Open access peer-reviewed chapter

Synthesis of Low Abundant Vitamin D Metabolites and Assaying Their Distribution in Human Serum by Liquid Chromatography-Tandem Mass Spectrometry (LC-MS/MS) as a New Tool for Diagnosis and Risk Prediction of Vitamin DRelated Diseases

Written By

Lars Kattner and Dietrich A. Volmer

Submitted: 11 November 2015 Reviewed: 06 June 2016 Published: 12 April 2017

DOI: 10.5772/64518

From the Edited Volume

A Critical Evaluation of Vitamin D - Basic Overview

Edited by Sivakumar Gowder

Chapter metrics overview

2,239 Chapter Downloads

View Full Metrics

Abstract

This chapter provides an overview of versatile and efficient chemical syntheses of vitamin D derivatives by application of either linear or convergent synthesis approaches. Synthesis of the most relevant naturally occurring vitamin D metabolites and their deuterated counterparts to use as calibration and reference standards in LC-MS/MS assays is also shown. The chapter then summarizes the most important mass spectrometric approaches to quantify important vitamin D metabolites in human biofluids. In addition, new developments are described that are aimed at the pathobiological interpretation of the measured vitamin D metabolite distributions in various human diseases.

Keywords

  • vitamin D deficiency
  • biomarkers
  • low abundant vitamin D metabolites
  • assay development
  • LC-MS/MS
  • diagnosis
  • disease risk prediction

1. Introduction

Vitamin D is mostly known for its role in the regulation of calcium and phosphorous homeostasis [13]. Consequently, vitamin D deficiency may cause various disorders related to bone mineralization [4]. Drugs based on vitamin D analogs are commonly used to treat bone diseases (osteoporosis, osteomalacia, and rickets) or psoriasis. More recently, it has been suggested that vitamin D deficiency is also connected to a wide range of other diseases beyond bone mineralization, such as diabetes, autoimmune diseases, cardiovascular diseases, and cancer, as various clinical and epidemiological studies have shown [59]. However, the development of drugs for treatment of these diseases based on appropriate vitamin D analogs has mostly failed, either due to their rapid metabolic clearance or their calcemic effects. Vitamin D analogs are usually hormonally active compounds with pleiotropic functions and their levels in the body are strictly regulated by the hormonal system. In cases of oversupply, they are enzymatically degraded to avoid harmful effects such as calcemia, leading mainly to inactivation and conversion into water soluble degradation products suitable for renal clearance. Consequently, prevention of vitamin D deficiency rather than therapy of a vitamin D-related disease is a promising approach. Due to low concentration and short half-life of the active metabolite, 1α,25-dihydroxyvitamin D3 (1,25(OH)2D3, calcitriol) (4) in serum, the biosynthetic precursor 25-hydroxyvitamin D3 (25(OH)D3) (3) is usually measured as a marker of vitamin D status. Additionally, more than 50 naturally occurring but low abundant vitamin D metabolites of mostly unknown physiological function have been identified. Common assays, particularly ELISA and radio-immunoassays (RIA), are often restricted to 25(OH)D3 because of sensitivity and specificity limitations [1012]. Fortunately, recent mass spectrometry advances have permitted reaching deeper into the metabolic cascade of vitamin D, including low abundant species [1317]. Consequently, a wide variety of naturally occurring metabolites of potential biological activity can be analyzed simultaneously by liquid chromatography-tandem mass spectrometry (LC-MS/MS), which is currently considered the “gold standard” technique. However, this method has suffered in the past from the limited availability of the required vitamin D metabolites needed as calibration and reference standards. Consequently, as a prerequisite, all metabolites of interest have to be available for chemical synthesis, as described in this chapter.

Advertisement

2. Metabolism of vitamin D

Vitamin D3 (2) is mainly synthesized in the skin from 7-dehydrocholesterol (1) by UV irradiation and subsequent thermal isomerization of the intermediate product previtamin D3. Vitamin D3 is then hydroxylated in the liver to 25(OH)D3 (3), followed by further oxidation in the kidneys to 1,25(OH)2D3 (calcitriol) (4), which is usually considered the biologically active metabolite (Scheme 1). Calcitriol 4 is metabolized via CYP24 to other oxidative products, mainly 1α,24(R),25-trihydroxyvitamin D3 (1,24,25(OH)3D3) (5), followed by subsequent enzymatic oxidation and degradation of the carbon side chain to water soluble calcitroic acid [1, 18] for final clearance from the body. 25(OH)D3 (3) is degraded in an analogous manner via 24(R),25-dihydroxyvitamin D3 (24,25(OH)2D3) 6. Alternatively, 25(OH)D3 (3) and 1,25(OH)2D3 (4) can be metabolized to their corresponding 26,23-lactones (7,8) [19]. Vitamin D metabolites also have the potential to be metabolized through a C-3 epimerization pathway, leading to C-3-epi-metabolites such as 9-12, with an inversion of the stereogenic center at position C-3 [20, 21]. These reactions lead to products that are believed to be inactive. Since high concentrations of 3-epi-25-hydroxyvitamin D3 (3-epi-25(OH)D3) (10) were found in infants [22], it was initially assumed that this pathway is only a consequence of an immature vitamin D metabolism. It has also been shown, however, that the epimerization pathway is favored if, for any reason, side chain degradation by CYP24 is inhibited [23].

Scheme 1.

Metabolic pathways of vitamin D.

Finally, the corresponding metabolites of vitamin D2 (13) also have to be considered [18], because food from plant origin (in particular mushrooms) and food supplements may contain vitamin D2. Vitamin D2 metabolites have similar physiological function as compared to their corresponding vitamin D3 analogs, although their potency is apparently lower [24]. A detailed comparison of the metabolism of vitamin D2 and D3 [18] reveals that 25(OH)D3 (3) and 1,25(OH)2D3 (4) are preferentially metabolized at their side chains, particularly at C-23, C-24, and C-26, leading to metabolites susceptible to further oxidation, which is similar for vitamin D2, even though there are slight differences as a consequence of the C-22/23 carbon double bond and a methyl group at C-24. Thus, C-24 hydroxylation of vitamin D2 may occur initially, leading to 24-hydroxyvitamin D2, following by C-1 hydroxylation in the kidneys, giving 1,24-dihydroxyvitamin D2, which is subsequently oxidized in the side chain. Furthermore, 1,25-dihydroxyvitamin D2, formed in an analogous manner to its D3 counterpart, may be metabolized by side chain oxidation either to C-1,24,25 or C-1,25,26 trihydroxylated vitamin D2, again followed by further side chain oxidation. Each of these both pathways again leads to calcitroic acid as final metabolite.

Advertisement

3. Synthesis of low abundant vitamin D metabolites

The commercial unavailability of many relevant vitamin D3 and D2 species has limited the scope of LC-MS/MS assays in the past. For use as reference standards, vitamin D3 and D2 metabolites, as well as their corresponding stable isotopes (labeled with 2H (D) or 13C), have to be synthesized by application of a versatile and cost-effective methodology. A large variety of chemical syntheses of vitamin D derivatives has been developed in the last few decades by several academic and commercial groups [1, 2527], mainly with the aim of synthesizing new analogs for drug discovery and development purposes. These efforts have resulted in a large number of more than 3000 synthesized compounds [28]. Critical evaluation of these methods reveals, however, that only a few of them are suitable to reproducibly generate metabolites of interest in gram quantities at reasonable costs within short time frames. In this chapter, we will review some of the more suitable strategies that have been successfully applied and optimized in our laboratory.

The synthesis of deuterated vitamin D3 and D2 metabolites is mostly accomplished using the same procedures as those developed for nondeuterated metabolites.

The biosynthetic and technical synthesis (starting from an appropriate steroid precursor (analog to 1) is not favorable, due to low yields reached in the photochemical ring-opening reaction. It usually fails if a 1-hydroxy substituent is present. By far the most suitable methods for synthesis of vitamin D3 and D2 metabolites start with readily available Vitamin D2 (13). Two alternative strategies are applied, either by keeping the vitamin D skeleton intact during the course of the reaction sequence (linear synthesis) (Scheme 2), or by first cleaving the molecule to obtain the A-ring and the CD-ring building block, and subsequently reconnecting both molecules after separate appropriate chemical modifications (convergent synthesis) (Scheme 3) [1, 26, 27, 29, 30].

