You are on page 1of 25

6032 J. Med. Chem.

2009, 52, 6032–6041


DOI: 10.1021/jm900775q

Hydroxamic Acids As a Novel Family of Serine Racemase Inhibitors: Mechanistic Analysis Reveals
Different Modes of Interaction with the Pyridoxal-50 -phosphate Cofactor

)


)
Hillary E. Hoffman,†,‡, Jana Jir a,†,‡, Petr Cı́gler,†,^ Miloslav Sanda,
askov †
Jan Schraml,§ and Jan Konvalinka*,†,‡

Gilead Sciences and IOCB Research Center, Institute of Organic Chemistry and Biochemistry of the ASCR, v. v. i., Flemingovo n. 2, 166 10
Prague 6, Czech Republic, ‡Department of Biochemistry, Faculty of Science, Charles University, Hlavova 8, Prague 2, Czech Republic, and

)
§
Institute of Chemical Process Fundamentals of the ASCR, v. v. i., Rozvojov a 135, 165 02 Prague 6, Czech Republic. These two authors
contributed equally to this work. ^ Current address: The Scripps Research Institute, 10550 North Torrey Pines Road, La Jolla, California 92037.

Received June 1, 2009

Mammalian serine racemase (SR) is a pyridoxal-50 -phosphate (PLP) dependent enzyme responsible for
the biosynthesis of the neurotransmitter D-serine, which activates N-methyl-D-aspartate (NMDA)
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

receptors in the CNS. Aberrant regulation of NMDA receptor signaling has been implicated in a variety
of neuropathologies, and inhibitors of SR would therefore be a worthwhile tool for further investigation
or treatment of such conditions. Here, we identify a series of small aliphatic hydroxamic acids (HAs)
that act as potent SR inhibitors. However, specificity studies showed that some of these HAs can act as
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

nonspecific inhibitors of PLP-dependent enzymes. We employed NMR, MS, and UV/vis spectroscopic
techniques to reveal that the nonspecific effect is likely due to irreversible interaction of the HA moiety
with PLP to form aldoxime species. We also characterize L-aspartic acid β-hydroxamate as a competitive
and selective SR inhibitor that could be used as a scaffold for further inhibitor development.

Introduction of active-site directed inhibitors with high specificity is a


a 0 particular challenge.6 Nevertheless, several PLP-dependent
Mammalian serine racemase (SR ) is a unique pyridoxal-5 -
enzymes have been recognized as pharmaceutical targets, as
phosphate (PLP)-dependent enzyme responsible for the bio-
synthesis of D-serine in the central nervous system. D-Serine recently reviewed by Amadasi et al.7 In the case of SR,
acts as a neurotransmitter and coagonist of N-methyl- however, very few potent and specific inhibitors have been
1
D-aspartate (NMDA) receptors, ionotropic glutamate recep-
identified to date; the two most potent competitive inhibitors,
tors that are found throughout the brain but predominantly malonic acid and L-erythro-3-hydroxyaspartic acid,8 exhibit
within the forebrain.2 NMDA-receptor-mediated glutamate inhibitory constants (Ki) in the micromolar range.
excitotoxicity has been implicated in a plethora of neuro- With the aim of identifying more effective SR inhibitors, we
pathological conditions, but treatment of patients with high- screened a panel of structurally diverse small molecules for the
affinity NMDA receptor blockers often results in undesirable inhibition of recombinant mouse SR (mSR). As a result of the
side effects, such as hallucinations. Thus, alternative methods screening, we identified a series of hydroxamic acids (HAs)
of decreasing NMDA receptor overactivation are highly with potent inhibitory activity. Since their discovery in 1869,9
sought after. Excitotoxic D-serine levels have been implicated the HAs have remained an intriguing family of organic
in Alzheimer’s disease3 and amyotrophic lateral sclerosis bioligands.10 Their bioactivity is usually associated with their
(Lou Gehrig’s disease),4 while low D-serine levels result in metal-binding properties [e.g., natural or artificial sidero-
NMDA receptor hypofunction, which may lead to symptoms phores for Fe(III) sequestration and strong chelators of Zn-
of schizophrenia.5 SR inhibitors therefore offer a novel (II)], but recent interest has focused on the use of HAs as
and potentially highly specific approach for attenuation of specific enzyme inhibitors and potential pharmacophores.
NMDA-receptor-mediated excitotoxicity and for further HAs have been shown to inhibit a number of enzyme targets
study of the pathway. including metalloproteases, hydrolases, ureases, lipoxy-
Since almost all PLP-dependent enzymes share a genases, cyclooxygenases, and peptide deformylases (for a
common transition state, the external aldimine, identification recent review, see refs 11 and 12). Several of the HAs have
been shown to be cell-permeable, and in at least one case, they
*Corresponding author. Tel: 420-220183218. Fax: 420-220183578. have been shown to cross the blood-brain barrier.13
E-mail: konval@uochb.cas.cz. In the present work, we identify several HAs, including a
a
Abbreviations: AR, alanine racemase; ATP, adenosine triphos- series of dihydroxamic acids (DHAs), with inhibitory activity
phate; DHA, dihydroxamic acid; DNPH, 2,4-dinitrophenylhydrazine;
DTT, dithiothreitol; EDTA, ethylenediaminetetraacetic acid; FDAA,
toward SR. In fact, one of them, succinodihydroxamic acid
1-fluoro-2,4-dinitrophenyl-5-L-alanine amide; HA, hydroxamic acid; (3, Chart 1), appears to be the most potent competitive
HEPES, 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid; IPTG, SR inhibitor identified to date. However, screening the HAs
isopropyl-β-D-1-thiogalactopyranoside; NMDA, N-methyl-D-aspar- for inhibition of a small panel of structurally and func-
tate; OD, optical density; PLP, pyridoxal-50 -phosphate; SDH, serine
dehydratase; SR, serine racemase (mSR, mouse SR; hSR, human SR); tionally diverse PLP-dependent enzymes revealed that the
TA, transaminase. DHAs capable of inhibiting SR also inhibit all the other
pubs.acs.org/jmc Published on Web 09/11/2009 r 2009 American Chemical Society
Article Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 6033

Chart 1. Structures of the Hydroxamic Acids Studied Split into


Four Structurally Diverse Groupsa
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

Figure 1. Graph of relative remaining SR activity in the presence of


selected hydroxamic acids compared to the noninhibited reaction.
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

Latin numerals indicate the structurally diverse groups of hydro-


xamic acids described in Chart 1. Activity was assayed at pH 8.0 in
the presence of 10 μM PLP (see Experimental Section for details)
with both the substrate (L-serine) and the inhibitor at 5 mM
concentration. A known potent inhibitor, L-erythro-3-hydroxyas-
partic acid,8 was used as the positive inhibition control (inhibitor).
Control represents the noninhibited reaction. The data are averages
of duplicate or triplicate measurements; the error bars show the
standard deviation from the mean. For the effective compounds,
IC50 values (μM) are shown. IC50 values were calculated from
two independent data sets. The standard deviation is within 25%,
except in the case of 4 (within 60%). *No remaining activity was
observed.

Results
Selected Hydroxamic Acids Potently Inhibit Mouse Serine
Racemase. A panel of hydroxamic acid analogues of car-
boxylic and amino acids was analyzed for the inhibition of
mouse serine racemase (mSR). We have recently shown that
mSR and its human orthologue (hSR) exhibit comparable
inhibitor sensitivity.14 Therefore, we used mSR for the initial
a screening to facilitate comparison of the results with our
The groups are (I) dihydroxamic acids (DHAs) and (II) hydroxamic
derivatives of amino acids (compound 8 is a mixture of two isomers: the
previously published mSR data.8 Compounds were screened
D-isomer with 2R,3S configuration and the L-isomer with 2S,3R config-
for inhibition of mSR-catalyzed L-serine racemization.
uration), (III) three representatives of benzohydroxamic acids, and (IV) L-Serine and the test compound were present at equimolar
formohydroxamic acid and the clinical compounds acetohydroxamic concentration (5 mM);9 this setup was chosen to reveal
acid (Lithostat) and vorinostat (Zolinza). ligands of comparable or higher affinity toward serine
racemase than the substrate L-serine. The inhibition screen-
enzymes tested. Although the DHAs are effective SR inhibi- ing data is summarized in Figure 1.
tors, their compromised specificity renders them unattractive The first group of test compounds (I) comprised two to six
candidates for further drug development. We employ NMR, and eight carbon chain dihydroxamic acids (Chart 1).
MS, and UV/vis spectrophotometry to show that the DHA Because of the unavailability of its synthetic precursor, the
inhibitors act as covalent modifiers of PLP, forming an seven carbon chain dihydroxamic acid was not included. The
aldoxime species and in effect depleting the available pool of three, four, and five carbon chain DHAs inhibit racemiza-
cofactor. tion significantly and with comparable IC50 values: malono-
Notably, not all of the inhibitory HAs identified by our dihydroxamic acid (2), 370 ( 30 μM; succinodihydroxamic
screening experiments are able to act nonspecifically via PLP- acid (3), 90 ( 20 μM; and glutarodihydroxamic acid (4),
aldoxime formation. L-Aspartic acid β-hydroxamate (10, 170 ( 110 μM. The two, six, and eight carbon chain
Chart 1) acts as a competitive SR inhibitor and forms an dihydroxamic acids (1, 5, and 6, respectively) do not affect
aldimine species with PLP that may act as a transition state enzyme activity. The second group of compounds (II) in-
mimetic. This compound is selective for SR and closely related cluded hydroxamic acid derivatives of the amino acid sub-
enzymes and could be used as a lead compound in further strates serine and threonine and the inhibitors glycine (Ki =
inhibitor development studies. 1640 μM8) and L-aspartic acid (Ki = 1900 μM15). Neither the
6034 Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 Hoffman et al.
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

