N-acetyl-[14C]-L-glutamate ([14C]-NAG). Subsequently, the production of [14C]-NAG was determined indirectly by the measurement of CPS activity via liquid scintillation counting (Coudé et al., 1982). Lund and Wiggins (1984) as well as Alonso and Rubio (1985) suggested the determination of NAGS via hydrolysis of hepatic NAG followed by measurement of release L-glutamate via HPLC. All described methods proved to be time-consuming and non-specific. A GC-MS stable isotope application was later developed to directly determine enzymatic NAG production, but column separation and absolute quantification (due to multiple peak formation) were problematic rendering it unsuitable for large-scale kinetic experiments (Tuchman and Holtzknecht, 1990a; Tuchman and Holtzknecht, 1990b).

(5)

The most promising methodology appears to be HPLC- or UPLC-MS/MS, in which direct measurement of NAGS and absolute quantification is possible with NAG stable isotope used as an internal standard (Caldovic et al., 2002). Specific reaction conditions have been investigated by several research groups which have shown that optimal substrate concentrations, pH, and the addition of L-arginine (which stimulates the activity of NAGS) are needed for accurately determining NAGS activity. Although the latter assay has proven to be successful, it suffers from high backgrounds in samples with low enzyme activity due to the non-enzymatic (chemical) formation of NAG during in vitro NAGS assays (Coudé et al., 1979; Coudé et al., 1982; Caldovic et al., 2003; Lund and Wiggens 1984; Alonso and Rubio 1985; Tuchman and Holtzknecht 1990a)

The aim of our study was to develop an optimized assay for the determination of NAGS activity in liver homogenates. To this end, we first determined the optimal reaction conditions and kinetic constants using purified human NAGS. The optimized method was subsequently used to measure the activity of NAGS in mouse liver homogenates, but we observed low undesirable signal-to-noise ratios. To circumvent this problem, we have now devised a novel method. The results are described in this paper.

Materials and Methods Materials

All materials were supplied by Sigma-Aldrich (St. Louis, MO) unless otherwise specified.

Methods

Purified NAGS enzyme assay

Human NAGS was expressed and purified as described before (Dercksen et al., 2013). The activity of NAGS, as expressed and purified from E.coli, was measured essentially as described by Caldovic et al. (2002) and Aires et al. (2011) with modifications. These modifications comprise an optimized sample preparation protocol and UPLC-MS/MS detection of the product (NAG). Several conditions of the assay, including time of reaction, pH optimum (7-10.5) and buffer systems (sodium phosphate, triethanolamine, Tris and ammonium acetate) were evaluated. Subsequently, the Km values for acetyl-CoA and L-glutamate were determined with

GraphPad Prism version 5 (GraphPad Software Inc., La Jolla, CA, USA), which led to an optimized 75 µl reaction mixture containing 50 mM ammonium acetate buffer

(6)

(final pH = 9.0), 10 mM L-glutamate, 2.5 mM acetyl-CoA and 1 mM L-arginine. The reaction was started by adding 0.05 µg purified enzyme (prepared in 0.1 mg/ml BSA/PBS solution). After a 15 min incubation period at 30 ºC, reactions were stopped with 175 μl acetonitrile (70% solution) and 10 μl of a solution containing the 1 mM internal standard, i.e. N-acetyl-L-glutamic-2,3,3,4,4-d5 acid (CDN isotopes: Germany), was added. All steps, except the incubation, were done on ice to prevent non-enzymatic formation of NAG. The precipitated protein was removed by centrifugation (20 000 g, 5 minutes, 4 ºC) and the supernatant was dried under nitrogen at 40 ºC. The reaction product was resuspended in 110 μl of 0.4% heptafluorobutyric acid (HFBA) solution (pH = 4).

The amount of NAG formed was quantified using a Quattro Premier XE tandem mass spectrometer (MS/MS) from Waters (Milford, MA, USA) with an Acquity sample manager and an Acquity binary solvent manager. NAG and NAG-[2,3,3,4,4-d5] were analyzed on a Waters C18-BEH column (100 mm-length x 2.1 mm-diameter, 1.7 µm particle size), using a linear gradient from 100 % solvent A (0.1 % HFBA) to 50 % solvent B (acetonitrile/water, 4/1, v/v) in 5 min. The flow rate was 500 µl/min with a total run time of 9 min. Both compounds were detected and quantified by MRM acquisition electron positive ionization mode (ESI+), using the transitions m/z 190 > 84 for NAG, and 195 > 88 for NAG-[2,3,3,4,4-d5]. The MS conditions furthermore consisted of a capillary voltage of 3.00 kV, cone voltage of 30 V, desolvation gas flow of 900 L/h as well as a source temperature and desolvation temperature of 120 °C and 300 °C, respectively. The optimized collision energy was achieved at 25 eV.