3.1. Linear synthesis of vitamin D metabolites

Following a linear synthesis (Scheme 2), the cis-diene moiety of vitamin D2 (13) is protected by sulfur dioxide after silylation of the hydroxyl group at C-3 as a tert-butyldimethylsilyl ether, leading to the SO2 adduct 14. Labeling with deuterium at C-6,19,19 can be carried out conveniently by treatment with D2O or deuterated alcohol (e.g., isopropanol). Finally, thermal removal of SO2 and photochemical isomerization leads to threefold deuterated vitamin D2 (d-13). This simple and convenient method to obtain C-6,19,19 deuterated metabolites [31] can also be applied to a wide variety of other vitamin D3 and D2 derivatives. In order to obtain metabolites of the corresponding C-3-epi series of vitamin D3 or D2 metabolites, the configuration of the C-3 OH group of vitamin D2 (13) can be epimerized under the so-called “Mitsunobu conditions” [32], by treatment with an aromatic acid and an azodicarboxylate, resulting in the corresponding ester 15 with concomitant inversion of the stereogenic center at C-3 (orientation of the substituent has changed from β to α). The ester is finally cleaved by saponification with potassium hydroxide or by reduction with lithium aluminum hydride to afford C-3-epi vitamin D2 (16), which can be used to synthesize a wide variety of C-3 epimers, including 3-epi-25-hydroxyvitamin D3 (3-epi-25(OH)D3) (10, d-10) and 3-epi-1α,25-dihydroxyvitamin D3 (3-epi-1,25(OH)2D3) (11, d-11). Yields are only moderate due to the favored formation of a C-3,4-5,6-7,8 all-trans triene system by elimination. To further proceed with the synthesis, the C-22/23 double bond in the side chain of SO2 adduct 14 is cleaved by ozonolysis, leading to alcohol 17 after reductive workup with sodium boronhydride, which can be used to add a wide variety of modified side chains as appropriate. For instance, 17 can be converted to an iodide, which is coupled in a nickel-zinc-mediated reaction with ethyl acrylate after prior removal of the SO2 protective group to afford 18. At this stage, an allylic oxidation by selenium dioxide can be carried out, leading to a 1α-hydroxy series of metabolites, such as 19 as appropriate precursor. Silylation of the newly generated hydroxyl group, followed by Grignard reaction with methylmagnesium bromide, leads to 1α,25-dihydroxy metabolites. If a deuterated Grignard reagent, labeled with 3 deuterium atoms at their methyl groups, is employed, the corresponding sixfold deuterium labeled metabolite is obtained, containing 3 deuterium atoms at C-26 and C-27, respectively. Finally, the C-5/6 double bond has to be isomerized photochemically, and the silyl protective groups have to be removed, ending up with nondeuterated or deuterated 25(OH)D3 (3, d-3) or 1,25(OH)2D3 (4, d-4) [33], as shown in Scheme 2. If 3-epi-vitamin D2 (16) is employed in this sequence, 3-epi-25(OH)2D3 (10, d-10) is obtained. 3-epi-1,25(OH)2D3 (11, d-11) is obtained, if 1α-hydroxylation by selenium dioxide was carried out, and the resulting C-1-OH-group is epimerized by oxidation and final reduction.

Vitamin D2 metabolites are synthesized by application of analogous strategies as applied for their corresponding D3 counterparts, although they are more challenging to synthesize, for two reasons: vitamin D2 metabolites contain an olefinic double bond in C-22/23-position, which has to be arranged in trans geometry during the course of a suitable coupling reaction of a CD-ring precursor and an alternated side chain building block by olefination. Most synthetic olefination methods yield a mixture of cis and trans double bonds, leading to product mixtures that are difficult to separate due to their similar polarity. If the reaction conditions are too basic, C-20 epimerization can also occur, leading to even more complex product mixtures. Additionally, vitamin D2 metabolites contain a chiral methyl group in C-24 position that has to be ranged in a defined configuration. When a suitable building block bearing this methyl group is employed in the synthesis, epimerization of the chiral center may occur under basic reaction conditions, along with the risk for epimerization at C-20 in the coupling reaction. Thus, alcohol 17 is converted to its corresponding phenylsulfone, which is coupled by olefination with aldehyde 20, a method originally invented by Julia [34]. Note that 20 can be synthesized in a few routine steps from commercially available (S)-(+)-3-hydroxyisobutyric acid methyl ester, which is submitted to a Grignard reaction with methyl-magnesium bromide. Analogously to the synthesis of the other sixfold deuterated metabolites labeled at C-26/27, the addition of deuterated Grignard reagent in the course of the reaction leads to sixfold deuterated vitamin D2 metabolites. Finally, the double bond at C-5/6 of 21 has to be photochemically isomerized and the silyl protective groups have to be removed, leading to nondeuterated or deuterated 25(OH)D2 (22, d-22) [35].

Scheme 2.

Linear synthesis of vitamin D metabolites. TBDMSCl: tert-butyldimethylsilyl chloride, SO2: sulfur dioxide, O3: ozone, NaBH4: sodium borohydride, iPrOD: deuterated isopropanol, Bu4NF: tetra-n-butylammonium fluoride, NaHCO3: sodium hydrogen carbonate, TsCl: p-toluenesulfonyl chloride, PhSH: thiophenol, MCPBA: 3-chloroperbenzoic acid, I2: iodine, PCC: pyridinium chlorochromate, D3CMgBr: deuterated methylmagnesium bromide, SeO2: selenium dioxide, DIAD: diisopropyl azodicarboxylate, PPh3: triphenylphosphine, KOH: potassium hydroxide, LiAlH4: lithium aluminium hydride.

Scheme 3.

Convergent synthesis of 24,25(OH)xD metabolites. TBDMSCl: tert-butyldimethylsilyl chloride, O3: ozone, NaBH4: sodium borohydride, I2: iodine, Cu: copper, Zn: zinc, D3CMgBr: deuterated methylmagnesium bromide, Me2CO: acetone, PCC: pyridinium chlorochromate, KMnO4: potassium permanganate, DIAD: diisopropyl azodicarboxylate, NCS: N-chlorosuccinimide, LiPPh2: lithium diphenylphosphide, H2O2: hydrogen peroxide. BuLi: n-butyllithium, Bu4NF: tetra-n-butylammonium fluoride, Pb(OAc)4: lead tetraacetate, SeO2: selenium dioxide.

3.2. Convergent synthesis of vitamin D metabolites

Alternatively, a convergent approach can be applied (Scheme 3), which is more versatile than a linear synthesis and allows for more harsh reaction conditions and wider scope of suitable substrates and reagents. In a classical and widely applied strategy, vitamin D2 (13) is submitted to ozonolysis, leading to Inhoffen-Lythgoe diol 23 and allylic alcohol 24 after reductive workup with sodium boronhydride. After appropriate chemical modification of the CD-ring precursor, this compound is coupled with an A-ring building block. Two strategies can be distinguished. In the most common approach, invented by Lythgoe and further developed by the Hoffmann-La Roche group [36], allylic alcohol 24 is converted via an allylic chloride 25 in diphenylphosphine oxide 26, which is coupled with a C-8 ketone by a Wittig-Horner olefination. The A-ring diphenylphosphine oxide 26 can be coupled before or after appropriate alteration of the side chain at the CD-ring building block. To employ an A-ring diphenylphosphine oxide such as 26 is favourable, if no further substituents except a C-3 OH group and a C-10 exo-methylene group in the A-ring are needed, as it is already present in vitamin D2 (13). As example for an efficient application of this strategy, the synthesis of 24,25(OH)2D3 (6) and its deuterated counterpart (d-6), their 3-epi-analogs (9/d-9) and 1α-analogs (5/d-5), is shown in Scheme 3 [37]. Here, diol 23 is converted to its corresponding iodide, which is coupled with enone 27 (obtained in a few routine steps from (R)-(-)-lactic acid using a zinc-copper-catalyzed reaction mediated by ultrasound, leading to 28. By Grignard reaction of 28 with nondeuterated or deuterated methyl-magnesium bromide, 29 or d-29 are obtained. The 1,2-diol moiety is protected as isopropylidene ether, followed by oxidation with pyridinium chlorochromate to obtain ketone 30/d-30, which can be coupled with diphenylphosphine oxide 26 to obtain d-24,25(OH)2D3 (d-6) after removal of silyl- and isopropylidene protective groups. To access the corresponding C-3-epi-analog, vitamin D2 is first bis-hydroxylated at C-7/8 using potassium permanganate. Inversion of the stereogenic center at C-3 under “Mitsunobu-conditions” using picolinic acid leads to the corresponding C-3-epi-triol 31. The yields of this reaction are much higher as compared to the synthesis of 16 used in the linear synthesis, because an elimination reaction leading to a favored C-3,4-5,6-7,8-all-trans system is avoided [38]. Silylation and cleavage of the diol moiety of triol 31 by lead tetraacetate and reductive workup with sodium boronhydride lead to the CD-ring fragment 32 and C-3-epi allylic alcohol 33, which can be used after conversion to the corresponding diphenylphosphine oxide 34. Note that 34 is coupled to d-30 to obtain d-3-epi-24,25(OH)2D3 (d-9) or its corresponding nondeuterated counterpart 9. 3-epi- diphenylphosphine oxide 34 can be used as a versatile A-ring building block to be employed in a Wittig-Horner reaction for the synthesis of a wide variety of 3-epi vitamin D3 and D2 metabolites, by connection with appropriate CD-ring building blocks with modified side chain, as shown for the synthesis of 3-epi-24,25(OH)2D3 (9, d-9) as representative example. Finally, permanganate oxidation of vitamin D2 (13) leads to triol 35, which is converted in a few routine steps, including a stereoselective 1α-allylic oxidation with SeO2 via 36 to diphenylphosphine oxide 37, which is then coupled with d-30 to afford 1,24,25(OH)3D3 (5/d-5).