Figure 2. Mechanism of action of 2 (A), 3 (B), and 10 (C,D) depicted as Lineweaver-Burk (A-C) and Eadie-Hofstee (D) plots. Data for the
plots were analyzed using global analysis nonlinear fit followed by the linear transformation provided by GraFit 5 software.19 Kinetic data were
recorded using an HPLC-based end point assay as described in Experimental Section. The substrate L-serine is abbreviated as LS. Inhibitor
concentrations (μM) are indicted on the right y-axes. Reaction velocity is in arbitrary units of product (D-serine) peak area relative to the
internal standard (glycine) peak in time. The data points represent triplicate experiments with an average standard deviation of 9%.

substrate (7 and 8) nor the glycine (9) R-hydroxamic acid the data to defined kinetic models (in this case, models for
derivatives showed significant inhibition. However, L-aspar- competitive, mixed-type, uncompetitive, or noncompetitive
tic acid β-hydroxamate (10) inhibited serine racemase with inhibition).18 DynaFit4 accomplishes precise numerical and
an IC50 of 1300 ( 200 μM. graphical tests to choose the best fitting model. The best
Additionally, we screened several other structurally di- fitting model for 2 is noncompetitive inhibition with a Ki
verse hydroxamic acids for SR inhibition (III and IV), value of 417 ( 52 μM. Data for 3 best match the competitive
including a panel of benzohydroxamic acids, exemplified model. The Ki value calculated for 3 is 3.6 ( 0.6 μM,
by 11, 12, and 13. None of these compounds significantly indicating that 3 is the most potent competitive SR inhibitor
affected serine racemization. characterized to date. Compound 10 is also a competitive
Hydroxamic acids often act by chelating functionally inhibitor with a Ki of 97.5 ( 23.7 μM. The results of the
important metal ions, and divalent cations are important mechanistic analysis are presented in Figure 2 as Linewea-
serine racemase activators.16,17 Since the screening reactions ver-Burk or Eadie-Hofstee plots prepared with GraFit5
were performed in the presence of 1 mM MgCl2, we were software.19
concerned that the observed SR inhibition might be an effect Specificity of the HAs. In order to test the specificity of the
of metal chelation. Therefore, we tested SR inhibition by 2 HA inhibitors for serine racemase, we screened 2, 3, and 10
(as a representative inhibitory compound) in the presence of for inhibition of a small panel of PLP-dependent enzymes.
varying concentrations of MgCl2. The inhibition of SR by 2 The noninhibitory compound 1 was included in the screen as
remained unaffected even in the presence of a 5-fold molar a control, as was the previously identified SR inhibitor
excess of Mg2þ over the inhibitor (data not shown). malonic acid.8 The enzyme panel comprised mouse SR,
Mechanism of Serine Racemase Inhibition by HAs. We human SR, rat liver serine dehydratase (SDH),20 alanine
selected 2, 3, and 10, hydroxamic acid analogues of known racemase from Bacillus stearothermophilus (AR), and bac-
competitive inhibitors,8,15 as representative compounds for terial transaminase (TA).
further analysis. Compound 4 was also an effective inhibitor, The results of the specificity screen are shown in Figure 3.
but since our synthetic preparation was less than 95% pure, All enzymatic reactions were performed at pH 8.0 in the
we excluded it from the detailed kinetic analyses. We ana- presence of 10 μM PLP using 5 mM compound (see Experi-
lyzed the mechanisms of action of 2, 3, and 10 by determina- mental Section for details). Compounds 2 and 3 showed an
tion of mSR racemization kinetics at various substrate/ inhibitory effect on all enzymes tested. Interestingly, malonic
inhibitor concentrations (see Experimental Section for acid, the dicarboxylic acid analogue of 2, is specific for mouse
details). The data were processed with DynaFit4 software, and human SR. Compound 1, which did not inhibit mSR,
which uses least-squares nonlinear regression analysis to fit did not inhibit any of the other enzymes tested. Alone among
Article Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 6035
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

Figure 3. Inhibition of a panel of PLP-dependent enzymes by selected hydroxamic acids. Inhibition was assayed at pH 8.0 in the presence of 10
μM PLP. The inhibitor concentration was kept at 5 mM, while the substrate concentration was chosen to be near the KM of the enzyme (see
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

Experimental Section for details). *No remaining activity was observed. #The tested compound acts as a substrate.

Figure 4. Absorbance spectra of PLP and selected compounds recorded at pH 8.0. The final concentration of the test compound was 5 mM,
and the PLP concentration was 0.3 mM.

the hydroxamic acids tested, the L-aspartic acid analogue 10 phosphate buffer, pH 8.0, and absorbance spectra were
exhibits some selectivity: it inhibits mSR, hSR, and their recorded in the 310-500 nm range. Results of the screening
homologue SDH (20% identity). It does not show any are shown in Figure 4 (1, 2, 3, 10, and 16) and Figure 2S
significant inhibitory activity toward AR, and it acts as a (Supporting Information) (4, 5, and 6). The absorbance
substrate of TA (the natural substrate of which is L-aspartic spectrum of PLP alone (Figure 4, black curve) exhibits
acid). maxima at 325 and 390 nm. The major chromophore at
The only common feature of the enzymes included in the 390 nm corresponds to PLP with an unsubstituted aldehyde
panel is the presence of a PLP-Schiff base in the catalytic moiety, and the minor peak at 325 nm corresponds to PLP-
center. Therefore, we hypothesized that the DHAs might act hydrate.21 As a positive control, we added an excess of
via modification of PLP, and to test this hypothesis, we hydroxylamine to PLP. Hydroxylamine reacts rapidly with
analyzed SR inhibition in the presence of various concentra- the aldehyde of PLP to form a PLP-aldoxime (see Scheme 1).
tions of PLP. When the concentration of PLP added to the The absorbance spectrum of this PLP-aldoxime is shown in
reaction mixture exceeded the concentration of 2 or 3 in Figure 4 (red curve); as expected, one major peak is visible at
molar equivalents, we observed quenching of the inhibition 335 nm, while the peak at 390 nm has disappeared entirely.
by excess PLP (data not shown), suggesting that certain HAs Addition of the inhibitory DHAs 2, 3, or 4 to PLP resulted in
may inhibit SR and other PLP-dependent enzymes by inter- similar spectral shifts. Addition of noninhibitory DHAs 1, 5,
acting with the cofactor. or 6 had no significant effect on the absorbance profile of free
Some Hydroxamic Acids React with PLP in Buffered PLP, indicating that these compounds do not react with the
Aqueous Solution. In order to investigate the putative inter- aldehyde of PLP. Addition of 10, a selective and competitive
action of HAs with PLP in aqueous solution, we analyzed HA inhibitor of SR, to PLP resulted in a unique absorbance
the reaction of PLP and HAs using UV/vis spectrophoto- spectrum (Figure 4, yellow curve). In addition to the forma-
metry. Solutions of PLP and various HAs were prepared in tion of a peak at 330 nm (similar to spectra of PLP in the
6036 Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 Hoffman et al.

