Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Extended diagnosis of purine and pyrimidine disorders from urine: LC MS/MS assay development and clinical validation

  • Péter Monostori ,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    monostoripeter@gmail.com

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Glynis Klinke,

    Roles Data curation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Jana Hauke,

    Roles Data curation, Formal analysis, Investigation, Methodology, Validation, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Sylvia Richter,

    Roles Data curation, Methodology, Validation, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Jörgen Bierau,

    Roles Methodology, Validation, Writing – review & editing

    Affiliation Department of Clinical Genetics, Maastricht University Medical Center, Maastricht, The Netherlands

  • Sven F. Garbade,

    Roles Data curation, Formal analysis, Methodology, Software, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Georg F. Hoffmann,

    Roles Conceptualization, Funding acquisition, Methodology, Project administration, Resources, Supervision, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Claus-Dieter Langhans,

    Roles Conceptualization, Funding acquisition, Methodology, Project administration, Resources, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Dorothea Haas ,

    Contributed equally to this work with: Dorothea Haas, Jürgen G. Okun

    Roles Conceptualization, Methodology, Project administration, Resources, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

  • Jürgen G. Okun

    Contributed equally to this work with: Dorothea Haas, Jürgen G. Okun

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Project administration, Resources, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of General Pediatrics, Division of Neuropediatrics and Metabolic Medicine, Center for Pediatric and Adolescent Medicine, University Hospital Heidelberg, Heidelberg, Germany

Abstract

Background and aims

Inborn errors of purine and pyrimidine metabolism are a diverse group of disorders with possible serious or life-threatening symptoms. They may be associated with neurological symptoms, renal stone disease or immunodeficiency. However, the clinical presentation can be nonspecific and mild so that a number of cases may be missed. Previously published assays lacked detection of certain diagnostically important biomarkers, including SAICAr, AICAr, beta-ureidoisobutyric acid, 2,8-dihydroxyadenine and orotidine, necessitating the use of separate assays for their detection. Moreover, the limited sensitivity for some analytes in earlier assays may have hampered the reliable detection of mild cases. Therefore, we aimed to develop a liquid chromatography–tandem mass spectrometry (LC-MS/MS) assay that allows the simultaneous and sensitive detection of an extended range of purine and pyrimidine biomarkers in urine.

Methods

The assay was developed and validated using LC-MS/MS and clinically tested by analyzing ERNDIM Diagnostic Proficiency Testing (DPT) samples and further specimens from patients with various purine and pyrimidine disorders.

Results

Reliable determination of 27 analytes including SAICAr, AICAr, beta-ureidoisobutyric acid, 2,8-dihydroxyadenine and orotidine was achieved in urine following a simple sample preparation. The method clearly distinguished pathological and normal samples and differentiated between purine and pyrimidine defects in all clinical specimens.

Conclusions

A LC-MS/MS assay allowing the simultaneous, sensitive and reliable diagnosis of an extended range of purine and pyrimidine disorders has been developed. The validated method has successfully been tested using ERNDIM Diagnostic Proficiency Testing (DPT) samples and further clinical specimens from patients with various purine and pyrimidine disorders. Sample preparation is simple and assay duration is short, facilitating an easier inclusion of the assay into the diagnostic procedures.

Introduction

Purine and pyrimidine bases, nucleosides and nucleotides are essential components of the nucleic acids DNA and RNA, and are associated with metabolic regulation, synthesis of numerous biomolecules and other vital processes in cell physiology [13] (S1 Fig). Accordingly, inborn errors of purine and pyrimidine metabolism can manifest with serious or life-threatening symptoms that may include immunological, hematological, neurological and renal pathology [13].

Purine and pyrimidine deficiencies are generally considered rare; however, their prevalence is unknown and probably underestimated [1, 2]. One reason is that their clinical spectrum is diverse and frequently overlapping with other disorders. Moreover, there is considerable heterogeneity in the phenotype and concentrations of biomarkers even within an affected family. Third, although previously considered pediatric diseases, these disorders are now increasingly being diagnosed in adults as milder presentations and mutations are recognized [1, 2]. This underlines the importance of analytical assays that can detect biomarkers of purine and pyrimidine metabolism with sufficient sensitivity and specificity [1, 2]. For clinical diagnosis, liquid chromatography–tandem mass spectrometry (LC-MS/MS) is the technique of choice [4], predominantly using urine as sample matrix [1, 2].