Two other low abundant metabolites, 23(S),25(R)-dihydroxyvitamin D3 26,23-lactone 7 and 1α-1,23(S),25(R)-trihydroxyvitamin D3 26,23-lactone 8, are synthesized in an analogous manner [29] (Scheme 4), by coupling a CD-ring ketone carrying an appropriate side chain with diphenylphosphine oxide 26 or 37. Here CD-ring-building block 32 obtained from vitamin D2 13 is used, which is converted to aldehyde 38. In order to prepare an appropriate modified side chain, isopentenol 39 is converted using a few routine steps to sulfone 40, which is then coupled with 38 by a Julia olefination to afford 41. Removal of the isopropylidene protective group, oxidation to an unsaturated α-hydroxy carboxylic acid and subsequent iodolactonization leads to ketone 44 via 42 and 43, which is coupled with diphenylphosphine oxide 26 or 37, respectively, to give 7 and 8.

Scheme 4.

Convergent synthesis of 23,26(OH)x-lactone metabolites. TBDMSCl: tert-butyldimethylsilyl chloride, O3: ozone, LDA: lithium diisopropylamide, Na: sodium, Hg: mercury, PPTS: pyridinium p-toluenesulfonate, DMSO: dimethyl sulfoxide, Py: pyridine, SO2: sulfur dioxide, I2: iodine, KOH: potassium hydroxide, Bu3SnH: tributyltin hydride, Bu4NF: tetra-n-butylammonium fluoride, PCC: pyridinium chlorochromate, BuLi: n-butyllithium.

3.3. Convergent Palladium-catalyzed synthesis of vitamin D metabolites

The A-ring building block is preferably synthesized de novo, if further substituents apart from a C-3-OH group and a C-10 exo methylene group in the A-ring (already present in vitamin D2) are needed. This is particularly valuable for synthetic analogs containing substituents at C-2, which were developed in the past as new drugs [30]. However, natural metabolites may also contain additional substituents in the A-ring, such as 4β,25-dihydroxy vitamin D3 [39]. In these cases, Pd catalyzed tandem reactions can be applied (Scheme 5). Initially invented by Trost [40], an acyclic enyne 46 is coupled with a CD-ring vinyl bromide 47 obtained by Peterson olefination from the corresponding ketone. Closure of the A-ring and its connection with CD-ring is preferably carried out in one pot. Alternatively, the corresponding triflate 48 can be employed in a Zn-mediated reaction, which was invented by Negishi [41]. Additionally, CD-ring alkene boranates 49 can be used via an analogous coupling method originally invented by Suzuki [42]. Finally, a cyclic A-ring enynes, such as 50, can be coupled with CD-ring triflate 50, a coupling method invented by Sonogashira [43]. In all these cases, the CD-ring vinyl bromide 47, boranate 49, or triflate 51 are usually carrying the modified side chain already, before coupling with the acyclic or cyclic A-ring enyne. A major drawback of these methods is the fact that synthesis of acyclic or cyclic enynes or other A-ring building blocks not derived from vitamin D2 involves many synthesis steps, and in some cases separation of diastereomeric mixtures is also necessary to obtain enantiomerically and diastereomerically pure product. However, the suitable starting materials can usually be obtained from the natural chiral pool (i.e., terpenes such as carvone [44], malic acid [45], quinic acid [43], or carbon hydrates such as D-glucose [46] or D-xylose [47]), which are usually readily available. A wide range of substrates are tolerated. Consequently, these advanced strategies involving Pd catalyzed reactions have widely been applied in recent years and are by now well established.

Scheme 5.

Convergent Pd-catalyzed synthesis of vitamin D metabolites via de novo-A-ring synthesis. Pd: palladium, ZnCl2: zinc chloride.

Advertisement

4. LC-MS/MS assays for vitamin D metabolites

In the second part of this chapter, an overview of the mass spectrometric analysis of vitamin D metabolites in biological samples is presented. Because of the chemical nature of these secosteroidal molecules, liquid chromatography-tandem mass spectrometry (LC-MS/MS) currently provides the optimum analytical platform for analysis of vitamin D metabolites. While mass spectrometry assays are often not as rugged and more expensive than most non-mass spectrometric assays (in particular immunoassays), they provide the ability to capture multiple metabolites simultaneously at very low concentration levels. Immunoassays measure the 25(OH)D3 and 25(OH)D2 metabolites only and have limitations with respect to detection sensitivity as well as selectivity and specificity issues.

In most published assays, the metabolites are determined from serum or plasma matrices. Other biological matrices have rarely been used, although vitamin D metabolites have successfully been quantified in saliva [48] and various soft tissues [49]. In the following sections, we discuss the requirements and characteristics of LC-MS/MS of vitamin D metabolites; that is, chromatographic separation, ionization, m/z analysis, interferences and accuracy issues, multimetabolite screening applications, and the clinical role of vitamin D fingerprinting methods.

4.1. Liquid chromatographic separation of vitamin D metabolites

Due to the hydrophobic structure, vitamin D metabolites are generally easily separated on reversed-phase liquid chromatography stationary phases (e.g., octadecyl C18) materials utilizing hydrophobic interactions. The vitamin D metabolites generally elute in the order trihydroxylated < dihydroxylated < monohydroxylated metabolites, that is, for vitamin D3-related molecules, the order of chromatographic retention is 1,25(OH)2D3 < 25(OH)D3 < D3. Between corresponding D2 and D3 analogs, the D2 metabolites elute marginally later than the D3 versions. For the two important isomers of dihydroxylated vitamin D metabolites, 24,25(OH)2D3 and 1,25(OH)2D3, the order of retention is 24,25(OH)2D3 < 1,25(OH)2D3.

Possibly more interesting—from a chromatography point of view—are the biochemically formed isomers after stereochemical reversal (β→α) at C-3; for example, the epimers 25(OH)D3 and 3-epi-25(OH)D3. The subtle diastereomeric differences between these species can be distinguished using specialized columns, such as combined C18/chiral [50], pentafluorophenyl (PFP) [5153], and cyano (CN) [5457].

4.2. Ionization of vitamin D metabolites for mass spectrometric analysis

Mass spectrometric determination of vitamin D metabolites using liquid chromatography-mass spectrometry (LC-MS) is not trivial because of the structural limitations that the analytes provide with respect to attaching a charge to the molecules. Gas chromatography-mass spectrometry techniques have been used in the past for qualitative analysis and structure determinations, but these methods have been almost completely replaced by modern LC-MS methods. While analysis of transformation products such as vitamin D sulfates is relatively easy by LC-MS as the metabolites can be simply analyzed as deprotonated molecules, ionizing the relatively nonpolar vitamin D metabolites is not straightforward. For LC-MS, electrospray ionization (ESI) or atmospheric pressure chemical ionization (APCI) are the most common ionization techniques, which usually result in the formation of [M+H]+ ions for most biological molecules. APCI has been successfully applied to steroids, as gas-phase chemical ionization often efficiently transfers a proton to these types of molecules. ESI usually relies on a charging mechanism in the liquid phase and one would therefore expect ESI to be less efficient than APCI for vitamin D metabolites. In reality this is not the case, however, and both techniques have equally been applied to vitamin D analysis, with the analytical figures of merit being quite similar between the two ionization techniques. There appears to be a slight trend toward application of ESI rather than APCI in recent years.