Scheme 1 Formation of PLP-aldoximes in mixtures of PLP and


DHAs was further confirmed by ESI-MS analysis of the
reaction mixtures. A signal of 261.2 m/z, which corresponds
to PLP-aldoxime, was detected in the spectra of the inhibi-
tory DHAs (2, 3, and 4) mixed with PLP. Additionally, an
artificial ionization signal of 243.2 m/z was observed, which
corresponds to the aldoxime with one water molecule re-
moved. These two signals correspond to the major compo-
nents visible in the spectrum of a mixture of hydroxylamine
with PLP, and thus, both signals can be attributed to PLP-
aldoxime. Most of the spectra also contained a signal of
246.1 m/z, which corresponds to unmodified PLP. The
aldoxime signals did not appear in the spectra of noninhibi-
tory compounds 1, 5, 6, or 15 mixed with PLP; these spectra
contained only the signal corresponding to unmodified PLP.
Interaction of 10 with PLP Occurs by a Different Mechan-
ism. 1H NMR analysis showed that, like the DHAs, 10 reacts
with the aldehyde group of PLP, as evidenced by a decrease
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

presence of 2, 3, 4, and hydroxylamine), a second broad peak in the intensity of the CHO signal over time. However, the
centered around 405 nm appears in the spectrum. This PLP-aldoximes shown in Scheme 1 are not formed in this
second peak could indicate that the inhibitor interacts with case. Rather, the reaction of PLP and 10 results in the
PLP via its amino group, as PLP-amino acid Schiff bases are appearance of two new broad signals at 8.73 and 5.75 ppm
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

known to exhibit absorbance maxima in this region.22,23 (as well as a variety of minor side-products). A slow sub-
Two pharmacologically active compounds containing sequent reaction affects the CH-CH2 moiety of 10. The
a hydroxamic acid moiety, acetohydroxamic acid (15, proton on the R-carbon is abstracted, keeping the two CH2
Lithostat), an inhibitor of bacterial urease,24 and suberoyl- protons nonequivalent. Exchange of the R-proton with D
anilide hydroxamic acid (16, vorinostat, see Chart 1), a can be excluded (no line broadening of the CH2 signals was
histone deacetylase inhibitor marketed under the brand observed). These data do not allow for a conclusive determi-
name Zolinza,25 were added to the series as controls. Neither nation of the structure of the product(s) formed from the
5 mM 16 (data not shown) nor 5 mM 15 (see Figure 1) reaction of PLP and 10. However, on the basis of the known
exhibits inhibitory activity toward SR, and no change in the interactions of PLP with amino acids26-28 and on the avail-
spectrum of PLP upon the addition of 15 or 16 was observed. able NMR data, we propose that one of the major products
PLP as a Reactive Target for DHAs: Characterization of of the reaction of 10 with PLP is the aldimine shown in
Reaction Products. In order to analyze the product(s) result- Scheme 1.
ing from reactions of HAs with PLP, we monitored the MS analysis of the mixture of 10 and PLP offers support
reactions with 1H NMR spectroscopy (see Supporting In- for this proposal. At PLP concentrations of 4-5 mM, the
formation for background and details). All solutions were major feature of the spectrum is a peak at 246.1 m/z,
prepared in 100 mM phosphate buffer in D2O, and the pD corresponding to unmodified PLP, and only trace amounts
was kept as close to 8.0 as possible. Upon addition of 2, 3, or of the proposed PLP-10-aldimine (as a potassium adduct,
the noninhibitory compound 1 to PLP, the chemical shift of 416.3 m/z) can be observed. However, at higher PLP con-
PLP’s aldehydic proton does not change, but the signal centration (30 mM), the spectra contained a significant
intensity decreases until the signal disappears completely signal corresponding to the PLP-10-aldimine. Under these
(except in the case of 1). The rate of the decrease in intensity conditions, the corresponding PLP-amino acid-aldimines
depends on the nature of the DHA and decreases in the order can be observed in the spectra of mixtures of PLP and 7, 8,
3 > 2 > 1. The reaction of 3 with PLP went to completion in 9, L-serine, or glycine. The fact that the R-HAs 7, 8, and 9
less than a minute and resulted in the formation of two have no effect on SR activity (see Figure 1) suggests that
products. A thorough structural analysis based on 1H and these compounds probably do not form aldimine species in
15
N NMR (see Supporting Information Tables 2S and 3S for the active site, perhaps because of steric or electrostatic clash
a detailed explanation) resulted in the unambiguous identi- of the R-hydroxamic acid moiety with active site residues. To
fication of the major and minor products as the syn- and anti- address this hypothesis, it would be interesting to test the
aldoxime, respectively, as shown in Scheme 1. The reaction R-hydroxamic acid analogue of 10; however, since this
of PLP and 2 led to the same final products but proceeded compound is not commercially available and its synthesis
more slowly, and an intermediate product could be observed. is rather challenging, we did not include it here. In the case of
PLP and 1 underwent a very slow reaction (after 12 days, the β-HA 10, a competitive and selective SR inhibitor,
20% of the starting material remained), which eventually aldimine formation is a probable mechanism of inhibition.
resulted in the formation of syn- and anti-PLP-aldoxime via
Discussion
an unidentified intermediate (with chemical shifts different
from those of the intermediate observed in the reaction of Specific and potent inhibitors of mammalian serine race-
PLP and 2). The NMR analysis was also carried out with mase would be advantageous tools for further research of the
DHAs 4-6, with spectra recorded immediately and after 20 enzyme function in vivo and might have therapeutic potential
h. The noninhibitory compounds 5 and 6 did not react with for the treatment of neuropathologies associated with gluta-
PLP during this time period, while the inhibitory DHA 4 was matergic excitotoxicity. Recently published experiments with
reactive. After 20 h, the starting material had been comple- SR knockout mice support this hypothesis, showing that SR
tely consumed, and the PLP-aldoxime was formed. deficiency is correlated with a reduction in neurotoxic effects
Article Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 6037

caused by overexcitation of NMDA receptors.29 SR inhibitor repeating the screening assay with a panel of PLP-dependent
development has been addressed by several groups in recent enzymes [mouse SR, human SR, rat serine dehydratase
years.30,31,15,8 The most effective competitive inhibitors iden- (SDH), bacterial alanine racemase (AR), and bacterial trans-
tified by these groups are small amino and dicarboxylic acids, aminase (TA)]. Mouse and human SR share 89% sequence
with the most potent being L-erythro-3-hydroxyaspartic acid identity and are predicted to be members of the fold type II
(Ki = 43 μM) and malonic acid (Ki = 71 μM).8 In order to subfamily of PLP-dependent enzymes.33 SDH, which elim-
conduct experiments in animal models or tissue cultures, more inates L-serine to pyruvate in the liver, shares approximately
potent compounds are needed, and identification of novel 20% sequence identity with mammalian SR, and its crystal
structures capable of specifically inhibiting SR is therefore structure reveals that it is fold type II.34 AR shares no
very relevant. significant sequence homology with mammalian SR and is a
During our screening of various structurally diverse small member of the fold type III family,35 a family that is quite
molecules for SR inhibition, we identified a series of hydro- structurally distinct from other PLP-dependent enzymes. As
xamic acids as potent SR inhibitors. Hydroxamic acids have its name suggests, AR catalyzes the interconversion of D- and
been shown to inhibit a variety of enzymes, often by chelating L-alanine. TA produces oxaloacetate and L-glutamate from L-
functionally important metal ions. We immediately addressed aspartate and R-ketoglutarate and is a member of the fold type
the possibility that the observed inhibition was due to chela- I subfamily, also referred to as the asparate aminotransferase
tion of Mg2þ, a potent SR activator present in our screening family. The enzymes included in the panel are therefore quite
assay. Gratifyingly, the HAs retained their inhibitory activity structurally and functionally diverse and are representative of
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

even in the presence of a large molar excess of Mg2þ over the the variety of PLP-dependent enzymes. Direct comparison of
test compound, indicating that metal chelation does not play a the potency of the inhibitors toward the different enzymes is
significant role in SR inhibition. not possible under this setup, as potency is affected by the kcat/
Kinetic analysis of the mechanism of action of inhibitory KM of the enzymes toward their respective substrates and by
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