Despite the availability of clinical LC-MS/MS methods for purines and pyrimidines [59], there is still significant need for improvement. First, previously published assays lacked detection of certain diagnostically important biomarkers, including SAICAr (succinyl-5-aminoimidazole-4-carboxamide-1-ribonucleoside; a marker of adenylosuccinate lyase deficiency, ADSL), not included in [59]; AICAr (5-aminoimidazole-4-carboxamide ribonucleoside; a marker of AICAR transformylase/IMP cyclohydrolase deficiency, ATIC), not assayed in [59]; beta-ureidoisobutyric acid (beta-ureidopropionase deficiency, UPB1), not detected in [5, 7, 9]; 2,8-dihydroxyadenine (adenine phosphoribosyltransferase deficiency, APRT), not included in [5, 79] and orotidine (uridine monophosphate synthetase deficiency (UMPS) type III), not assayed in [59]. Thus, in order to avoid missing cases, separate assays such as the Bratton-Marshall-test for SAICAr [10] were recommended to be performed additionally for detection of these additional metabolites. In addition, some analytes could only be measured with limited sensitivity in those earlier LC-MS/MS assays [59], potentially preventing reliable detection of milder disease forms.

Therefore, we aimed to develop an LC-MS/MS assay that allows the reliable diagnosis of an extended range of purine and pyrimidine disorders while still maintaining sample preparation simple and assay duration short for easier inclusion of the assay in the routine workflow. To verify the clinical performance of the assay, we set out to analyze ERNDIM Diagnostic Proficiency Testing (DPT) samples and further specimens from known patients.

Materials and methods

Reagents

Unlabeled standards were purchased from Sigma-Aldrich (St. Louis, MO, USA), except for 2,8-dihydroxyadenine, dihydrothymine, orotidine and succinyladenosine (Toronto Research Chemicals, Toronto, ON, Canada) and pseudouridine (Berry & Associates, Dexter, MI, USA). SAICAr was obtained from Dr. Marie Zikánová (Charles University, Prague, Czech Republic). Deuterated internal standards (ISs) were acquired from Cambridge Isotope Laboratories (Andover, MA, USA). Methanol, isopropanol and formic acid (all ULC-MS grade) were purchased from Biosolve (Valkenswaard, The Netherlands). Ultrapure water (18.2 MΩ.cm), filtered through a 0.22 μm pore size membrane, was obtained from a Thermo Barnstead GenPure Pro UV-TOC/UF system (Thermo Fisher Scientific, Waltham, MA, USA).

Calibrators and quality controls

The analyte levels in calibrators Cal01 to Cal10 prior to sample preparation were as follows: 0, 1, 2, 4, 8, 16, 32, 64, 128 and 256 μM, respectively. During method validation and in subsequent analyses, we used the „Control Purines and Pyrimidines” material from ERNDIM (supplier: MCA Laboratory, Queen Beatrix Hospital, Winterswijk, The Netherlands) as an officially approved external quality control (EQC; the LOT 2015.010 was used for method development). In order to monitor further analytes not included in the EQC, we also prepared an internal QC (IQC) in-house that contained all metabolites of the EQC plus beta-ureidopropionic acid, allopurinol, beta-ureidoisobutyric acid, 2,8-dihydroxyadenine and succinyladenosine. Since analyte levels in the EQC represented the highly pathologic range, concentrations in the IQC were chosen to be lower but still analytically relevant, i.e. 20 μM. All calibrators, the IQC and EQC samples and the IS working solution were stored in aliquots at -20°C. The preparation of standard stock solutions, calibrators, IS stock and working solutions and QCs are further detailed in the S1 Text.

Sample preparation

Urinary creatinine concentrations were determined with a Beckman Coulter AU480 Chemistry System, using a kinetic modification of the Jaffe procedure (Beckman Coulter Diagnostics, Brea, CA, USA). Urine samples were centrifuged for 5 min at 400 × g in a Hettich Rotina 420R centrifuge (Tuttlingen, Germany) at ambient temperature. Sample preparation was performed in a 96-well Merck MultiScreen filter plate (pore size: 0.45 μm, volume: 2 ml), with a Merck MultiScreen collection plate (volume: 2 ml). In order to obtain creatinine levels of 0.25 mM in all samples at the first stage of sample preparation, appropriate volumes of urine samples (based on creatinine levels) were diluted to 400 μl with water, whereas 100 μl aliquots of the calibrators and QCs were mixed with 300 μl synthetic urine (creatinine 0.333 mM, Synthetic Urine e. K., Nußdorf, Germany). Then, 100 μl 25 μM IS was added to all samples. Accordingly, the final concentrations of creatinine and IS in the samples were 0.20 mM and 5 μM, respectively, with a final sample volume of 500 μl.

The samples were sealed with an adhesive foil (ChromSystems, Gräfelfing, Germany), mixed and incubated for 10 min at ambient temperature. Then, the samples were filtered using a 96-well Plate Vacuum Manifold (Phenomenex, Torrance, CA, USA). The collection plate was sealed with a pre-slit adhesive foil (ChromSystems, Gräfelfing, Germany) and placed directly into the autosampler. Results of the LC-MS/MS analysis were reported in mmol/mol creatinine.

LC-MS/MS instruments and settings

A Waters Acquity I-Class UPLC system (Binary Solvent Manager, thermostatic Column Manager and FTN Sample Manager) and a Waters TQ-S triple quadrupole MS/MS were used, controlled by MassLynx 4.1 software (all Waters, Milford, MA, USA). MS/MS data were evaluated with TargetLynx 4.1 software (Waters, Milford, MA, USA).