An important, often neglected consideration in the choice of ionization source is the (usually detrimental) impact of coionized sample components that coelute with vitamin D metabolites from the LC column. Here, ionization suppression effects caused by these coeluting molecules may impact the outcome even stronger than the differences seen in ionization efficiencies between ESI and APCI [58, 59]. However, no systematic comparison of ionization efficiency/ionization suppression effects between ESI and APCI has been performed yet. Furthermore, coionization of coeluting components also leads to the formation of isobaric interferences when sample matrices such as plasma or serum are analyzed, which will be discussed in more detail below.

Another useful ionization technique is atmospheric pressure photoionization (APPI). This technique is very common in fields such as petroleum analysis, but surprisingly, it is virtually unknown in the clinical community. To our knowledge, only one study has systematically compared APPI with APCI and shown that APPI generated significantly higher ion currents for 25(OH)D3 than APCI [60]. A second study applied APPI to quantification of 25(OH)D3 without comparison to ESI or APCI [61].

Finally, another option to overcome problems of ionization efficiency and resulting problems with detection sensitivity is derivatization of vitamin D metabolites, to convert them into better responding transformation products. Such procedures are now quite common in the vitamin D analytical field; they comprise introduction of a chemical group that is readily ionized or is permanently charged. Cookson-type triazoline-diones and triazoline-dione-related reagents (e.g., 4-phenyl-1,2,4-triazoline-3,5-dione [PTAD]) are most often applied, which utilize the reaction of the reagent’s dienophile with the cis-diene group at C-5/C-6 and C-10/C-19 of vitamin D (Diels-Alder [4+2] reaction) (Scheme 6).

Several promising PTAD assays have been described in the literature [39, 49, 6264]. The advantage of using the cis-diene moiety is that all vitamin D metabolites contained in the sample are converted when the reagent is present at sufficient concentrations (i.e., in large excess quantities) as all relevant vitamin D metabolites possess this structural motif. Ionization efficiency of 25(OH)D-PTAD has been reported to be 100-fold higher than the nonderivatized molecules in ESI mode [65]. A structural variation in PTAD includes a permanently charged quaternary ammonium moiety, which enhances detection sensitivity even further [66]. This reagent is commercially available. A recent chemotyping assay for multiple vitamin D metabolites has shown detection limits in the fg/mL range for vitamin D metabolites from human serum (see discussion below).

Scheme 6.

Diels-Alder [4+2] derivatization of vitamin D metabolites using PTAD.

4.3 Mass analyzers for vitamin D analysis

Most clinical applications for vitamin D utilize quadrupole-based mass analyzers with low resolving powers. Because these mass spectrometers measure m/z signals at “unit mass resolution,” they cannot be operated in full scan mode for the accurate quantification of vitamin metabolites when complex sample matrices such as serum are analyzed. Serum contains multiple components of identical nominal molecular mass to vitamin D metabolites, several of which have been shown to coelute during High Performance Liquid Chromatography (HPLC) [67]. The coeluting metabolites then generate so-called “isobaric” ions in the mass spectra after ESI, which cannot be distinguished from the vitamin D target molecules using low resolution MS. Therefore, tandem mass spectrometry (MS/MS) is implemented, which activates the ionized species further and forces them to undergo collision-induced dissociation (CID) for generating structure-specific product ions. The metabolite-specific product ions are then measured and their response used for quantification of the target metabolites.

By far the most common low resolution tandem mass spectrometer for vitamin D analysis in clinical laboratories is the triple quadrupole (QqQ) instrument. In a QqQ mass spectrometer, the [M+H]+ ions of vitamin D metabolites are selected in the first quadrupole, fragmented in a quadrupole collision cell, and the products analyzed in the final quadrupole mass analyzer. The acquisition mode that is almost always applied is the so-called selected reaction monitoring (SRM) mode. In this mode, the [M+H]+ precursor ion is isolated in Q1 (e.g., m/z 401 for 25(OH)D3), dissociated in q2, and one specific product ion selected in Q3 (e.g., m/z 383). The quality of the precursor/product ion selection determines the selectivity of analysis; that is, the product ion structure should ideally directly reflect the chemical structure of the chosen precursor ion for maximum specificity. For vitamin D metabolites, the selection of an appropriate product ion is surprisingly difficult because of some inherent structural limitation of the precursor structure, leading primarily to simple dehydration reactions at low collision energies, which turn into complex mass spectra at higher collision energies (Figure 1) [68]. Unfortunately, because of multiple, overlapping series of fragmentation reactions, leading to “picket fence” type product ion spectra, fragment ions for vitamin D metabolites in the diagnostic m/z range often resemble each other.

Figure 1.

Collision-induced dissociation (CID) spectrum of the [M+H]+ ion of 1,25(OH)2D3 (spectrum after fully completed dehydration reactions after electrospray ionization and CID; m/z analysis by Fourier-transform ion cyclotron resonance, FTICR).

A few studies also report the use of a low resolution quadrupole ion trap (IT) [69, 70] in MS/MS mode for 25(OH)D3 and 25(OH)D2, but these instruments are rarely implemented in clinical environments. IT instruments are often not fit for purpose for quantitation of low abundant biological molecules, because of the large detrimental contributions of acquisition time overhead to the duty cycle, thus limiting detection sensitivity. There is also a general decline of this instrument type in modern mass spectrometry labs.

The most anticipated future development for vitamin D analysis is the introduction of high resolution mass spectrometry (HRMS), which is now firmly established in many other fields of modern mass spectrometry, in particular, in pharmaceutical applications and the proteomics field [71]. This trend is mostly due to the availability of robust quadrupole-quadrupole-time-of-flight (QqTOF) and orbitrap mass analyzers in recent years, which have transformed many analytical approaches to mass spectrometry. The use of mass defect as metabolite-specific property, for example, is now an integral part of many metabolite identification routines for drug metabolites [72]. While the number of applications of HRMS in the vitamin D field is still limited, the existing work has clearly demonstrated the potential of HRMS for vitamin D analysis. For example, orbitrap mass spectrometers in full scan [73, 74] and MS/MS [75, 76] modes have been applied successfully to the analysis of 25(OH)D3 in human serum and analytical performance has been shown to be equivalent or better than triple quadrupole and immunoassays. The important role of HRMS in the separation of isobaric interferences will be shown below.

4.4. Interferences during LC-MS/MS

The LC-MS/MS analysis of vitamin D metabolites is affected by various sources of error, which can affect both precision and accuracy. As with any other LC-MS/MS analysis from biological samples, ion suppression by coeluting sample components or chemical modifiers from the sample preparation or chromatography can lead to reduced analyte signals. There are several options to assess whether or not ion suppression is present, which have been summarized in many review articles, e.g., by Matuszewski et al. [77]. Stable isotope standards of vitamin D metabolites (mostly deuterated analogs) can usually correct for accuracy errors from ion suppression effects, as long as it is guaranteed that protein binding for the isotope standard is the same as for the endogenous analyte, which requires implementation of a careful incubation routine [78]. Importantly, deuterated isotope standards are commercially available for most of the relevant vitamin D metabolites but unfortunately not for all.

A second important source of analytical error originates from isobaric noise; that is, endogenous or exogenous metabolites that coelute with the vitamin D analytes and erroneously contribute to the analytical signal, if unspecific ions are used for mass spectral analysis. This has recently been described in detail by Qi et al. [67], who clearly demonstrated the presence of multiple isobars of 25(OH)D3 in human serum. The isobars have to be carefully removed in low resolution mass spectrometry, by application of appropriate MS/MS or ion mobility spectrometry routines [67]. A number of exogenous and endogenous molecules have been identified as relevant metabolites in serum samples [50, 67, 79].