HAs shows that 10 is a classical competitive inhibitor with a Ki the reaction conditions. The results shown in Figure 3 should
of 97.5 ( 23.7 μM. Compound 10 is therefore a roughly 20- therefore be interpreted qualitatively rather than quantita-
fold more potent mSR inhibitor than its carboxylic acid tively. However, it can be concluded that the DHA not
analogue L-aspartic acid (competitive mechanism, Ki = capable of inhibiting mouse SR (1) did not inhibit any of the
1900 μM15). There could be several reasons for the increased other enzymes tested, while the most active DHAs (2 and 3)
potency of 10 compared to that of L-aspartic acid. Hydro- significantly inhibit all of the enzymes tested. Strikingly,
xamic acids generally have much higher pKa values than their malonic acid is a specific inhibitor of SR, while its dihydroxa-
corresponding carboxylic acids,32 which could lead to differ- mic acid analogue 2 inhibits all of the PLP-containing en-
ent electrostatic interactions with active site residues depend- zymes tested indiscriminately. Unlike 2 and 3, the L-aspartic
ing on the local pH in the active site. Furthermore, the acid analogue 10 is rather selective; it inhibits mouse and
hydroxamic acid moiety has the potential to form more human SR as well as their homologue SDH quite potently but
hydrogen bonding interactions with active site residues than does not inhibit the more distantly related AR. Compound 10
its carboxylic acid counterpart. However, since the 3-D acts as an inefficient substrate for TA and competes with the
structure of SR has not yet been solved, it is difficult to natural substrate L-aspartic acid.
speculate about which possibility is more likely. The ability of 2 and 3 to inhibit such a diversity of PLP-
Succinodihydroxamic acid (3), which behaves as a compe- dependent enzymes, together with the observed inhibition
titive inhibitor with a Ki of 3.6 ( 0.6 μM, is also more potent quenching in the presence of excess PLP, prompted us to
than its dicarboxylic acid analogue succinic acid, which only analyze the chemical mechanism of the observed enzyme
moderately inhibits mSR.8 In contrast, malonic acid is one of inhibition in more detail. Hydroxylamine, NH2OH, which is
the most potent competitive SR inhibitors identified to date used in the synthesis of hydroxamic acids, is a reagent
(Ki = 71 μM8), while its dihydroxamic acid analogue 2 acts as commonly used to remove PLP from PLP-dependent holoen-
a noncompetitive inhibitor with a Ki value of 417 ( 52 μM. It zymes. The presence of hydroxylamine in the inhibitor pre-
is surprising that 2, which differs from 3 only in the length of parations could explain their nonspecific interaction with
the carbon chain, behaves as a noncompetitive inhibitor, while PLP. However, all of the synthetic compounds described in
3 behaves as a competitive inhibitor. It should be noted that this work were purified as described in Experimental Section,
the inhibition mechanisms analyzed in the cases of 2 and 3 are and analysis of both 15N labeled and unlabeled compounds36
very complex. Both compounds are able to irreversibly inter- by NMR, mass spectrometry (MS), and elemental analysis
act with the PLP cofactor, as our NMR, MS, and UV/vis revealed that no significant hydroxylamine was present in the
analyses indicate, and therefore, the kinetic models used may final preparations. Furthermore, hydroxamic acids are gen-
not be entirely suitable to describe the inhibitor-SR interac- erally believed to be stable in aqueous solution, although they
tion. It is noteworthy, however, that in conditions approach- are known to undergo acid-catalyzed hydrolysis.37
ing the initial velocity of the enzymatic reaction (for details, We set out to analyze this putative interaction of HAs and
see Experimental Section) 3 and 2 behave as potent competi- PLP by UV/vis spectroscopy, NMR, and mass spectrometry.
tive and noncompetitive inhibitors, respectively. We introduce a rapid and simple spectroscopic assay that
In addition to the mechanism of action and potency, we also enables facile identification of compounds with the ability to
addressed the target specificity of the HAs. The development cross-react with PLP’s aldehyde moiety. Addition of the
of specific inhibitors for PLP-dependent enzymes is always a inhibitory DHAs (2, 3, and 4) to PLP caused a spectral shift
challenging task since the mechanisms of almost all PLP- consistent with the formation of PLP-aldoxime. As controls,
dependent enzymes, regardless of their substrate or reaction we used clinically available compounds containing a hydro-
specificity, proceed via condensation of the amino acid sub- xamic acid moiety, acetohydroxamic acid (15, Lithostat) and
strate with PLP, forming an external aldimine intermediate. vorinostat (16, Zolinza). These compounds do not inhibit SR,
We tested the selectivity of 2, 3, and 10 for mammalian SR by and we expected the compounds to be unreactive in the
6038 Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 Hoffman et al.

presence of PLP since there have been no reports of vitamin 5-membered rings at mildly alkaline pH has been demon-
B6 deficiency in patients treated by these drugs. Indeed, the strated.38 Cyclization of 4 would result in a 6-membered ring,
addition of noninhibitory compounds (1, 5, 6, 15, and 16) had which is sterically favored. Noninhibitory compounds 1, 5,
no significant influence on the UV/vis spectrum, suggesting and 6 would form 3-, 7-, and 9-membered rings, respectively;
that, as expected, these compounds are stable at physiological these conformations are less likely to occur. The only outlier in
pH and do not cross-react with PLP-dependent enzymes. this scheme is the inhibitory compound 2. Intramolecular con-
Again, 10 is an outlier. The major feature of the spectrum of densation of 2 would result in the formation of a 4-membered
the mixture of 10 and PLP is a peak at 330 nm, consistent with ring, which is not a sterically favorable conformation but is
PLP-aldoxime or PLP-hydrate formation. A second broad also not impossible to imagine. Alternatively, 2 may undergo
peak centered around 405 nm indicates that 10 might interact hydrolysis or interaction with PLP via a different mechanism,
with PLP via its amino group, forming the PLP-aldimine perhaps reflected in its different mode of inhibition.
species shown in Scheme 1. The findings presented here underscore the notion that
In order to analyze the structures of the putative products of caution must be taken when analyzing the results of large-
the reaction of PLP and HAs, we used 1H NMR and MS. scale inhibitor screening involving HAs and their derivatives.
NMR analysis of the interaction of DHAs 1, 2, and 3 with Indeed, the limited stability of some HAs under physiological
PLP allowed us to unambiguously identify the reaction conditions has been reported in other biomedicinal applica-
products as the syn- and anti-PLP-aldoximes shown in tions. Although HAs have been identified as one of the most
Scheme 1. It is interesting to note that the experimentally powerful inhibitors of Zn2þ-containing metalloproteases, in-
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

observed rates of reaction of DHAs with PLP (i.e., 1 , 2 , 3) corporation of this pharmacophore into drugs is not common,
agrees with the published order of hydrolytic stability of the partly due to the rapid metabolism of the hydroxamic acid
DHAs in acidic solution (1 > 2 > 3).37 One could speculate moiety.39,25
that the aldoximes reported here are in fact products of the In this work, we identify a series of DHAs capable of
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

reaction of PLP with hydroxylamine liberated upon hydro- nonspecific interaction with PLP-dependent enzymes via
lysis of DHAs. While HAs are generally known to be stable in modification of the cofactor to form a catalytically inactive
aqueous solutions, acid- or base-catalyzed hydrolysis is a aldoxime species. In contrast to the DHAs, one of the newly
possibility. Hydrolysis experiments in 100 mM phosphate identified HA inhibitors, 10, derived from L-aspartic acid, is a
buffer at pH 8.0 and pH 7.4 (see Supporting Information) potent competitive SR inhibitor and exhibits moderate selec-
offer support for this proposal, showing that the DHAs that tivity for mouse and human SR. Although it also slowly reacts
react rapidly with PLP (2, 3, and 4) are prone to hydrolysis, with the cofactor PLP, it does so less efficiently than the
while nonreactive DHAs (1, 5, and 6) are hydrolytically stable DHAs and yields a different mixture of products, in particular
under these conditions. However, while the UV/vis, MS, and an aldimine species that may act as a transition state mimetic.
NMR results draw a picture of the interaction of DHAs and Compound 10 could serve as a lead molecule for the devel-
PLP in solution, it is important to note that it remains unclear opment of the next generation of SR inhibitors, though its cell
whether the observed enzymatic inhibition by DHAs results permeability and ability to cross the blood-brain barrier
from this nonspecific depletion of PLP in solution or from must be addressed in the future. One can speculate that its
specific interaction of the inhibitor and enzyme followed by similarity to L-aspartic acid would allow it to use the same
cofactor depletion within the active site. physiological pathway.
The information gleaned from the NMR analysis further
highlights the complexity of interpreting the kinetic data Experimental Section
presented in Figure 2A and B. The enzymatic reaction velo- Materials. Enzymatic reaction mixtures were resolved on
cities were recorded at a single time point ranging from 8 to a Zorbax Extend C18 reversed phase HPLC column (4.6 
15 min. However, the NMR measurements revealed that 250 mm, particle size 5 μm, Agilent Technologies, USA)
hydroxylamine and 3 react with PLP to form PLP-aldoxime mounted on an Alliance 2795 HPLC system (Waters Co.,
within seconds, while the chemical reaction of PLP and 2 takes Milford, MA, USA) coupled to a Waters 2487 dual wavelength
considerably longer (on the order of minutes or hours). Like 3, absorbance detector. For NMR experiments, the pH was mea-
hydroxylamine behaves as a competitive inhibitor in our sured using a glass electrode (Spintrode, Hamilton) and a
assay conditions (data not shown), while 2 appears noncom- PHH2220 pH-meter (Radiometer) calibrated with Fisher Scien-
petitive. Adding a time dimension to the enzymatic assay may tific Buffers; the pH values were converted into pD values by
adding 0.40 as recommended.40 The 1H, 13C, and 15N NMR
help to further unravel the differences between 2 and 3,
spectra were measured on a Varian INOVA 500 spectrometer.
but such a detailed kinetic analysis is beyond the scope of The instrumental setup and experimental parameters have been
this work. described.41 Mass spectral data were recorded using a Q-TOF
Another interesting point raised in this study is the observa- microspectrometer from Waters Co. (Milford, MA, USA).
tion that compounds that are closely structurally related UV/vis absorbance scans were conducted on a Unicam UV500
(differing only in the length of the carbon chain) can have UV/vis spectrophotometer.
such different reactivities, both in buffered solution and in the D,L-Serine hydroxamate (7), D,L-threonine hydroxamate (8),
presence of PLP and enzyme. In this case, the intermediate glycine hydroxamate (9), L-aspartic acid β-hydroxamate (10),
chain-length DHAs 2, 3, and 4 proved to be inhibitory and acetohydroxamic acid (15), and pyridoxal-50 -phosphate were
able to react with PLP in solution according to UV/vis, MS, purchased from Sigma Aldrich and used without further pur-
ification. Experimental procedures for the synthesis of 11-14
and NMR analysis, while the compounds with very short (1)
can be found in Supporting Information. The purity and
and very long (5 and 6) carbon chains were not. One possibi- identity of synthetic compounds were accessed by elemental
lity is that some DHAs may undergo intramolecular conden- analysis and NMR spectroscopy. All compounds were found to
sation, resulting in the release of hydroxylamine, a notion that be at least 95% pure unless otherwise indicated.
is supported by hydrolysis experiments (see Supporting In- Synthesis of Dihydroxamic Acids (DHAs; 1-6). Both 15N
formation). The ability of analogues of 3 to cyclize to form labeled and nonlabeled DHAs were prepared from dicarboxylic
Article Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 6039