The chromatographic separation of the analytes was performed on an ACE Excel C18-AR 100×3.0mm; 1.7 μm analytical column (Advanced Chromatography Technologies, Aberdeen, Scotland). Eluent A consisted of ultrapure water plus 0.4% formic acid (ULC-MS grade). Eluent B consisted of methanol/water 50/50 (methanol ULC-MS grade). Both eluents were prepared weekly. The column was tempered to 25°C. The autosampler settings were the following: injection volume: 1 μl; sample compartment temperature: 5°C. Gradient elution at a flow of 200 μl/min was performed by changing %B as follows: 0.0-1.1 min: 10%; 1.1-6.0 min: 10% to 80%; 6.0-6.1 min: 80% to 100%; 6.1-8.0 min: 100%; 8.0-8.5 min: 100% to 10%; 8.5-14.0 min: 10%. The flow was directed to waste in the first 2 min and the last 5.5 min of each acquisition. Further experience with columns, eluents and centrifugal filter cups obtained during method development are provided in the S1 Table.

MS/MS settings were optimized by means of consecutive syringe infusions of 1 μM solutions for each analyte and IS, dissolved in methanol/water 50/50. Optimized MS/MS settings, including electrospray ion (ESI) modes of purines and pyrimidines and related metabolites are summarized in Tables 1 and 2, respectively. In the final MS/MS method, all MRMs were acquired within 3 min acquisition windows (retention time ±1.5 min), with dwell times of 5 ms for each MRM. The final ion source settings were the following: desolvation gas flow = 800 L/h; cone gas flow = 150 L/h; nebulizer = 6.0 bar; capillary voltage = 2.5 kV; desolvation temperature = 650°C; source temperature = 150°C.

thumbnail
Table 1. Optimized MS/MS settings of purines and related metabolites.

https://doi.org/10.1371/journal.pone.0212458.t001

thumbnail
Table 2. Optimized MS/MS settings of pyrimidines and related metabolites.

https://doi.org/10.1371/journal.pone.0212458.t002

Application of the LC-MS/MS assay on ERNDIM Diagnostic Proficiency Testing (DPT) samples and further patient specimens

Diagnostic testing of the LC-MS/MS assay was first performed by analyzing the levels of purines and pyrimidines in ERNDIM Diagnostic Proficiency Testing (DPT) samples (n = 10). Subsequently, further urine specimens from patients (n = 10) referred for routine purine-pyrimidine determination using a previously published method [6] were re-assayed with the present method. All procedures followed were in accordance with the ethical standards of the Helsinki Declaration of 1975, as revised in 2000 and approved by the Ethical Committee of the University of Heidelberg (071/2005). Written consent was obtained from ERNDIM as an approval to use the DPT samples in method development.

Results

Method validation

The linearity (coefficient of determination, R2), the limit of detection (LOD) and the lower and upper limits of quantitation (LLOQ, ULOQ) were tested after serial dilution and subsequent sample preparation of calibrators (n = 10) (S2 Table). LOD was defined as S/N = 3; LLOQ was defined as the lowest concentration measured with a coefficient of variation (CV) lower than 20%; ULOQ was defined as the upper end of the range where the response of a given analyte was linear, i.e. where the coefficient of determination (R2) was higher than 0.99.

The chromatographic resolution (Rs) of critical analyte pairs showing MS/MS interference was calculated via the following equation: Rs = 1.18(t2-t1)/(w0.5,1+w0.5,2) where t1 and t2 are the retention times of the respective peaks and w0.5,1 and w0.5,2 are the peak widths at half peak height [11]. Baseline resolution (Rs≥1.5) was achieved for all critical pairs (S2 Table).

Carryover was tested by injecting blank samples (methanol) after the highest calibrator Cal10. No peaks were detected in the blank samples.

Interday and intraday precisions, variations between multiple injections of a single sample extract and recoveries were examined using both the IQC and the EQC samples (n = 10) (S3 Table and Table 3, respectively). To highlight the reliability of the method, the interday reproducibility data were calculated and reported from 10 measurements on 10 consecutive weeks, not 10 consecutive days. Note, that the IQC sample included additional analytes, as compared with the ERNDIM EQC sample.

thumbnail
Table 3. Assay validation using the ERNDIM „Control Purines and Pyrimidines” external quality control samples (EQC; LOT used for method development: 2015.010) (n = 10).

https://doi.org/10.1371/journal.pone.0212458.t003

Unprepared calibrators and QCs were stable with CV<15% for all analytes for at least 6 months if stored aliquoted at -20°C.

The stability of the prepared samples during storage was evaluated by aliquoting a set of calibrators and QCs immediately after sample preparation, with one aliquot measured both immediately and following 48h of storage at 5°C in the autosampler and another aliquot assayed after 7 days of storage at -20°C. There were no significant differences in the peak intensities, the linear curve equations or the calculated concentrations of the QCs.