Many of these interferences can be eliminated by application of high resolution mass spectrometry using sufficiently high resolving powers; for example, through implementation of orbitrap or Fourier-transform ion cyclotron resonance (FTICR) mass spectrometers. For example, Liebisch and Matysik demonstrated that the orbitrap MS instrument in their study was able to separate an isobaric interference of 25(OH)D3 in the MS/MS domain; this interference was caused by fragmentation of the d6-25(OH)D2 isotope standard [75].

Importantly, the issues relating to isobaric noise have only been studied for the 25(OH)D3 analyte; other metabolites will likely be affected by similar interferences, the impact of which during quantification, in particular, in multimetabolite assays (see below), remains unknown.

4.5. Method accuracy and certified reference materials

The vitamin D analytical community is supported through the Vitamin D External Quality Assessment Scheme (DEQAS), a nonprofit organization that evaluates the performance of analytical assays of member laboratories for 25(OH)D3 and 1,25(OH)2D3 [80] through round-robin analyses. DEQAS has clearly demonstrated that analytical performance of vitamin D analysis has greatly improved over the years between 1994 and 2009 [81].

The United States National Institute of Standards and Technology (NIST) provides certified standard solutions for 25(OH)D3 and 25(OH)D2. Furthermore, the NIST and the National Institutes of Health (NIH) Office of Dietary Supplements (ODS) have established the Vitamin D Metabolites Quality Assurance Program (VitDQAP), for interlaboratory comparison of measurement of 25(OH)D2, 25(OH)D3, and 3-epi-25(OH)D3 in serum and plasma and provide method-appropriate control materials.

A number of commercial reference, calibration, and quality control materials are available from several companies that allow rapid implementation and validation of vitamin D analytical methodologies.

4.6. Vitamin D multimetabolite assays

One of the most important advantages of LC-MS/MS assays over clinical immunoassays is the ability to determine multiple vitamin D species independently and simultaneously. As a result, there are now several very capable LC-MS/MS assays described in the literature that provide the capability for profiling the most relevant vitamin D metabolites at the same time, within a single analytical run, and with sufficient dynamic range to allow measuring the required physiological levels, down to the picomolar range. The topic has recently been reviewed in detail and the interested reader is referred to Ref. [82]. Briefly, most of these multimetabolite assays utilize derivatization techniques that transform all vitamin D species into better responding analogs (vide supra). This procedure in turn permits analysis of both high and low abundant vitamin D species with similar analytical figures of merit [62, 83].

The simultaneous acquisition of all important metabolite levels then provides the possibility of using these vitamin D metabolite distributions (“chemotypes”) as complex diagnostic or prognostic biomarkers for correlation with disease phenotype or clinical outcome of treatment. A few examples for such correlations are summarized in the last section.

Advertisement

5. Vitamin D fingerprints (chemotypes) in clinical applications

A number of recent studies have gone beyond the usual determination of 25(OH)D3—as marker for vitamin D status [84]—and 1,25(OH)2D3—for diagnosis of renal diseases, hypercalcemic syndromes, and disorders of 25(OH)D3 metabolism [85]. These broader profiling techniques are aimed at discovering dynamic effects of metabolites and catabolites, which are located further downstream the 25(OH)D3 metabolic cascade. Current studies highlight 24,25(OH)2D3 as important diagnostic marker, which was previously only considered a clearance product of vitamin D without activity. In fact, it has been shown that 24,25(OH)2D3 has crucial roles in bone metabolism [86] and renal diseases [87].

Capturing multiple vitamin D species and their dynamic changes allows for a better understanding of interindividual variations after vitamin D supplementation. Müller et al. recently demonstrated an inverse linear correlation between baseline 25(OH)D3 and response to supplementation for patients with chronic liver disease [83]. The study also showed that lower baseline 24,25(OH)2D3 levels were linked to larger changes of 25(OH)D3 levels, and that those patients who exhibited greater response to vitamin D supplementation had lower levels of 3-epi-25(OH)D3 (Figure 2).

Berg et al. implemented the ratio of 24,25(OH)2D3 and 25(OH)D3 as a novel status marker for vitamin D [88]. Wagner et al. used the same ratio and demonstrated that it was predictive of 25(OH)D3 response to supplementation [89]. Binkley et al. measured multiple vitamin D species and developed a model to describe interindividual variation of 25(OH)D3 levels after supplementation [90]; the authors highlighted the role of absorption (as measured by the nonmetabolized vitamin D3 species) and degradation (via the 24,25(OH)2D3 species) and presented a treat-to-target regime for tailored serum levels of 25(OH)D3 [90].

Figure 2.

Nonparametric correlations between baseline vitamin D metabolites and response to vitamin D supplementation for patients with chronic liver diseases after 6-month treatment: (a) the relative change in serum 25(OH)D3 (in response to vitamin D supplementation) correlated inversely with baseline 25(OH)D3 concentrations; (b) similarly, an inverse correlation between relative change in serum 25(OH)D3 and baseline 24,25(OH)2D3 was observed; (c) baseline 24,25(OH)2D3 concentrations were nonsignificantly higher in women as compared to men; (d) patients with lower 3-epi-25(OH)D3 concentrations at baseline tended to have a larger response to vitamin D supplementation (reprinted with permission from Ref. [83]).

Other important studies include the work by Bosworth et al. [87] and Stubs et al. [91], who utilized multimetabolite LC-MS/MS methods to characterize chronic kidney disease (CKD). Similarly, Duan et al. [64] studied patients with multiple sclerosis and observed comparable levels of 25(OH)D3 in the healthy control subjects and the patients; however, serum levels for 1,25(OH)2D3 and 24,25(OH)2D3 were lower in patients than controls.

Advertisement

6. Conclusions

The availability of assays for simultaneous capturing multiple vitamin D metabolites combined with reliable techniques for synthesis of the required metabolite standard compounds makes accurate measurement of metabolite distributions and subsequent correlation with disease phenotype readily possible; the outcome of these strategies is expected to be useful for diagnosis and risk prediction for various diseases. It may also allow for specific supplementation strategies in the future, which consider patient-specific dosage requirements and use of selected vitamin D metabolites or analogs.