acid ethylesters by treatment with hydroxylamine.42 Detailed SR Inhibition Screening. Inhibition of mSR was analyzed as
procedures have been described.41 Briefly, the esters were con- described.8 Briefly, the racemization reactions were performed
verted to sodium dihydroxamates using an excess of hydroxy- in a pH 8.0 buffer supplemented with 10 μM PLP, 1 mM ATP,
lamine and sodium methoxide in dry methanol. The free DHAs 1 mM MgCl2, and 5 mM DTT. For the initial screening, a 5 mM
were liberated from aqueous solutions of dihydroxamates on a concentration of both the substrate L-serine and the test com-
weakly acidic catex column in Hþ-form. Lyophilization of pound was used. Reactions proceeded for 30 min at 37 C and
eluates and drying over P2O5 provided the final DHAs in good were stopped by the addition of 0.3 M HClO4. After deri-
yields (80-90%). The final preparation of compound 4 was vatization with 1-fluoro-2,4-dinitrophenyl-5-L-alanine amide
roughly 85% pure. (FDAA, Marfey’s reagent), the substrate L-serine and the
1-Hydroxy-2,5-pyrrolidinedione (N-hydroxysuccinimide) product D-serine were resolved and detected by HPLC.
was identified as a minor side-product (about 8 molar %) For the determination of IC50 values, various inhibitor con-
formed during the synthesis of 3. It was purchased in pure centrations were used while the L-serine concentration was kept
form from Sigma Aldrich. A freshly prepared solution of constant at 5 mM, and the data were analyzed with GraFit
N-hydroxysuccinimide (5 mM) did not significantly inhibit 5 software.19 For the determination of the mechanism of
mSR. inhibition, 4 or 5 substrate concentrations (from 1.25 to
Detailed NMR spectral data has been published elsewhere 12 mM L-serine) and a minimum of 2 inhibitor concentrations
for compounds 1-6.41,43 NMR parameters for the predominant and a noninhibited reaction set were used. We performed a
Z-conformers are given here. preliminary experiment with multiple reaction time points in
Oxalodihydroxamic acid (1): 1H NMR (300 MHz, DMSO- order to choose reaction times that approach the initial velocity
d6) δ 11.47 (bs, 2H), 9.16 (bs, 2H); 13C NMR (75 MHz, DMSO- of the enzymatic reaction and result in detectable production of
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

d6) δ 157.02; 15N NMR (50 MHz, DMSO-d6) δ -212.5. D-serine (not shown). The reaction time of the experiments
Malonodihydroxamic acid (2): 1H NMR (300 MHz, DMSO- presented here ranged from 8 to 15 min, and the conversion of
d6) δ 10.39 (bs, 2H), 8.93 (bs, 2H), 2.75 (s, 2H); 13C NMR L-serine to D-serine was 0.5-2%. All reactions were performed
(75 MHz, DMSO-d6) δ 163.65 (CO), 38.46 (CH2); 15N NMR in triplicate, and the averages and corresponding standard
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

(50 MHz, DMSO-d6) δ -213.4. deviations were analyzed with DynaFit4 software, which calcu-
Succinodihydroxamic acid (3): 1H NMR (300 MHz, DMSO- lates the least-squares nonlinear regressions and compares the
d6) δ 10.36 (bs, 2H), 8.69 (bs, 2H), 2.17 (s, 4H); 13C NMR possible mechanisms of action.18 The Ki values were calcu-
(75 MHz, DMSO-d6) δ 168.38 (CO), 27.99 (CH2); 15N NMR lated with DynaFit4 from one representative independent
(50 MHz, DMSO-d6) δ -216.1. experiment. The following inhibitor concentrations were used
Glutarodihydroxamic acid (4): 1H NMR (300 MHz, DMSO- for the determination of inhibitory constants: 2 (0, 150, and
d6) δ 10.36 (bs, 2H), 8.67 (bs, 2H), 1.94 (t, J = 7.2 Hz, 4H), 1.70 400 μM), 3 (0, 7.5, 10, and 30 μM), and 10 (0, 100, and 800 μM).
(p, J = 7.2 Hz, 2H); 13C NMR (75 MHz, DMSO-d6) δ 168.88 Lineweaver-Burk and Eadie-Hofstee transformations were
(CO), 31.92 (CH2, 2C), 21.58 (CH2, 1C); 15N NMR (50 MHz, performed with GraFit5 software.19
DMSO-d6) δ -216.1. Inhibition Screening with Other PLP-Dependent Enzymes.
Adipodihydroxamic acid (5): 1H NMR (300 MHz, DMSO- Alanine racemase from Bacillus stearothermophilus (AR) and
d6) δ 10.34 (bs, 2H), 8.68 (bs, 2H), 1.93 (unresolved multiplet, broad-range bacterial transaminase (TA) were purchased from
4H), 1.4 (unresolved multiplet, 4H); 13C NMR (75 MHz, Sigma Aldrich as lyophilized powders.
DMSO-d6) δ 169.14 (CO), 32.28 (CH2), 25.02 (CH2); 15N The plasmid pCW-SDH, which encodes rat liver serine
NMR (50 MHz, DMSO-d6) δ -215.0. dehydratase (SDH), was a generous gift from H. Ogawa.20 To
Suberodihydroxamic acid (6): 1H NMR (300 MHz, DMSO- express SDH, pCW-SDH was transformed into E. coli BL21-
d6) δ 10.37 (bs, 2H), 8.71 (bs, 2H), 1.92 (t, J = 7.4 Hz), 1.45 RIL cells, and the cells were cultured in 1 L of Luria-Bertani
(unresolved multiplet, 4H), 1.19 (unresolved multiplet, 4H); broth at 37 C. When the culture reached an OD600 of approxi-
13
C NMR (75 MHz, DMSO-d6) δ 169.26 (CO), 32.41, 28.50, mately 0.5, IPTG was added to a final concentration of 0.5 mM.
25.21; 15N NMR (50 MHz, DMSO-d6) δ -215.2. The cells were induced for 17 h at 37 C, then harvested at 2000g
Mouse Serine Racemase Preparation. Recombinant mouse for 15 min at 4 C. The cell pellet was resuspended in 20 mL of
serine racemase was prepared as previously described.8 Briefly, 20 mM Tris-HCl buffer, pH 8.0, containing 2 mM EDTA and
calcium chloride competent E. coli MC1061 cells were trans- 5 mM DTT. Ten milligrams of lysozyme and one protease
formed with pKS-mSR.8 Cell cultures were grown in nutrient- inhibitor tablet (Complete Mini EDTA free, Roche) were
rich medium at 37 C and 300 rpm. Serine racemase expression added, and cells were subjected to freeze/thaw and sonication.
was induced by 1 mM L-arabinose (Sigma) at OD600 = 0.8. The soluble fraction was clarified by centrifugation, and
After 5 h of induction, bacteria were harvested by centrifugation SDS-PAGE analysis of the supernatant revealed that SDH
(10000g, 10 min, 4 C) and stored at -70 C. was heavily overexpressed. To partially purify SDH, the lysate
The purification procedure was carried out in QA buffer was diluted with buffer to a final volume of 25 mL, and 5 g of
[20 mM triethanolamine hydrochloride-NaOH, pH 7.0, ammonium sulfate was added. After incubation at 4 C for 1 h,
1 mM MgCl2, 20 μM PLP, 0.1 mM D,L-dithiothreitol, and the SDH-containing precipitate was separated by centrifugation
0.02% (w/v) NaN3]. Pellets of induced bacteria were suspended and then redissolved in lysis buffer. This resulted in an SDH
in QA buffer supplemented with 1 mM phenylmethanesulfonyl preparation of about 50% purity, as judged by SDS-PAGE
fluoride, 0.2 mg/mL chicken egg lysozyme, and 0.05% (w/v) stained by Coomassie Brilliant Blue G-250.
sodium deoxycholate and homogenized using a glass Dounce The preparation of recombinant human serine racemase has
homogenizer. After the addition of 10 mM MgCl2 and DNase I been described,14 and its activity was assayed according to the
(Roche Molecular Biochemicals) to 10 μg/mL, the cell suspen- protocol for mouse SR described above.
sion was sonicated and centrifuged at 20000g for 30 min at 4 C. AR, SDH, and TA were assayed in 50 mM phosphate buffer,
The supernatant was subjected to the following three chro- pH 8.0, containing 10 μM PLP. The enzymes were preincubated
matographic steps: reverse phase (Phenyl-Sepharose FastFlow, with 5 mM inhibitor at room temperature for 15-20 min.
Pharmacia), ion-exchange (Q-Sepharose, FastFlow, Pharma- Reactions were started by the addition of substrate at a con-
cia), and affinity chromatography (ATP-agarose, Sigma). centration close to the KM of the enzyme (5 mM D-alanine for
The resulting 90% pure protein preparation was dialyzed three AR, 50 mM L-serine for SDH, and 5 mM L-aspartic acid and
times against 200 volumes of QA buffer at 4 C, concentrated to 5 mM R-ketoglutarate for TA). Reactions were allowed to
1 mg/mL (as determined by quantitative amino acid analysis), proceed for 10-40 min at 50 C (AR) or 37 C (SDH and
and stored at -70 C. TA). Reactions were quenched by the addition of TCA to a final
6040 Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 Hoffman et al.