Reference intervals

For determination of normal biomarker ranges, urine samples (n = 251) referred for routine purine-pyrimidine measurements using a previously published method [6] were re-assayed with the present method. The specimens (collected between 0-18 years of age) were randomly chosen and anonymized prior to the analysis. Reference intervals and cutoffs were determined according to the following statistical procedure.

The R environment for statistical computing and graphics was used to analyze data [12]. Initially, for each metabolite, values above 99.5th percentile were excluded. The 97.5th percentile of the remaining data defined the upper cutoff of an analyte. For pseudouridine, a lower cutoff was also defined as the 2.5th percentile. The possible age dependency of the cutoffs was tested as follows. A one-way analysis of variance (ANOVA) based on the 97.5th percentile and, additionally for pseudouridine, on the 2.5th percentile was computed [13] to test if the cutoffs differed significantly between the age groups 0-1, 1-3, 3-6 and 6-18 years, respectively. Where quantile ANOVA detected significantly different 97.5th (or 2.5th) percentiles between certain age groups and metabolic specialists also rated the difference as clinically relevant, age-related cutoffs were defined for the respective groups. The cutoff values are provided in Table 4.

Testing of the present assay with ERNDIM Diagnostic Proficiency Testing (DPT) samples and further clinical specimens

The levels of purines and pyrimidines were measured in ERNDIM Diagnostic Proficiency Testing (DPT) samples (n = 10). The new assay clearly identified the pathological and the normal metabolite patterns and correctly differentiated between purine-pyrimidine disorders in all cases (Fig 1).

thumbnail
Fig 1. Results of the ERNDIM Diagnostic Proficiency Testing (DPT) samples (n = 10).

Numbers on the cylinders show how many times a given result exceeded the respective cutoff for that biomarker. PP: purine-pyrimidine disorder. Further abbreviations are provided in the List of Abbreviations Section. Note that orotidine is not pathognomonic for HGPRT (sample E04) but is accumulated as a result of allopurinol treatment.

https://doi.org/10.1371/journal.pone.0212458.g001

Results of the analyses of further patient specimens (n = 10) are depicted in Fig 2. The new LC-MS/MS assay clearly distinguished the pathological samples from normal ones and correctly diagnosed purine-pyrimidine disorders in all tested specimens. Selected clinical data of the patients are given in S2 Text.

thumbnail
Fig 2. Results of urine samples from patients with known diagnosis (n = 10).

Numbers on the cylinders show how many times a given result exceeded the respective cutoff for that biomarker. Abbreviations are provided in the List of Abbreviations Section. The Y-axis has been cut at 45 for a better visibility. Note that orotidine is not pathognomonic for HGPRT (sample P05) but is accumulated as a result of allopurinol treatment.

https://doi.org/10.1371/journal.pone.0212458.g002

To investigate the possible effects of urine pH on the LC-MS/MS results, we tested urine samples with very low creatinine levels, which allowed testing the highest possible amount of urine in the sample preparation. Three aliquots were prepared in each of two urine samples. 20 μl of 0.1M formic acid, 0.1M NaOH or water was added to the aliquots, respectively, resulting in a ±1 unit shift of the original urine pH (originally 6 and 7 for the two urine samples, respectively). There were no differences between the calculated concentrations of the three corresponding aliquots.

Representative LC-MS/MS chromatograms of a calibrator, a healthy proband and patients with ADSL, HGPRT, UPB1 and APRT deficiencies, respectively, are presented in Fig 3.

thumbnail
Fig 3.

LC-MS/MS chromatograms of A: the Cal06 calibrator (16 μM) and urine specimens from B: a healthy proband; C: a patient with ADSL deficiency; D: a patient with HGPRT deficiency; E: a patient with UPB1 deficiency; and F: a patient with APRT deficiency. All intensities (Y-axes) have been normalized to 1.5*108 cps for better comparability and have been plotted against retention time (X-axes). Analytes have been numbered as follows: 1. URA; 2. DHU; 3. Thy; 4. DHT; 5. BUP; 6. ADN; 7. Hypo; 8. ALLOP; 9. BUIB; 10. Xan; 11. DHA; 12. dAdo; 13. AICAr; 14. ADS; 15. HMU; 16. Oro; 17. dUrd; 18. THD; 19. PSU; 20. dIno; 21. dGui; 22. Ino; 23. Guo; 24. Ord; 25. SAICAr; 26. SAdo. The ISs have been marked by an asterisk (*) after the number of the respective unlabeled analyte. Note that guanosine-IS has been measured in both ESI+ and ESI-. ADSL: Adenylosuccinate lyase deficiency; HGPRT: Hypoxanthine-guanine phosphoribosyltransferase deficiency; UPB1: beta-Ureidopropionase deficiency; APRT: Adenine phosphoribosyltransferase deficiency. Further abbreviations are provided in the List of Abbreviations Section. Chromatograms were generated by MassLynx 4.1 (Waters, Milford, MA, USA) and modified using Inkscape 0.91pre4 (Open Source Software licensed under the GNU General Public License).