References

  1. 1. Bouillon R, Okamura WH, Norman A: Structure and function in vitamin D endocrine system. Endocr Rev. 1995;16(2):200–257. DOI: 10.1210/edrv-16-2-200
  2. 2. Dusso AS, Brown AJ, Slatopolsky E: Vitamin D. Am J Physiol Renal Physiol. 2005;289:F8–F28. DOI: 10.1152/ajprenal.00336.2004
  3. 3. DeLuca H: Overview of general physiologic features and functions of vitamin D. Am J Clin Nutr. 2004;80(suppl):1689S–1696S. ISSN: 0002-9165
  4. 4. Holick M: Vitamin D and bone health. J Nutr. 1996;126:1159S–1164S. ISSN: 0022-3166
  5. 5. von Hurst PR, Stonehouse W, Coad J: Vitamin D supplementation reduces insulin resistance in South women living in New Zealand who are insulin resistant and vitamin D deficient – a randomized, placebo-controlled trial. Br J Nutr. 2010;103:549–555. DOI: 10.1017/S0007114509992017
  6. 6. Hyppönen E, Läärä E, Reunanen A et al.: Intake of vitamin D and risk of type 1 diabetes: a birth-cohort study. Lancet. 2001;358:1500. ISSN: 0099-5355
  7. 7. Lappe JM, Travers-Gustafson D, Davies KM et al: Vitamin D and calcium supplementation reduces cancer risk: results of a randomized trial. Am J Clin Nutr. 2007;85:1586–1591. Print ISSN: 0002-9165; Online ISSN: 1938-3207
  8. 8. Souberbielle JC, Body JJ, Lappe JM et al: Vitamin D and muskoskeletal health, cardiovascular disease, autoimmunity and cancer: recommendations for clinical practice. Autoimmun Rev. 2010;9:709–715. DOI: 10.1016/j.autrev.2010.06.009
  9. 9. Wang TJ, Pencina MJ, Booth SL et al: Vitamin D deficiency and risk of cardiovascular disease. Circulation. 2008;117:503–511. DOI: 10.1161/CIRCULATIONAHA.107.706127
  10. 10. Farrell J, Herrmann M: Determination of vitamin D and its metabolites. Best Pract Res Clin Endocrinol Metab. 2013;27(5):675–688. DOI: 10.1016/j.beem.2013.06.001
  11. 11. Zerwekh JE: The measurement of vitamin D: analytical aspects. Ann Clin Biochem. 2004;41:272–281. DOI: 10.1258/0004563041201464
  12. 12. Fraser WD, Milan AM: Vitamin D assays: past and present debates, difficulties, and developments. Calcif Tissue Int. 2013;92(2):118–127. DOI: 10.1007/s00223-012-9693-3
  13. 13. Geib T, Meier F, Schorr P, Lammert F, Stokes CS, Volmer DA: A simple micro-extraction plate assay for automated LC-MS/MS analysis of human serum 25-hydroxyvitamin D levels. J Mass Spectrom. 2015;50:275–279. DOI: 10.1002/jms.3522
  14. 14. Granado Lorencio F, Blanco-Navarro I, Pérez-Sacrsitán B: Critical evaluation of assays for vitamin D status. Curr Opin Clin Nutr Metab Care. 2013;16(6):734–740. DOI: 10.1097/MCO.0b013e328364ca96
  15. 15. Stokes CS, Volmer DA, Grünhage F, Lammert F: Vitamin D in chronic liver disease. Liver Int. 2013;33:338–352. DOI: 10.1111/liv.12106
  16. 16. Higashi T, Shimadab K, Toyo’oka T: Advances in determination of vitamin D related compounds in biological samples using liquid chromatography–mass spectrometry: a review. J Chromatogr B. 2010;878:1654–1661. DOI: 10.1016/j.jchromb.2009.11.026
  17. 17. Mena-Bravo A, Ferreiro-Vera C, Priego-Capote F, Maestro MA, Mouriño A, Quesada-Gómez JM, Luque de Castro MD: Quantitative analytical method to evaluate the metabolism of vitamin D. Clin Chim Acta. 2015;442:6–12. DOI: 10.1016/j.cca.2014.12.039
  18. 18. Horst RL, Reinhardt TA, Reddy GS. Chapter: Vitamin D metabolism. In: Feldman D, Pike JW, Glorieux FH, editors. 20 Vitamin D. 2nd ed. Burlington: Elsevier; 2005. p. 15–36.
  19. 19. Jones G: Metabolism and biomarkers of vitamin D. Scand J Clin Lab Invest Suppl. 2012;243:7–13. ISSN: 0036-5513
  20. 20. Kamao M, Tatematsu S, Sawada N, Sakaki T, Hatakeyama S, Kubodera N, Okano T: Cell specificity and properties of the C-3 epimerization of vitamin D3 metabolites. J Steroid Biochem Mol Biol. 2004;89–90:39–42. DOI: 10.1016/j.jsbmb.2004.03.048
  21. 21. Bailey D, Veljkovic K, Yazdanpanpanah M, Adeli K: Analytical measurement and clinical relevance of vitamin D3 C3-epimer. Clin Biochem. 2013;46:190–196. DOI: 10.1016/j.clinbiochem.2012.10.037
  22. 22. Singh RJ, Taylor RL, Reddy GS, Grebe SKG: C-3 Epimers can account for a significant proportion of total circulating 25-hydroxyvitamin D in infants, complicating accurate measurement and interpretation of vitamin D status. J Clin Endocrinol Metab. 2006;91(8):3055–3061. DOI: 10.1210/jc.2006-0710
  23. 23. Reddy GS, Rao DS, Siu-Caldera ML, Astecker N, Weiskopf A, Vouros P, Sasso GJ, Manchand PS, Uskokovic MR: 1a25-Dihydroxy-16-ene-23-yne-vitamin D3 and 1α,25-Dihydroxy-16-ene-23-yne-20-epi-vitamin D3: analogs of 1α,25-dihydroxyvitamin D3 that resist metabolism through the C-24 oxidation pathway are metabolized through the C-3 epimerization pathway. Arch Biochem Biophys. 2000;383(2):197–205. DOI: 10.1006/abbi.2000.2074
  24. 24. Houghton L, Vieth R: The case against ergocalciferol (vitamin D2) as a vitamin supplement. Am J Clin Nutr. 2006;84:694–697. Print ISSN: 0002-9165; Online ISSN: 1938-3207
  25. 25. Nadkarni S, Chodynski M, Corcoran A, Marcinkowska E, Brown G, Kutner A: Double point modified analogs of vitamin D as potent activators of vitamin D receptor. Curr Pharm Des. 2015;21:1741–1763. DOI: 10.2174/1381612821666141 205125113
  26. 26. Zhu GD, Okamura WH: Synthesis of vitamin D (calciferol). Chem Rev. 1995;95:1877–1952. DOI: 10.1021/cr00038a007
  27. 27. Dai H, Posner GH: Synthetic approaches to vitamin D. Synthesis 1994; 1994(12):1383–1398. DOI: 12 10.1055/s-1994-25697
  28. 28. Milne GWA, Delander M, editors. Vitamin D Handbook: Structures, Synonyms, and Properties. Hoboken: Wiley; 2008.
  29. 29. Kattner L, Bernardi D, Rauch E: Development of efficient chemical syntheses of vitamin D degradation products. Anticancer Res. 2015;35(2):1205–1210. ISSN: 0250-7005
  30. 30. Nadkarni S, Chodyński M, Krajewski K, Cmoch P, Marcinkowska E, Brown G, Kutner A: Convergent synthesis of double point modified analogs of 1α,25.dihydroxyvitamin D2 for biological evaluation. J Steroid Biochem Mol Biol. 2015 in press. DOI: 10.1016/j.jsbmb.2015.08.022
  31. 31. Ray R, Vicchio D, Yergey A, Holick MF: Synthesis of 25-hydroxy-[6,19,19´-2H3]vitamin D3. Steroids. 1992;57:142–146. DOI: 10.1016/0039-128X(92)90072-H
  32. 32. Mitsunobu O: The use of diethyl azodicarboxylate and triphenylphosphine in synthesis and transformation of natural products. Synthesis. Synthesis 1981; 1981(1): 1–28. DOI: 10.1055/s-1981-29317
  33. 33. Machand PS, Yiannikouros GP, Belica P, Madan P: Nickel-mediated conjugate addition. Elaboration of calcitriol from ergocalciferol. J Org Chem. 1995;60:6574–6581. DOI: 10.1021/jo00125a051
  34. 34. Ihara M, Suzuki S, Taniguchi T, Tokunaga Y, Fukumoto K: Modification of the Julia alkenation using SmI2-HMPA. Synlett. 1994 (10):859–860. DOI: 10.1055/s-1994-23033
  35. 35. Pietraszek A, Malinska M, Chodynski M et al: Synthesis and crystallographic study of 1,25-dihydroxyergocalciferol analogs. Steroids. 2013;78:1003–1014. DOI: 10.1016/j.steroids.2013.06.001
  36. 36. Baggiolini EG, Jacobelli JA, Hennessy BM, Batcho AD, Sereno JF, Uskokovic MR: Stereocontrolled total synthesis of 1α,25-dihydroxycholecalciferol and 1α,25-dihydroxyergocalciferol. J Org Chem. 1986;51:3098–3108. DOI: 10.1021/jo00366a004