concentration of 5% (w/v), and protein pellets were separated toxicity in amyotrophic lateral sclerosis. EMBO J. 2007, 26 (18),
by centrifugation. The supernatants were extracted twice with 4149–4159.
water-saturated diethyl ether prior to derivatization. SDH (5) Krystal, J. H.; D0 Souza, D. C. D-serine and the therapeutic
challenge posed by the N-methyl-D-aspartate antagonist model
reaction mixtures were derivatized with 2,4-dinitrophenylhy- of schizophrenia. Biol. Psychiatry 1998, 44 (11), 1075–1076.
drazine (DNPH), and pyruvate formation was monitored on (6) Toney, M. D. Reaction specificity in pyridoxal phosphate enzymes.
HPLC as previously described.8,44 AR and TA reaction mix- Arch. Biochem. Biophys. 2005, 433 (1), 279–287.
tures were derivatized with FDAA and separated according to (7) Amadasi, A.; Bertoldi, M.; Contestabile, R.; Bettati, S.; Cellini, B.;
the protocol for separation of L- and D-serine described above. di Salvo, M. L.; Borri-Voltattorni, C.; Bossa, F.; Mozzarelli, A.
AR mixtures were analyzed for the formation of L-alanine, and Pyridoxal 50 -phosphate enzymes as targets for therapeutic agents.
Curr. Med. Chem. 2007, 14 (12), 1291–1324.
TA mixtures were analyzed for the formation of L-glutamic acid. (8) Strisovsky, K.; Jiraskova, J.; Mikulova, A.; Rulisek, L.; Konva-
Pure L-alanine, D-alanine, L-aspartic acid, and L-glutamic acid linka, J. Dual substrate and reaction specificity in mouse serine
were derivatized and used as calibration standards. racemase: identification of high-affinity dicarboxylate substrate
NMR Measurements. An approximately 8 mM stock solution and inhibitors and analysis of the beta -eliminase activity. Bio-
of PLP was prepared in 100 mM phosphate buffer in D2O, pD chemistry 2005, 44 (39), 13091–13100.
8.4. An appropriate amount of hydroxamic acid (chosen to (9) Lossen, H. Uber die oxalohydroxamsaure. Justus Liebigs Ann.
Chem. 1869, 150 (3), 314–322.
ensure an approximately 1:1 molar ratio between the acid and (10) Marmion, C. J.; Griffith, D.; Nolan, K. B. Hydroxamic acids: An
PLP) was weighed out and mixed into 2 mL of PLP stock intriguing family of enzyme inhibitors and biomedical ligands. Eur.
solution. The final pD was adjusted to 7.1-8.1 by the addition J. Inorg. Chem. 2004, 15, 3003–3016.
of NaOH powder. For 13C NMR measurements (see Supporting (11) Muri, E. M. F.; Nieto, M. J.; Sindelar, R. D.; Williamson, J. S.
Information), an approximately 10-fold higher PLP concen- Hydroxamic acids as pharmacological agents. Curr. Med. Chem.
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

2002, 9 (17), 1631–1653.


tration was used. (12) Muri, E. M. F.; Nieto, M. J.; Williamson, J. S. Hydroxamic acids as
Mass Spectrometry. A 4, 5, or 30 mM solution of PLP (pH pharmacological agents: an update. Med. Chem. Rev. 2004, 1 (4),
adjusted to 8.5 with NaOH) was mixed with a 10-fold molar 385–394.
excess of the test compound and incubated at room temperature (13) Hockly, E.; Richon, V. M.; Woodman, B.; Smith, D. L.; Zhou, X.;
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q