https://doi.org/10.1371/journal.pone.0212458.g003

Discussion

Inborn errors of purine and pyrimidine metabolism are characterized by a broad phenotypic spectrum, in line with the ubiquitous presence and biochemical importance of these metabolites [13]. Clinically, purine and pyrimidine deficiencies can present with serious immunological, hematological, neurological and renal symptoms including delayed development, epilepsy, muscular hyper- or hypotonicity, mental retardation, dysmorphic features, neurogastrointestinal symptoms, ophthalmoplegia, muscle atrophy and polyneuropathy [13]. However, the clinical presentation can be nonspecific and mild which may result in missing the diagnosis [1, 2].

Nucleotides within cells break down to form the respective nucleosides and bases, which are transported into the extracellular fluids [1]. Thus, the diagnosis of purine and pyrimidine disorders is generally performed by detecting the related nucleosides, bases or degradation products associated with the enzyme defect. Urinary excretion profiles are most commonly used but application of plasma, serum or cerebrospinal fluid have also been published [1, 2]. Only very few disorders were reported to be detectable from dried blood spots which could possibly allow their inclusion in newborn screening [1416].

Traditionally, common methods for the detection of purine and pyrimidine metabolites applied HPLC combined with diode array or UV detection or, for pyrimidine degradation products, GC-MS. 1H-NMR was also reported to be appropriate for the diagnosis of many but not all of these disorders [17], still the most applied technique currently is LC-MS/MS [4].

To the best of our knowledge, the LC-MS/MS assay developed and presented here is the first one published to include the biomarkers SAICAr, AICAr, beta-ureidoisobutyric acid, 2,8-dihydroxyadenine and orotidine in a single method, in addition to the previously reported panel of purines and pyrimidines. This extension is expected to facilitate the detection of the disorders ADSL [1, 3], ATIC [1, 3], UPB1 [2, 3], APRT [1, 3] and UMPS [2, 18, 19], respectively, together with the deficiencies of HGPRT [1, 3], molybdenum cofactor (MoCoD) [1, 3], thymidine phosphorylase (TP) [2, 3], dihydropyrimidine dehydrogenase (DPD) [2, 3] and dihydropyrimidinase (DHP) [2, 3]. Based on the analyte panel, the presented assay should additionally be able to detect deficiencies of adenosine deaminase (ADA) [1, 3], purine nucleoside phosphorylase (PNP) [1, 3] and xanthine dehydrogenase (XDH) [1, 3], as well as phosphoribosylpyrophosphate synthetase (PRPPS) superactivity [1, 3]. However, this should be confirmed using patient samples with these disorders, which were not available for the authors of the present paper.

Many previously published LC-MS/MS methods aimed to assay several metabolites simultaneously in order to be able to diagnose multiple disorders [59] but all of them lacked detection of one or other of the above biomarkers. Even in assays using emerging techniques for the detection of a broad range of analytes, such as LC-QTOF [2022], the number of currently diagnosable purine and pyrimidine disorders was substantially lower than in the present assay. As an example, the detection of SAICAr, a marker of ADSL deficiency, is commonly performed using the Bratton-Marshall-test [10]. However, this color reaction test is neither quantitative nor specific as it may give positive results with a number of structurally related purine intermediates [10]. Previously, only few assays were reported for the quantitative measurement of SAICAr [14; 23; 24] which, in turn, allowed the detection of a much lower number of additional analytes.

To improve the quantitation of purines and pyrimidines, we have used a large number of isotopically labeled ISs (total: 11) as compared with earlier reports [5, 6, 8, 9]. The use of multiple ISs has not only improved the reliability of quantitation for analytes having an own labeled analog, but also allowed us to select the most appropriate IS for those analytes where a labeled analog was not commercially available. The selection was based on numerous factors, including the long-term stability of calibration, linearity (absence of saturation) and similarities in structure and retention time [25, 26].

In line with the above approach, we have decided to omit the labeled analogs of adenine and adenosine following extensive testing, despite being commercially available. The exclusion of labeled adenine-IS and adenosine-IS eliminated crosstalk between these single labeled ISs and their unlabeled counterparts, having mass differences of only one Da [25] and also helped to decrease ion suppression for high analyte levels [26]. Instead, xanthine-IS and guanosine-IS, respectively, were found to perform well as ISs, exhibiting an improved stability of calibration and a wider linear range.