  37. 37. Sestelo JP, Cornella I, Una O, Mouriño A, Sarandeses LA: Stereoselective convergent synthesis of 24,25-dihydroxyvitamin D3 metabolites: a practical approach. Eur J Chem. 2002;8(12):2747–2752. ISSN: 0947-6539
  38. 38. Achmatowicz B, Gorobets E, Marczak S, Przezdziecka A, Steinmeyer A, Wicha J, Zügel U: The first synthesis and biological testing of the enantiomer of 1α,25-dihydroxyvitamin D3. Tetrahedron Lett. 2001;42:2891–2895. DOI: 10.1016/S0040-4039(01)00317-3
  39. 39. Wang Z, Senn T, Kalhorn T, Zheng X, Zheng S, Davis C et al: Simultaneous measurement of plasma vitamin D3 metabolites, including 4β,25-dihydroxyvitamin D3, using liquid chromatography-tandem mass spectrometry. Anal Biochem. 2011;418:126–133. DOI: 10.1016/j.ab.2011.06.043
  40. 40. Trost BM, Dumas J, Villa M: New strategies for the synthesis of vitamin D metabolites via Pd-catalyzed reactions. J Am Chem Soc. 1992;114:9836–9845. DOI: 10.1021/ja00051a016
  41. 41. Gómez-Reino C, Vitale C, Maestro M, Mouriño A: Pd-catalized carbocyclization – Negishi cross-coupling cascade: a novel approach to 1α,25-dihydroxyvitamin D3 and analogues. Org Lett. 2005;7(26):5885–5887. DOI: 10.1021/ol052489c
  42. 42. Gogoi P, Sigüeiro R, Eduardo S, Mouriño A: An expeditious route to 1α,25-dihydroxyvitamin D3 and its analogues by an aqueous tandem palladium-catalyzed a-ring closure and Suzuki coupling to the C/D unit. Chem Eur J. 2010;16:1432–1435. DOI: 10.1002/chem.200902972
  43. 43. Sibilska IK, Szybinski M, Sicinski RR, Plum LA, DeLuca HF: Highly potent 2-methylene analogs of 1,a,25-dihydroxyvitamin D3: synthesis and biological evaluation. J Steroid Biochem Mol Biol. 2013;136:9–13. DOI: 10.1016/j.jsbmb.2013.02.001
  44. 44. Antony P, Sigüeiro R, Huet T et al.: Structure-function relationships and crystal structures of the vitamin D receptor bound 2α-methyl(20S,23RS)- and 2α-methyl-(20S,23R)-epoxymethano-1α,25dihydroxyvitamin D3. J Med Chem. 2010;53:1159–1171. DOI: 10.1021/jm9014636
  45. 45. Kato Y, Hashimoto Y, Naggasawa K: Novel heteroatom containing vitamin D3 analogs: efficient synthesis of 1α,25-dihydroxyvitamin D3-26,23-lactam. Molecules. 2003;8:488–499. DOI: 10.3390/80600488
  46. 46. Honzawa S, Suhara Y, Nihei KI et al.: Concise synthesis and biological activities of 2α-Alkyl- and 2α-(ω-Hydroxyalkyl)-20-epi-1α,25-dihydroxyvitamin D3. Bioorg Med Chem Lett. 2003;13:3503–3506. DOI: 10.1016/S0960-894X(03)00739-X
  47. 47. Moriarty RM, Kim J, Brumer III H: A general synthetic route to A-ring hydroxylated vitamin D analogs from pentoses. Tetrahedron Lett. 1995;36:51–54. DOI: 10.1016/0040-4039(94)02209-T
  48. 48. Higashi T, Shibayama Y, Fuji M, Shimada K: Liquid chromatography tandem mass spectrometric method for the determination of salivary 25-hydroxyvitamin D3: a noninvasive tool for the assessment of vitamin D status. Anal Bioanal Chem. 2008;391:229–238. DOI: 10.1007/s00216-007-1780-3
  49. 49. Lipkie TE, Janasch A, Cooper BR, Hohman EE, Weaver CM, Ferruzzi MG: Quantification of vitamin D and 25-hydroxyvitamin D in soft tissues by liquid chromatography-tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci. 2013;932:6–11. DOI: 10.1016/j.jchromb.2013.05.029
  50. 50. Shah I, James R, Barker J, Petroczi A, Naughton DP: Misleading measures in vitamin D analysis: a novel LC-MS/MS assay to account for epimers and isobars. Nutr J. 2011;10:46. DOI: 10.1186/1475-2891-10-46
  51. 51. Schleicher RL, Encisco SE, Chaudhary-Webb M, Paliakov E, McCoy LF, Pfeiffer CM: Isotope dilution ultra performance liquid chromatography-tandem mass spectrometry method for simultaneous measurement of 25-hydroxyvitamin D2, 25-hydroxyvitamin D3 and 3-epi-25-hydroxyvitamin in human serum. Clin Chim Acta. 2011;412:1594–1599. DOI: 10.1016/j.cca.2011.05.010
  52. 52. Baecher S, Leinenbach A, Wright JA, Pongratz S, Kobold U, Thiele R: Simultaneous quantification of four vitamin D metabolites in human serum using high performance liquid chromatography tandem mass spectrometry for vitamin D profiling. Clin Biochem. 2015;45(16–17):1491–1496. DOI: 10.1016/j.clinbiochem.2012.06.030
  53. 53. Strathmann FG, Sadilkova K, Laha TJ, LeSourd SE, Bornhorst JA, Hoofnagle AN, Jack R: 3-epi-25 hydroxyvitamin D concentrations are not correlated with age in a cohort of infants and adults. Clin Chim Acta. 2012;413:203–206. DOI: 10.1016/j.cca.2011.09.028
  54. 54. Tai SS-C, Bedner M, Phinney KW: Development of a candidate reference measurement procedure for the determination of 25-hydroxyvitamin D3 and 25-hydroxyvitamin D2 in human serum using isotope-dilution liquid chromatography-tandem mass spectrometry. Anal Chem. 2010;82:1942–1948. DOI: 10.1021/ac9026862
  55. 55. Keevil B: Does the presence of 3-epi-25OHD3 affect the routine measurement of vitamin D using liquid chromatography tandem mass spectrometry? Clin Chem Lab Med. 2012;50:181–183. DOI: 10.1515/cclm.2011.755
  56. 56. Stepman HCM, Vanderroost A, Stöckl D, Thienpont LM: Full-scan mass spectral evidence for 3-epi-25-hydroxyvitamin D3 in serum of infants and adults. Clin Chem Lab Med. 2001;49:253–256. DOI: 10.1515/CCLM.2011.050
  57. 57. Lensmeyer G, Poquette M, Wiebe D, Binkley N: The C-3 epimer of 25-hydroxyvitamin D(3) is present in adult serum. J Clin Endocrinol Metab. 2012;97:163–168. DOI: 10.1210/jc.2011-0584
  58. 58. Xu X, Mei H, Wang S, Zhou Q, Wang G, Broske L, Pena A, Korfmacher WA: A study of common discovery dosing formulation components and their potential for causing time-dependent matrix effects in high-performance liquid chromatography tandem mass spectrometry assays. Rapid Commun Mass Spectrom. 2005;19:2643–2650. DOI: 10.1002/rcm.2102
  59. 59. Jessome LL, Volmer DA: Ion suppression: a major concern in mass spectrometry. LC GC N Am. 2006;24:498–510. ISSN: 1527-5949
  60. 60. Adamec J, Jannasch A, Huang J, Hohman E, Fleet JC, Peacock M, Ferruzzi MG, Martin B, Weaver CM: Development and optimization of an LC-MS/MS-based method for simultaneous quantification of vitamin D2, vitamin D3, 25-hydroxyvitamin D2 and 25-hydroxyvitamin D3. J Sep Sci. 2001;34:11–20. DOI: 10.1002/jssc.201000410
  61. 61. Herrmann M, Harwood T, Gaston-Parry O, Kouzios D, Wong T, Lih A, Jimenez M, Janu M, Seibel MJ: A new quantitative LC tandem mass spectrometry assay for serum 25-hydroxy vitamin D. Steroids. 2010;75:1106–1112. DOI: 10.1016/j.steroids.2010.07.006
  62. 62. Ding S, Schoenmakers I, Jones K, Koulman A, Prentice A, Volmer DA: Quantitative determination of vitamin D metabolites in plasma using UHPLC-MS/MS. Anal Bioanal Chem. 2010;398:779–789. DOI: 10.1007/s00216-010-3993-0
  63. 63. Aronov PA, Hall LM, Dettmer K, Stephensen CB, Hammock BD: Metabolic profiling of major vitamin D metabolites using Diels-alder derivatization and ultra-performance liquid chromatography-tandem mass spectrometry. Anal Bioanal Chem. 2008;391:1917–1930. DOI: 10.1007/s00216-008-2095-8