NAGS assay in mouse liver homogenate

Wild type and NAGS deficient mouse liver samples were kindly supplied by Dr. M. Tuchman and colleagues (Senkevitch et al., 2012). The activity of NAGS in mouse liver homogenates was measured using the same optimized conditions as described for purified human NAGS, with a final protein concentration of 2 mg/ml and one additional modification. At the end of the incubation (75 µl final volume), we quenched acetyl-CoA by adding 10.2 µl of a carnitine acetyltransferase (CRAT) (Roche Holding AG, Basel: Switzerland), L-carnitine and N-ethylmaleimide (NEM) solution, with a final concentration of 0.9 mU/ml, and 5.9 mM (for the latter two) respectively. This procedure allowed rapid conversion of all acetyl-CoA into acetylcarnitine, thereby avoiding further non-enzymatic formation of NAG.

(7)

Results and discussion

Optimization of the NAGS assay using purified NAGS

In order to determine the optimal assay conditions of purified human NAGS, we first expressed human NAGS in a His-tagged form in E.coli (Dercksen et al., 2013). We subsequently incubated the enzyme for 15 min at different pH values in the presence of 2.5 mM acetyl-CoA and 10 mM L-glutamate with the addition of L-arginine (1 mM). The results depicted in fig. 1 show that the enzyme activity reaches a maximum at a pH of 9.0. To further determine the optimal conditions of the assay, we performed a time dependence study as illustrated in fig. 2. The activity increased with time but the slope declined considerably after 20 min. Based on these results, we selected an optimal incubation time of 15 min. Under such conditions approximately 10% of the limiting substrate (acetyl-CoA) was used by NAGS to produce NAG.

Fig 1: The effect of the pH on the activity of NAGS. The pH of the standard reaction mixture

was adjusted with ammonium hydroxide accordingly. Purified NAGS was incubated for 15 min and the production of NAG was quantified as described in the methods section. NAGS activity was measured in duplicate, with each point representing the mean.

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 6.5 7 7.5 8 8.5 9 9.5 10 10.5 N A G S a c ti v it y ( µ m o l/ m in .m g ) Final pH

(8)

Fig. 2: The effect of the incubation time on production of NAG. The assay was performed as

described in the methods section using purified NAGS and the reaction was allowed to proceed for 5, 10, 20, 40 and 60 min after which the reactions were terminated with 70% acetonitrile. Values are the mean of duplicate incubations.

As reported previously (Tuchman and Holtzknecht, 1990a; Tuchman and Holtzknecht, 1990b), we detected a relatively high rate of N-acetylglutamate production in the absence of NAGS enzyme. First we studied the contribution of the type of buffer, used in the reaction mixture, on the non-enzymatic production of NAG. The results depicted in Fig. 3 show that the rate of non-enzymatic NAG-formation was lowest with ammonium acetate used as buffer and highest with triethanolamine. On the other hand, the rate of enzyme-driven NAG-formation was highest with triethanolamine followed by ammonium acetate, sodium phosphate and Tris. We obtained the best signal-to-noise ratio using ammonium acetate as buffer, and used this system in all subsequent experiments.

0 1 2 3 4 5 6 7 8 9 0 20 40 60 N A G S a c ti v it y ( µ m o l/ m in .m g ) Time (minutes)

(9)

Fig. 3: Influence of the type of buffer on the enzymatic and non-enzymatic production of NAG.

The formation of NAG from acetyl-CoA and L-glutamate was studied in a standard reaction medium containing each of the different buffers at a final concentration of 50 mM and a final pH of 9.0 in the absence (white bars) or presence (grey bars) of purified NAGS. The production of NAG was quantified as described in the material and methods section. The activity obtained with the original buffer conditions (Tris) as described by Aires et al. (2011), was set to 100%. The experiment was performed in duplicate and error bars indicate the range.

Finally we optimized the deproteinization step of the samples which must be effective in termination of the reaction as well as be compatible with an UPLC-MS/MS analysis. The use of perchloric acid/potassium carbonate (PCA/K2CO3) and

trichloroacetic acid/potassium hydroxide (TCA/KOH) resulted in severe ion suppression as well as inconsistent peak formation. Also, inappropriate peak sharpening was observed with the use of PCA/K2CO3 and peak broadening was

seen with the use of TCA/KOH, which complicated absolute quantitation. Acetonitrile or heat inactivation appeared to be better choices for the deproteinization of samples. Acetonitrile, which is easily removed by evaporation, was selected for all subsequent experiments. Furthermore, the resuspension of the sample in 0.4% HFBA, after evaporation under nitrogen (40ºC), was essential to avoid peak "splitting" and consequently absolute quantification (Piraud et al., 2005).