for one day prior to measurement. Mixtures were prepared both Rosa, E.; Sathasivam, K.; Ghazi-Noori, S.; Mahal, A.; Lowden, P.
with and without a pH control (200 mM ammonium sulfate, pH A.; Steffan, J. S.; Marsh, J. L.; Thompson, L. M.; Lewis, C. M.;
Marks, P. A.; Bates, G. P. Suberoylanilide hydroxamic acid, a
8.0). The mixture was diluted in water 40- to 100-fold and histone deacetylase inhibitor, ameliorates motor deficits in a mouse
subjected to ESI-MS in a negative ion mode. model of Huntington0 s disease. Proc. Natl. Acad. Sci. U.S.A. 2003,
Spectrophotometric Assay for the Reaction of PLP with 100 (4), 2041–2046.
Hydroxamic Acids. A 0.6 mM solution of pyridoxal-50 -phos- (14) Hoffman, H. E.; Jiraskova, J.; Ingr, M.; Zvelebil, M.; Konvalinka,
phate and 10 mM solutions of test compounds were prepared in J. Recombinant human serine racemase: enzymologic character-
100 mM phosphate buffer, pH 8.0. The PLP and inhibitor ization and comparison with its mouse ortholog. Protein Expres-
sion Purif. 2009, 63 (1), 62–67.
solutions were mixed in a 1:1 ratio so that the final concentration (15) Dunlop, D. S.; Neidle, A. Regulation of serine racemase activity by
of PLP was 300 μM, and the final concentration of inhibitor was amino acids. Brain Res. Mol. Brain Res. 2005, 133 (2), 208–214.
5 mM. The reaction mixtures were incubated at room tempera- (16) Strisovsky, K.; Jiraskova, J.; Barinka, C.; Majer, P.; Rojas, C.;
ture for 1 h prior to measurement. Absorbance spectra were Slusher, B. S.; Konvalinka, J. Mouse brain serine racemase cata-
collected in the 310-500 nm range. lyzes specific elimination of L-serine to pyruvate. FEBS Lett. 2003,
535 (1-3), 44–48.
(17) Cook, S. P.; Galve-Roperh, I.; Martinez del Pozo, A.; Rodriguez-
Acknowledgment. This work was financially supported by Crespo, I. Direct calcium binding results in activation of brain
the Ministry of Education of the Czech Republic (grant serine racemase. J. Biol. Chem. 2002, 277 (31), 27782–27792.
(18) Kuzmic, P. Program DYNAFIT for the analysis of enzyme kinetic
1M0508) and the Grant Agency of the Academy of Sciences data: application to HIV proteinase. Anal. Biochem. 1996, 237 (2),
of the Czech Republic (grant IAA 400720706).We would like 260–273.
to thank Mr. Roman Vanı́cek for providing us with the very (19) Leatherbarrow, R. J. GraFit Version 5; Erithacus Software Ltd.:
helpful Sample Reader1 program and Mrs. Ludmila Horley, U.K., 2001.
(20) Ogawa, H.; Takusagawa, F.; Wakaki, K.; Kishi, H.; Eskandarian,
Soukupova for the synthesis of some hydroxamic acids. We M. R.; Kobayashi, M.; Date, T.; Huh, N.-H.; Pitot, H. C. Rat liver
would also like to thank the anonymous reviewers of this serine dehydratase. Bacterial expression and two folding domains
manuscript for their inspiring comments. as revealed by limited proteolysis. J. Biol. Chem. 1999, 274 (18),
12855–12860.
(21) Harris, C. M.; Johnson, R. J.; Metzler, D. E. Band-shape analysis
Supporting Information Available: Synthetic procedures for and resolution of electronic spectra of pyridoxal phosphate and
compounds 11-14 and additional spectroscopic measurements other 3-hydroxypyridine-4-aldehydes. Biochim. Biophys. Acta
(NMR and UV/vis). 1976, 421 (2), 181–194.
This material is available free of charge via the Internet at (22) Leussing, D. L. Vitamin B6 Pyridoxal Phosphate: Chemical, Bio-
http://pubs.acs.org. chemical and Medical Aspects Part A; John Wiley & Sons: New York,
1986; Vol. 1.
(23) Kallen, R. G.; Korpela, T.; Martell, A. E.; Matsushima, Y.;
References Metzler, C. M.; Metzler, D. E.; Morozov, Y. V.; Ralston, I. M.;
(1) Mothet, J.-P.; Parent, A. T.; Wolosker, H.; Brady, R. O., Jr.; Savin, F. A.; Torchinsky, Y. M.; Ueno, H. Transaminases; John
Linden, D. J.; Ferris, C. D.; Rogawski, M. A.; Snyder, S. H. D- Wiley & Sons: New York, 1985.
Serine is an endogenous ligand for the glycine site of the N-methyl- (24) Rosenstein, I. J.; Hamilton-Miller, J. M. Inhibitors of urease as
D-aspartate receptor. Proc. Natl. Acad. Sci. U.S.A. 2000, 97 (9), chemotherapeutic agents. Crit. Rev. Microbiol. 1984, 11 (1), 1–12.
4926–4931. (25) Grant, S.; Easley, C.; Kirkpatrick, P. Vorinostat. Nature Rev. Drug
(2) Monaghan, D. T.; Bridges, R. J.; Cotman, C. W. The excitatory Discovery 2007, 6 (1), 21–22.
amino acid receptors: their classes, pharmacology, and distinct (26) Metzler, D. E.; Snell, E. E. Deamination of serine. I. Catalytic
properties in the function of the central nervous system. Annu. Rev. deamination of serine and cysteine by pyridoxal and metal salts.
Pharmacol. Toxicol. 1989, 29, 365–402. J. Biol. Chem. 1952, 198 (1), 353–361.
(3) Hashimoto, K.; Fukushima, T.; Shimizu, E.; Okada, S.; Komatsu, (27) Snell, E. E. Summary of known metabolic functions of nicotinic
N.; Okamura, N.; Koike, K.; Koizumi, H.; Kumakiri, C.; Imai, K.; acid, riboflavin and vitamin B6. Physiol. Rev. 1953, 33 (4), 509–524.
Iyo, M. Possible role of D-serine in the pathophysiology of (28) Olivard, J.; Metzler, D. E.; Snell, E. E. Catalytic racemization of
Alzheimer0 s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry amino acids by pyridoxal and metal salts. J. Biol. Chem. 1952, 199
2004, 28 (2), 385–388. (2), 669–674.
(4) Sasabe, J.; Chiba, T.; Yamada, M.; Okamoto, K.; Nishimoto, I.; (29) Inoue, R.; Hashimoto, K.; Harai, T.; Mori, H. NMDA- and
Matsuoka, M.; Aiso, S. D-Serine is a key determinant of glutamate beta-amyloid1-42-induced neurotoxicity is attenuated in serine
Article Journal of Medicinal Chemistry, 2009, Vol. 52, No. 19 6041

racemase knock-out mice. J. Neurosci. 2008, 28 (53), 14486– (36) The labeled and unlabeled compounds behaved identically in the
14491. inhibition screening assay.
(30) Dixon, S. M.; Li, P.; Liu, R.; Wolosker, H.; Lam, K. S.; Kurth, M. (37) Ghosh, K. K.; Patle, S. K.; Sharma, P.; Rajput, S. K. A comparison
J.; Toney, M. D. Slow-binding human serine racemase inhibitors between the acid-catalysed reactions of some dihydroxamic acids,
from high-throughput screening of combinatorial libraries. J. Med. monohydroxamic acids and desferal. Bull. Chem. Soc. Jpn. 2003, 76
Chem. 2006, 49 (8), 2388–2397. (2), 283–290.
(31) Koval, D.; Jiraskova, J.; Strisovsky, K.; Konvalinka, J.; Kasicka, (38) Notari, R. E. Hydroxamic acids. II. Kinetics and mechanisms of
V. Capillary electrophoresis method for determination of D-serine hydroxyaminolysis of succinimide. J. Pharm. Sci. 1969, 58 (9),
and its application for monitoring of serine racemase activity. 1064–1068.
Electrophoresis 2006, 27 (13), 2558–2566. (39) Breuer, E.; Frant, J.; Reich, R. Recent non-hydroxamate matrix
(32) Simanenko, Y. S.; Prokop0 eva, T. M.; Popov, A. F.; Bunton, C. A.; metalloproteinase inhibitors. Expert Opin. Ther. Pat. 2005, 15 (3),
Karpichev, E. A.; Savelova, V. A.; Ghosh, K. K. O-Nucleophilicity 253–269.
of hydroxamate ions in reactions with ethyl 4-nitrophenyl (40) Glasoe, P. K.; Long, F. A. Use of glass electrodes to mea-
ethylphosphonate, diethyl 4-nitrophenyl phosphate, and 4-nitro- sure acidities in deuterium oxide. J. Phys. Chem. 1960, 64,
phenyl 4-toluenesulfonate. Russ. J. Org. Chem. 2004, 40 (9), 1337– 188–190.
1350. (41) Schraml, J.; Cigler, P. 15N-1H and 15N-13C couplings in 15N-
(33) Wolosker, H.; Blackshaw, S.; Snyder, S. H. Serine racemase: a glial enriched dihydroxamic acids. Magn. Reson. Chem. 2008, 46 (8),
enzyme synthesizing D-serine to regulate glutamate-N-methyl- 748–755.
D-aspartate neurotransmission. Proc. Natl. Acad. Sci. U.S.A. (42) Hurd, C. D.; Botteron, D. G. Hydroxamic acids from aliphatic
1999, 96 (23), 13409–13414. dicarboxylic acids. J. Org. Chem. 1946, 11, 207–214.
(34) Yamada, T.; Komoto, J.; Takata, Y.; Ogawa, H.; Pitot, H. C.; (43) Schraml, J.; Kvı́calova, M.; Blechta, V.; Soukupova, L.; Exner, O.;
Takusagawa, F. Crystal structure of serine dehydratase from rat Boldhaus, H.-M.; Erdt, F.; Bliefert, C. Structure determination of
liver. Biochemistry 2003, 42 (44), 12854–12865. dihydroxamic acids and their trimethylsilyl derivatives by NMR
Downloaded by CZECH ACADEMY OF SCIENCES on November 5, 2009 | http://pubs.acs.org

(35) Shaw, J. P.; Petsko, G. A.; Ringe, D. Determination of the spectroscopy. Magn. Reson. Chem. 2000, 38 (9), 795–801.
structure of alanine racemase from Bacillus stearothermophilus (44) Neidle, A.; Dunlop, D. S. Allosteric regulation of mouse brain
at 1.9-Å resolution. Biochemistry 1997, 36 (6), 1329–1342. serine racemase. Neurochem. Res. 2002, 27 (12), 1719–1724.
Publication Date (Web): September 11, 2009 | doi: 10.1021/jm900775q
Supporting Information