Note that even if 2,8-dihydroxyadenine had been included in the calibrators during assay validation, it was omitted from the standard row later. The variable responses of 2,8-dihydroxyadenine, observable exclusively among distinct charges of calibrators, were presumably related to the very poor solubility of 2,8-dihydroxyadenine [1]. In turn, the responses of 2,8-dihydroxyadenine in urine samples were stable under the current assay conditions. Thus, this issue could have been resolved by using the same calibration settings obtained during validation to quantify 2,8-dihydroxyadenine in subsequent sample preparations. In accordance, we could correctly identify the ERNDIM Diagnostic Proficiency Testing (DPT) sample E08 having elevated adenine and 2,8-dihydroxyadenine levels, characteristic for APRT deficiency (Fig 2). The same approach was used for the quantitation of SAICAr due to difficulties in the availability of the standard substance. In line, two patient specimens with ADSL deficiency (P09 and P10) were correctly identified (S3 Table).

Another limitation of the present assay is that the inclusion and quantitation of uric acid, the end-product of purine degradation, was not possible under the current sample preparation and LC conditions. In addition to the low temperature in the autosampler, otherwise favorable in terms of stability for most analytes, a low pH was also associated with a decreased solubility of uric acid [27]. On the other hand, acidic conditions were shown to be important not only for the adequate chromatographic separation of the analytes in the present assay but were also critical for the stability of dihydrouracil and dihydrothymine which would have been degraded to beta-ureidopropionic acid and beta-ureidoisobutyric acid under higher pH (see also S1 Text) [28]. Since uric acid can routinely be measured using traditional techniques such as colorimetric or enzymatic assays, we opted to maintain the pH in the acidic range throughout sample preparation and measurement in favor of all other analytes.

The use of a well-chosen column and dedicated method development allowed us to achieve ultra-high performance in terms of sensitivity, resolution and clinical reliability and to still remain in pressure ranges close to but not exceeding the limits of conventional high-performance liquid chromatography (HPLC). This approach, together with simple sample preparation, is expected to facilitate an easier inclusion of this assay in the daily routine and to improve transferability among laboratories. The ongoing prospective evaluation of the presented method in our laboratory and a wider range of patient samples can provide further experience on the reliability of the assay to identify patients in the routine setting.

Conclusions

An LC-MS/MS assay allowing the simultaneous, sensitive and reliable diagnosis of an extended range of purine and pyrimidine disorders has been developed. The validated assay has been tested using ERNDIM Diagnostic Proficiency Testing (DPT) samples and further clinical specimens from patients with various purine and pyrimidine disorders. The method clearly distinguished the pathological and normal samples and differentiated between purine and pyrimidine defects in all tested specimens. Sample preparation is simple and assay duration is short, facilitating an easier inclusion of the method in a routine workflow. The prospective evaluation of the described method is currently being continued in our laboratory.

Supporting information

S1 Fig. Biochemical pathways of purine and pyrimidine metabolism (simplified).

Metabolites detected by the presented method are printed in bold.

Purines:

Enzymes of the de novo purine synthesis: PRPPS: phosphoribosyl pyrophosphate synthetase; ADSL: adenylosuccinate lyase (adenylosuccinase); ATIC: AICAR transformylase/IMP cyclohydrolase; ADSS: adenylosuccinate synthetase. Enzymes of purine catabolism: AMPA: AMP deaminase; 5NT: 5'-nucleotidase(s); ADA: adenosine deaminase; PNP: purine nucleoside phosphorylase; XDH: xanthine dehydrogenase (xanthine oxidase); GDA: guanine deaminase; IMPDH: IMP dehydrogenase. Enzymes of purine salvage: HGPRT: hypoxanthine-guanine phosphoribosyltransferase; dGUOK: deoxyguanosine kinase; APRT: adenine phosphoribosyltransferase; ADK: adenosine kinase.

Pyrimidines:

Enzymes of the de novo pyrimidine synthesis: CPS2: carbamoylphosphate synthetase II: ATC: aspartate transcarbamoylase; DHO: dihydroorotase (CAD is comprised of CPS2, ATC and DHO); DHODH: dihydroorotate dehydrogenase; OPRT: orotate phosphoribosyltransferase; OMPDC: OMP decarboxylase (UMPS = uridine monophosphate synthetase is comprised of OPRT and OMPDC). Enzymes of pyrimidine catabolism: 5NT: 5'-nucleotidase(s); UP: uridine phosphorylase(s); TP: thymidine phosphorylase; DPD: dihydropyrimidine dehydrogenase; DHP: dihydropyrimidinase; UPB1: beta-ureidopropionase. Enzymes of pyrimidine salvage: PUS: pseudouridine synthase; TDO: thymine dioxygenase. Ribonucleotide reductase (RNR) and thymidylate synthetase (TYMS) are used in the synthesis of deoxynucleotides.