(10)

Kinetic parameters of purified NAGS

Fig 4: Influence of the L-glutamate and acetyl-CoA concentration on the activity of NAGS. The

activity was measured using purified NAGS in a standard reaction mixture containing (A) different concentrations of L-glutamate with fixed acetyl-CoA concentration (2.5 mM) or (B) different concentrations of acetyl-CoA with fixed L-glutamate concentration (10 mM). Michaelis-Menten plots were derived from non-linear regression analyses. The data points represent the mean values of three experiments.

Using the optimal assay conditions, as described for purified human NAGS, we varied the substrate concentrations in the incubation mixture and determined the activity of purified human NAGS, respectively. Fig. 4 depicts Michaelis-Menten plots for the two substrates which were used for the calculation of the kinetic parameters by non-linear regression analysis. The Km values for L-glutamate and acetyl-CoA

(11)

were 1.43 ± 0.28 mM and 0.53 ± 0.04 mM, respectively. The values concur with previous results of Caldovic et al. (2006) for purified human NAGS.

Wild type and knock-out NAGS activity in mouse liver

We tested the feasibility of the new assay for the determination of the activity of NAGS in mouse liver homogenates. The fractional activity ratio which is the ratio between the rates of formation of NAG in the presence and absence of homogenate, turned out to be very low compared to the formation of NAG with purified enzyme. This was partly due to the low activity of NAGS in total liver homogenates but more importantly due to the non-enzymatic formation of NAG. We reasoned that it would be best if we only had acetyl-CoA and L-glutamate present during the incubation time and to quench acetyl-CoA and/or L-glutamate immediately after the reaction period. To this end acetyl-CoA was rapidly depleted from the mixture by adding a solution of carnitine acetyltransferase (CRAT), L-carnitine plus N-ethylmaleimide (NEM). The latter compound was added, since it reacts with SH- groups thereby trapping the coenzyme A (CoA) produced from acetyl-CoA and shifting the equilibrium towards acetylcarnitine. With this modification of the protocol, we observed much higher fractional activity ratios. The limit of detection (LOD) and quantitation (LOQ) were determined as 0.06 nmol/min.mg and 0.19 nmol/min.mg, respectively.

In order to provide final proof that our newly developed assay was functioning well, we determined the activity of NAGS in murine liver homogenates in which the gene coding for NAGS had been disrupted. The NAGS activity in the liver homogenate from the NAGS-/- mice (n=4) was below the LOD and LOQ values. Liver samples of wild type mice (n=2) showed NAGS activity of 1.55 and 1.71 nmol/min.mg, respectively which is almost 10 times the LOQ value. An additional advantage of our method was the use of low protein concentrations (2 mg/ml), compared to earlier studies which used 10-20 mg/ml cell lysate, which mostly consisted of enriched mitochondrial fractions (Caldovic et al., 2006; Aires et al., 2011).

Conclusions

We have developed an improved assay for the enzymatic determination of NAGS based on the use of UPLC-MS/MS for the detection of NAG. An important advantage of the current assay is the optimization of conditions for mass spectrometric analyses, which included the use of an appropriate buffer and the termination of the enzyme assay by adding a 70% acetonitrile solution. Furthermore, the non-enzymatic formation of NAG by rapidly converting acetyl-CoA to acetylcarnitine at the end of the incubation was maximally suppressed. The current method is of value for the enzymatic confirmation of NAGS deficiency and the study of patients with

(12)

unexplained hyperammonemia (Tuchman and Holzknecht 1990a; Caldovich et al., 2002; Morizono et al., 2004). In addition, the drug-induced inhibition of NAGS and other forms of secondary hyperammonemia can be explored with this assay.

Acknowledgements

The study was financially supported as a Carolina MacGillavry PhD Fellowship awarded by "Koninklijke Nederlandse Akademie van Wetenschappen". We thank the staff of the enzyme and mass spectrometry sections of the Laboratory Genetic Metabolic Diseases (GMD) at the Academic Medical Centre, University of Amsterdam, for their expert advice and assistance. We are also specifically grateful for the supply of the liver material by Dr M. Tuchman and Dr E. Senkevitch, Children's National Medical Center, Washington DC, USA. A special thanks go out to Tom Wagemans, from lab GMD, for the production of purified NAGS for the in vitro studies.