  64. 64. Duan X, Weinstock-Guttman B, Wang H, Bang E, Li J, Ramanathan M, Qu J: Ultrasensitive quantification of serum vitamin D metabolites using selective solid-phase extraction coupled to microflow liquid chromatography and isotope-dilution mass spectrometry. Anal Chem. 2010;82:2488–2497. DOI: 10.1021/ac902869y
  65. 65. Higashi T, Awada D, Shimada K: Determination of 24,25-dihydroxyvitamin D3 in human plasma using liquid chromatography-mass spectrometry after derivatization with a Cookson-type reagent. Biomed Chromatogr. 2001;15:133–140. DOI: 10.1002/bmc.43
  66. 66. Hedman CJ, Wiebe DA, Dey S, Plath J, Kemnitz JW, Ziegler TE: Development of a sensitive LC-MS/MS method for vitamin D metabolites: 1,25 dihydroxyvitamin D2&3 measurement using a novel derivatization agent. J Chromatogr B Analyt Technol Biomed Life Sci. 2014;953–954:62–67. DOI: 10.1016/j.jchromb.2014.01.045
  67. 67. Qi Y, Geib T, Schorr P, Meier F, Volmer DA: On the isobaric space of 25 hydroxyvitamin D in human serum: potential for interferences in liquid chromatography/tandem mass spectrometry, systematic errors and accuracy issues. Rapid Commun Mass Spectrom. 2015;29:1–9. DOI: 10.1002/rcm.7075
  68. 68. Volmer DA, Mendes LR, Stokes CS: Analysis of vitamin D metabolic markers by mass spectrometry: current techniques, limitations of the “gold standard” method, and anticipated future directions. Mass Spectrom Rev. 2015;34:2–23. DOI: 10.1002/mas.21408
  69. 69. Higashi T, Awada D, Shimada K: Simultaneous determination of 25-hydroxyvitamin D2 and 25-hydroxyvitamin D3 in human plasma by liquid chromatography-tandem mass spectrometry employing derivatization with a Cookson-type reagent. Biol Pharm Bull. 2001;24:738–743. DOI: 10.1248/bpb.24.738
  70. 70. Gören AC, Bilsel G, Bilsel M: Rapid and simultaneous determination of 25-OH-vitamin D2 and D3 in human serum by LC/MS/MS: validation and uncertainty assessment. J Chem Metrl. 2007;1:1–9.
  71. 71. Ramanathan R, Jemal M, Ramagiri S, Xia YQ, Humpreys WG, Olah T, Korfmacher WA: It is time for a paradigm shift in drug discovery bioanalysis: from SRM to HRMS. J Mass Spectrom. 2011;46:595–601. DOI: 10.1002/jms.1921
  72. 72. Sleno L: The use of mass defect in modern mass spectrometry. J Mass Spectrom. 2012;47:226–236. DOI: 10.1002/jms.2953
  73. 73. Bruce SJ, Rochat B, Béguin A, Pesse B, Guessous I, Boulat O, Henry H: Analysis and quantification of vitamin D metabolites in serum by ultra-performance liquid chromatography coupled to tandem mass spectrometry and high-resolution mass spectrometry – a method comparison and validation. Rapid Commun Mass Spectrom. 2013;27:200–206. DOI: 10.1002/rcm.6439
  74. 74. Abdel-Khalik J, Crick PJ, Carter GD, Makin HL, Wang Y, Griffiths WJ: Studies on the analysis of 25-hydroxyvitamin D3 by isotope-dilution liquid chromatography-tandem mass spectrometry using enzyme-assisted derivatisation. Biochem Biophys Res Commun. 2014;446:745–750. DOI: 10.1016/j.bbrc.2014.01.088
  75. 75. Liebisch G, Matysik S: Accurate and reliable quantification of 25-hydroxy-vitamin D species by liquid chromatography high-resolution tandem mass spectrometry. J Lipid Res. 2015;56:1234–1239. DOI: 10.1194/jlr.D058511
  76. 76. Raml R, Ratzer M, Obermayer-Pietsch B, Mautner A, Pieber TR, Sinner FM, Magnes C: Quantifying vitamin D and its metabolites by LC/Orbitrap MS. Anal Methods. 2015;7:8961–8966. DOI: 10.1039/C5AY01583A
  77. 77. Matuszewski BK, Constanzer ML, Chavez-Eng CM: Strategies for the assessment of matrix effect in quantitative bioanalytical methods based on HPLC-MS/MS. Anal Chem. 2003;75:3019–3030. DOI: 10.1021/ac020361s
  78. 78. Vogeser M, Kyriatsoulis A, Huber E, Kobold U: Candidate reference method for the quantification of circulating 25-hydroxyvitamin D3 by liquid chromatography-tandem mass spectrometry. Clin Chem. 2004;50:1415–1417. DOI: 10.1373/clinchem.2004.031831
  79. 79. Maunsell Z, Wright DJ, Rainbow SJ: Routine isotope-dilution liquid chromatography-tandem mass spectrometry assay for simultaneous measurement of the 25-hydroxy metabolites of vitamins D2 and D3. Clin Chem. 2005;51:1683–1690. DOI: 10.1373/clinchem.2005.052936
  80. 80. Carter GD, Carter CR, Gunter E, Jones J, Jones G, Makin HLJ, Sufi S: Measurement of vitamin D metabolites: an international perspective on methodology and clinical interpretation. J Steroid Biochem Mol Biol. 2004;89–90:467–471. DOI: 10.1016/j.jsbmb.2004.03.055
  81. 81. Carter GD, Berry JL, Gunter E, Jones G, Jones JC, Makin HLJ, Sufi S, Wheeler MJ: Proficiency testing of 25-hydroxyvitamin D (25-OHD) assays. J Steroid Biochem Mol Biol. 2010;121:176–179. DOI: 10.1016/j.jsbmb.2010.03.033
  82. 82. Müller MJ, Volmer DA: Mass spectrometric profiling of vitamin D metabolites beyond 25-hydroxyvitamin D. Clin Chem. 2015;61:1033–1048. DOI: 10.1373/clinchem.2015.241430
  83. 83. Müller MJ, Stokes CS, Lammert F, Volmer DA: Chemotyping the distribution of vitamin D metabolites in human serum. Sci Rep. 2016;6:21080. DOI: 10.1038/srep21080
  84. 84. Hollick MF: Vitamin D status: measurement, interpretation, and clinical application. Ann Epidemiol. 2009;19:73–78. E-ISSN: 1873-2585
  85. 85. Hollick MF: The D-lemma: to screen or not to screen for 25-hydroxyvitamin D concentrations. Clin Chem. 2010;56:729–731. DOI: 10.1373/clinchem.2009.139253
  86. 86. van Leeuwen JP, van den Bemd GJ, van Driel M, Buurman CJ, Pols HA: 24,25-dihydroxyvitamin D3 and bone metabolism. Steroids. 2011;66:375–80. DOI: 10.1016/S0039-128X(00)00155-0
  87. 87. Bosworth CR, Levin G, Robinson-Cohen C, Hoofnagle AN, Ruzinski J, Young B, Schwartz SM, Himmelfarb J, Kestenbaum B, de Boer IH: The serum 24,25-dihydroxyvitamin D concentration, a marker of vitamin D catabolism, is reduced in chronic kidney disease. Kidney Int. 2012;82:693–700. DOI: 10.1038/ki.2012.193
  88. 88. Berg AH, Powe CE, Evans MK, Wenger J, Ortiz G, Zonderman AB, Suntharalingam P, Lucchesi K, Powe NR, Karumanchi SA, Thadhani RI: 24,25-dihydroxyvitamin D3 and vitamin D status of community-dwelling black and white americans. Clin Chem. 2015;61:877–884. DOI: 10.1373/clinchem.2015.240051
  89. 89. Wagner D, Hanwell HE, Schnabl K, Yazdanpanah M, Kimball S, Fu L, Sidhom G, Rousseau D, Cole DE, Vieth R: The ratio of serum 24,25-dihydroxyvitamin D3 to 25-hydroxyvitamin D3 is predictive of 25-hydroxyvitamin D3 response to vitamin D3 supplementation. J Steroid Biochem Mol Biol. 2011;126:72–77. DOI: 10.1016/j.jsbmb.2011.05.003
  90. 90. Binkley N, Lappe J, Singh RJ, Khosla S, Krueger D, Drezner MK, Blank RD: Can vitamin D metabolite measurements facilitate a “treat-to-target” paradigm to guide vitamin D supplementation? Osteoporos Int. 2015;26:1655–1660. DOI: 10.1007/s00198-014-3010-0
  91. 91. Stubbs JR, Zhang S, Friedman PA, Nolin TD: Decreased conversion of 25-hydroxyvitamin D3 to 24,25-dihydroxyvitamin D3 following cholecalciferol therapy in patients with CKD. Clin J Am Soc Nephrol. 2014;9:1965–1973. DOI: 10.2215/CJN.03130314

Written By

Lars Kattner and Dietrich A. Volmer

Submitted: 11 November 2015 Reviewed: 06 June 2016 Published: 12 April 2017