AMP: adenosine-5’-monophosphate; AICAr: 5-aminoimidazole-4-carboxamide ribonucleoside; AICAR: 5-aminoimidazole-4-carboxamide ribonucleotide; CMP: cytidine-5’-monophosphate; GMP: guanosine-5’-monophosphate; IMP: inosine-5’-monophosphate; OMP: orotidine-5’-monophosphate; PRPP: phosphoribosylpyrophosphate; SAICAr: succinyl-5-aminoimidazole-4-carboxamide-1-ribonucleoside; SAICAR: succinyl-5-aminoimidazole-4-carboxamide-1-ribonucleotide; S-AMP: adenylosuccinate; TMP: thymidine-5’-monophosphate; UMP: uridine-5’-monophosphate; XMP: xanthosine-5’-monophosphate.

Drawn using Inkscape 0.91pre4 (Open Source Software licensed under the GNU General Public License).

https://doi.org/10.1371/journal.pone.0212458.s001

(TIF)

S1 Text. Preparation of unlabeled standard and internal standard (IS) solutions, calibrators and Quality Controls (QCs).

https://doi.org/10.1371/journal.pone.0212458.s002

(DOC)

S2 Text. Selected clinical data of the examined patients with known diagnosis (n = 10).

https://doi.org/10.1371/journal.pone.0212458.s003

(DOC)

S1 Table. Extended details of method development.

https://doi.org/10.1371/journal.pone.0212458.s004

(DOC)

S2 Table. Assay validation using calibrators (n = 10) and chromatographic resolution of critical analyte pairs with mass spectrometric interference.

https://doi.org/10.1371/journal.pone.0212458.s005

(DOC)

S3 Table. Assay validation using the internal quality control samples (IQC) (n = 10).

https://doi.org/10.1371/journal.pone.0212458.s006

(DOC)

Acknowledgments

The authors acknowledge the use of data derived from ERNDIM External Quality Assurance (EQA) materials in this manuscript. The use of ERNDIM EQA materials does not imply that ERNDIM endorses the methods used or the scientific validity of the findings in this manuscript. ERNDIM (www.erndim.org) is an independent, not–for–profit foundation that provides EQA schemes in the field of inborn errors of metabolism with the aim of improving diagnosis, treatment and monitoring of inherited metabolic diseases.

The authors thank the group of Dr. Marie Zikánová (Charles University, Prague, Czech Republic) for making the substance SAICAr available. The excellent technical assistance of Kathrin V. Schwarz, Dorina Magin and Carola Friedrich is greatly acknowledged.