References

Aires CPC, Van Cruchten A, IJlst, L, et al. 2011. New insight on the mechanism of valproate-induced hyperammonemia: inhibition of hepatic N-acetyl-glutamate synthase activity by valproyl-CoA. J Hepatol 55:426-434.

Alonso E, Rubio V. 1985. Determination of N-acetyl-L-glutamate using high-performance liquid chromatography. Anal Biochem 146:252-259.

Bachmann C, Colombo JP, Jaggi K. 1982a. N-acetylglutamate synthetase (NAGS)

deficiency: diagnosis, clinical observation and treatment. Adv Exp Med Biol 153:39-45.

Bachmann C, Krahenbuhl S, Colombo JP. 1982b. Purification and properties of

acetyl-CoA: L-glutamate N-acetyltransferase from human liver. Biochem J 205:123-127.

Caldovic L, Lopez GY, Haskins N, et al. 2006. Biochemical properties of recombinant human and mouse N-acetylglutamate synthase. Mol Genet Metab 87:226-232.

Caldovic L, Morizono H, Panglao MG et al. 2003. Null mutation in the N-acetylglutamate synthase gene associated with neonatal disease and hyperammonemia. Hum Genet 112:364-368.

Caldovic L, Morizono H, Pangloa MG et al. 2002. Cloning and expression of the human N-acetylglutamate synthase gene. Biochem Biophys Res Commun 299:581-586.

(13)

Coudé FX, Rabier D, Cathelineau L, Grimber G, Parvy P, Kamoun P. 1982. A mechanism for valproate-induced hyperammonemia. Adv Exp Med Biol 153:153-156.

Coudé FX, Sweetman L, Nyhan WL. 1979. Inhibition by propionyl-CoA of N-acetylglutamate synthetase in rat liver mitochondria. J Clin Invest 64:1544-1551.

Dercksen M, IJlst L, Duran M et al. 2013. Inhibition of N-acetylglutamate synthase by various monocarboxylic and dicarboxylic short-chain coenzyme A esters and the production of alternative glutamate esters. Biochem Biophys Acta. doi.org/10.1016/j.bbadis.2013.04.027.

Lund P, Wiggins D. 1984. Is N-acetylglutamate a short-term regulator of urea synthesis? Biochem J 218:991-994.

Morizono H, Caldovic L, Shi D, Tuchman M. 2004. Mammalian N-acetylglutamate synthase. Mol Genet Metab 81:S4-S11.

Piraud M, Vianey-Saban C, Petritis K, Elfakir C, Steghens J, Bouchu D. 2005. Ion-pairing reversed-phase liquid chromatography/electrospray ionization mass spectrometric analysis of 76 underivatized amino acids of biological interest: a new tool for the diagnosis of inherited disorders of amino acid metabolism. Rapid Commun Mass Spectrom 19:1587-1602.

Senkevitch E, Cabrera-Luque J, Morizono H, Caldovic L, Tuchman M. 2012. A novel biochemically salvageable animal model of hyperammonemia devoid of N-acetylglutamate synthase. Mol Genet Metab 106:160-168.

Tuchman M, Holzknecht RA. 1990a. Human hepatic N-acetylglutamate content and

N-acetylglutamate synthase activity. Biochem J 271:325-329.

Tuchman M, Holzknecht RA. 1990b. N-acetylglutamate in liver and gut of normal and

fasted mice, normal human livers and livers of individuals with carbamyl phosphate synthetase or ornithine transcarbamylase deficiency. Pediatr Res 27:408-412.

Referenties

GERELATEERDE DOCUMENTEN

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons.. In case of

Inn order to evaluate that an ultra-short loop feedback, mechanism controlling TSH secretion at thee pituitary level is responsible for the suppression of TSH levels in

Wee synthesised thyroidal cDNA that was used to produce a) two DIG-labelled TSHR cDNA probess to screen a human pituitary library, b) sense and antisense DIG-labelled TSHR RNA

goitree growth, low thyroxine and increasing triiodothyronine during PTU treatment.. Wenzel KW, Lente JR. 1983 Syndrome of persisting thyroid stimulating immunoglobulins and growth

Inn conclusion, long-term TSH suppression in patients with Graves' hyperthyroidism rendered euthyroidd with antithyroid drugs is correlated quantitatively with the presence of TBII

geeft het elektriciteitsgebruik en dieselverbruik per 1.000 kg melk weer van bedrijven die opstallen, gekeken naar de totale bedrijfsproductie.. Er is geen verband

De verschillende reacties van gsOVCs en wsOVCs op (1) blootstelling aan ratmyeline, (2) BAM- en microglia-geconditioneerd medium, en (3) de verschillende effecten van