References

  1. 1. Balasubramaniam S, Duley JA, Christodoulou J. Inborn errors of purine metabolism: clinical update and therapies. J Inherit Metab Dis. 2014;37: 669–686. pmid:24972650
  2. 2. Balasubramaniam S, Duley JA, Christodoulou J. Inborn errors of pyrimidine metabolism: clinical update and therapy. J Inherit Metab Dis. 2014; 37: 687–698. pmid:25030255
  3. 3. Bakker JA, Bierau J. Purine and pyrimidine metabolism: still more to learn. Ned Tijdschr Klin Chem Labgeneesk. 2012; 37: 3–6.
  4. 4. Dudley E, Bond L. Mass spectrometry analysis of nucleosides and nucleotides. Mass Spectrom Rev. 2014; 33: 302–331. pmid:24285362
  5. 5. Rebollido-Fernandez MM, Castiñeiras DE, Bóveda MD, Couce ML, Cocho JA, Fraga JM. Development of electrospray ionization tandem mass spectrometry methods for the study of a high number of urine markers of inborn errors of metabolism. Rapid Commun Mass Spectrom. 2012; 26: 2131–2144. pmid:22886809
  6. 6. Hartmann S, Okun JG, Schmidt C, Langhans CD, Garbade SF, Burgard P, et al. Comprehensive detection of disorders of purine and pyrimidine metabolism by HPLC with electrospray ionization tandem mass spectrometry. Clin Chem. 2006; 52: 1127–1137. pmid:16613999
  7. 7. la Marca G, Casetta B, Malvagia S, Pasquini E, Innocenti M, Donati MA, et al. Implementing tandem mass spectrometry as a routine tool for characterizing the complete purine and pyrimidine metabolic profile in urine samples. J Mass Spectrom. 2006; 41: 1442–1452. pmid:17061293
  8. 8. Schmidt C, Hofmann U, Kohlmüller D, Mürdter T, Zanger UM, Schwab M, et al. Comprehensive analysis of pyrimidine metabolism in 450 children with unspecific neurological symptoms using high-pressure liquid chromatography-electrospray ionization tandem mass spectrometry. J Inherit Metab Dis. 2005; 28: 1109–1122. pmid:16435204
  9. 9. Ito T, van Kuilenburg AB, Bootsma AH, Haasnoot AJ, van Cruchten A, Wada Y, et al. Rapid screening of high-risk patients for disorders of purine and pyrimidine metabolism using HPLC-electrospray tandem mass spectrometry of liquid urine or urine-soaked filter paper strips. Clin Chem. 2000; 46: 445–452. pmid:10759467
  10. 10. Laikind PK, Seegmiller JE, Gruber HE. Detection of 5'-phosphoribosyl-4-(N-succinylcarboxamide)-5-aminoimidazole in urine by use of the Bratton-Marshall reaction: identification of patients deficient in adenylosuccinate lyase activity. Anal Biochem. 1986; 156: 81–90. pmid:3740420
  11. 11. Foley JP. Resolution equations for column chromatography. Analyst. 1991; 116: 1275–1279.
  12. 12. R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Available from https://www.R-project.org/. Retrieved July 2018.
  13. 13. Wilcox R. Introduction to Robust Estimation and Hypothesis Testing. 3rd ed. Elsevier; 2012.
  14. 14. Zikanova M, Krijt J, Skopova V, Krijt M, Baresova V, Kmoch S. Screening for adenylosuccinate lyase deficiency using tandem mass spectrometry analysis of succinylpurines in neonatal dried blood spots. Clin Biochem. 2015; 48: 2–7. pmid:25445730
  15. 15. Jimmerson LC, Bushman LR, Ray ML, Anderson PL, Kiser JJ. A LC-MS/MS Method for Quantifying Adenosine, Guanosine and Inosine Nucleotides in Human Cells. Pharm Res. 2017; 34: 73–83. pmid:27633886
  16. 16. la Marca G, Canessa C, Giocaliere E, Romano F, Malvagia S, Funghini S, et al. Diagnosis of immunodeficiency caused by a purine nucleoside phosphorylase defect by using tandem mass spectrometry on dried blood spots. J Allergy Clin Immunol. 2014; 134: 155–159. pmid:24767876
  17. 17. Wevers RA, Engelke UF, Moolenaar SH, Bräutigam C, de Jong JG, Duran R, et al. 1H-NMR spectroscopy of body fluids: inborn errors of purine and pyrimidine metabolism. Clin Chem. 1999; 45: 539–548. pmid:10102915
  18. 18. Bailey CJ. Orotic aciduria and uridine monophosphate synthase: a reappraisal. J Inherit Metab Dis. 2009; 32 Suppl 1: S227–233.
  19. 19. Grohmann K, Lauffer H, Lauenstein P, Hoffmann GF, Seidlitz G. Hereditary orotic aciduria with epilepsy and without megaloblastic anemia. Neuropediatrics. 2015; 46: 123–125. pmid:25757096
  20. 20. Coene KLM, Kluijtmans LAJ, van der Heeft E, Engelke UFH, de Boer S, Hoegen B, et al. Next-generation metabolic screening: targeted and untargeted metabolomics for the diagnosis of inborn errors of metabolism in individual patients. J Inherit Metab Dis. 2018; 41: 337–353. pmid:29453510
  21. 21. Körver-Keularts IMLW, Wang P, Waterval HWAH, Kluijtmans LAJ, Wevers RA, Langhans CD, et al. Fast and accurate quantitative organic acid analysis with LC-QTOF/MS facilitates screening of patients for inborn errors of metabolism. J Inherit Metab Dis. 2018; 41: 415–424. pmid:29435781
  22. 22. Gevi F, Zolla L, Gabriele S, Persico AM. Urinary metabolomics of young Italian autistic children supports abnormal tryptophan and purine metabolism. Mol Autism. 2016; 7: 47. pmid:27904735
  23. 23. Krijt J, Skopova V, Adamkova V, Cermakova R, Jurecka A, Kmoch S, et al. The need for vigilance: false-negative screening for adenylosuccinate lyase deficiency caused by deribosylation of urinary biomarkers. Clin Biochem. 2013; 46: 1899–1901. pmid:24183879
  24. 24. Chen BC, McGown IN, Thong MK, Pitt J, Yunus ZM, Khoo TB, et al. Adenylosuccinate lyase deficiency in a Malaysian patient, with novel adenylosuccinate lyase gene mutations. J Inherit Metab Dis. 2010; 33 Suppl 3: S159–162.
  25. 25. Stokvis E, Rosing H, Beijnen JH. Stable isotopically labeled internal standards in quantitative bioanalysis using liquid chromatography/mass spectrometry: necessity or not? Rapid Commun Mass Spectrom. 2005; 19: 401–407. pmid:15645520
  26. 26. Remane D, Wissenbach DK, Meyer MR, Maurer HH. Systematic investigation of ion suppression and enhancement effects of fourteen stable-isotope-labeled internal standards by their native analogues using atmospheric-pressure chemical ionization and electrospray ionization and the relevance for multi-analyte liquid chromatographic/mass spectrometric procedures. Rapid Commun Mass Spectrom. 2010; 24: 859–867. pmid:20196193
  27. 27. Chhana A, Lee G, Dalbeth N. Factors influencing the crystallization of monosodium urate: a systematic literature review. BMC Musculoskelet Disord. 2015; 16: 296. pmid:26467213
  28. 28. House CH, Miller SL. Hydrolysis of dihydrouridine and related compounds. Biochemistry. 1996; 35: 315–320. pmid:8555190