Hydroxamic acids as a novel family of serine racemase

inhibitors: mechanistic analysis reveals different modes of

interaction with the pyridoxal-5’-phosphate cofactor

Hillary E. Hoffman1,2#, Jana Jirásková1,2#, Petr Cígler1,##, Miloslav Šanda1,

Jan Schraml3, and Jan Konvalinka1,2*

1
Gilead Sciences and IOCB Research Center, Institute of Organic Chemistry and

Biochemistry of the ASCR, v. v. i., Flemingovo n. 2, 166 10 Prague 6, Czech Republic;


2
Department of Biochemistry, Faculty of Science, Charles University, Hlavova 8, Prague

2, Czech Republic; 3Institute of Chemical Process Fundamentals of the ASCR, v. v. i.,

Rozvojová 135, 165 02 Prague 6, Czech Republic

#
These two authors contributed equally to this work.
##
Current affiliation: The Scripps Research Institute, 10550 North Torrey Pines Road, La

Jolla, California 92037, USA


*
corresponding author. Tel: 420-220183218; Fax: 420-220183578; E-mail:

konval@uochb.cas.cz

S1
Table of Contents

Synthesis of ring-substituted benzohydroxamic acids (11, 12, 13) S3

Synthesis of formohydroxamic acid (14) S3

NMR Spectra of PLP S4

Table 1S S5

Model experiment with NH2OH: NMR assignment of aldoxime signals S5

Table 2S S7

NMR analysis of the reaction of PLP and DHAs S8

Table 3S S9

Concentration-dependence of the reaction: spectrophotometric analysis S9

Figure 1S S10

UV/Vis data for compounds 4 – 6 S10

Figure 2S S11

Hydrolysis of DHAs at pH 8.0 and 7.4 S11

Figure 3S S12

References S13

S2
Synthesis of ring-substituted benzohydroxamic acids (11, 12, 13)

The synthesis and thorough NMR characterization of these compounds has been

published elsewhere.1 Briefly, the compounds were prepared from carboxylic esters in

methanol or from acyl chlorides in chloroform according to standard procedures.2, 3 In

the case of 11, the reaction of the methyl ester in aqueous solution was followed by

extraction of the acid into ethanol. All compounds were recrystallized from water,

aqueous acetone, acetone-pentane, or aqueous ethanol.

4-aminobenzohydroxamic acid (11) 1H NMR (300 MHz, DMSO-d6) δ 10.730 (bs, 1H),
13
8.643 (bs, 1H), 7.48 (d, J = 8.5 Hz, 2H), 6.53 (d, J = 8.5 Hz, 2H); C NMR (75 MHz,

DMSO-d ) δ 112.73, 119.26, 128.46, 151.81, 165.21; 15N NMR (50 MHz, DMSO-d6) δ -
6

218.3 (NHOH), -314.5 (NH2).

4-chlorobenzohydroxamic acid (12) 1H NMR (300 MHz, DMSO-d6) δ 11.301 (bs, 1H),
13
9.107 (bs, 1H), 7.76 (m, 2H), 7.54(m, 2H); C NMR (75 MHz, DMSO-d ) δ 128.62,
6

128.92, 131.70, 136.06, 163.34; 15N NMR (50 MHz, DMSO-d6) δ -215.3 (NHOH).

4-nitrobenzohydroxamic acid (13) 1H NMR (300 MHz, DMSO-d6) δ 11.558 (bs, 1H),

9.315 (bs, 1H), 8.31 (d, J = 8.6 Hz, 2H), 8.00 (d, J = 8.6 Hz, 2H);13C NMR (75 MHz,

DMSO-d ) δ 123.78, 128.57, 138.71, 149.18, 162.54; 15N NMR (50 MHz, DMSO-d6) δ -
6

213.2 (NHOH), -11.2 (NO2).

Synthesis of formohydroxamic acid (14)

Formohydroxamic acid (14) was prepared from formic acid ethyl ester using a similar

protocol as for DHA synthesis. Briefly, 60 mL of sodium methoxide was added to a

solution of 0.145 mol hydroxylamine hydrochloride in a 2°C water bath. The solution

was filtered to remove NaCl, and the filtrate was treated with 0.189 mol ethyl formate.

S3
The mixture was incubated overnight at room temperature, and methanol was removed in

a rotary evaporator (34°C, 15 mbar). The solid phase was extracted several times with

methylacetate, which resulted in the formation of white crystals. The crystals were

filtered, washed with dichloromethane, and dried. Excess formic acid was removed under

high vacuum. The spectroscopic parameters of the final product were consistent with the

published values.4

NMR Spectra of PLP

Literature 1H NMR data on PLP are summarized in Table 1S along with experimental
13
values obtained in this work. C NMR parameters and their pH dependence have also

been reported in literature.5-7 However, due to the sensitivity issue, they were obtained

either after selective 13C enrichment7 or at concentrations far exceeding those relevant to

biochemical studies.5,6

Our preliminary experiments indicated that 1H NMR spectroscopy is well suited to follow

the putative reaction of PLP and DHAs. The spectra consist of well separated singlets and
31
closely spaced doublets (due to CH2 proton coupling with the P nucleus of the

phosphate moiety), and the reaction with DHAs does not lead to any signal overlap in the

most significant region (δ = 10.5 – 7.0 ppm). A small overlap occurs with CH2O proton

signals, which have signals close to the residual solvent signal.

S4
Table 1S. 1H NMR chemical shifts (and 31P-1H coupling constants) of PLP in water

Solution CHO C6-H 5-CH2O 2-CH3 Reference

(arom)
8, 9
Acidic 10.50 8.20 5.05 2.60

(pD=4.1) 6.53(hydrate)
9
Neutral 10.40 7.77 5.03 (7 Hz) 2.43

(pD=7.25)
8
Neutral 10.40 7.72 5.00 (5 Hz) 2.40

(pD=7.8)
8
Alkaline 10.37 7.62 4.95 2.34

(pD=9.4)

100mM 10.28 7.61 4.94 2.31 this work

phosphate (0.5 Hz) (6.0 Hz)

buffer

(pD=8.42)

Model experiment with NH2OH: NMR assignment of aldoxime signals


15
Two products were formed from the reaction of PLP with NH2OH HCl. The products

were formed in an 88:12 ratio with signals at δ = 8.345 and 7.839 split into doublets by

coupling to the 15N nucleus [with 2J(15N-1H) of 2.4 and 14.0 Hz, respectively, (see Table
15
3S) as confirmed by a N decoupling experiment (1H{15N})]. The major product was

identified as PLP-aldoxime with syn geometry (see Scheme 1), as our values match those

previously reported for the syn isomer of PLP-aldoxime [δ = 8.43 and 2J(15N-1H) = 2.7

S5
Hz].10 Though the corresponding parameters for the anti isomer have not been reported,

the 1H chemical shifts of the minor product roughly agree with those reported for anti

isomers of other aromatic oximes (δ = 7.2 – 7.7 ppm)11 as does the 2J(15N-1H) coupling

[2J(15N-1H) = 13.8 – 18.8 Hz].10 The failure of previous authors10 to observe this anti

isomer is paralleled by similar observations on oximes of other aromatic aldehydes,

especially those containing phenolic OH groups (as found in PLP). The OH group is

thought to autocatalyze rearrangement of anti into syn isomer.12 Purification of the

reaction products by crystallization removes the less abundant isomer, while we directly

measured the unpurified reaction mixture, which enabled us to observe the anti isomer.

[An analogous situation was recognized for aliphatic aldoximes which appeared isolable

in a single form (anti) only.13 In a recent study of closely related PLP-aldimines

Limbach’s group observed only E isomers14 analogous to our syn isomer.]

The reaction could be carried out at sufficiently high concentration (81 mM PLP) to
13
allow measurements of C NMR spectra in a reasonable time. The results are

summarized in Table 2S.

S6
13 15
Table 2S. C NMR chemical shifts (and N-13C/31P-13C coupling constants in Hz in

parentheses/brackets, respectively) of products of reaction of PLP with 15NH2-OH (pD =

8.04, 81mM PLP); the shifts are referenced to 1,4-dioxane at δ(13C) = 66.30.

Product CH=N C2/3/4/5/6 5-CH2O 2-CH3

syn-PLP- 147.37 (5.3) 146.19/154.49/123.44(4.8)/ 61.35 16.39

oxime 131.38 <8.7>/131.91

anti-PLP- 146.29 (<2.5)a 146.86/157.98/126.58a/ 61.24 16.22

oxime 133.44<8.7>/128.32
a
Coupling not resolved

The assignment of the two products as the syn and anti PLP-aldoxime pictured in Scheme

1 is in agreement with the results of APT and gHMBC experiments. Additionally, the

small, unresolved 1J(15N-13C) coupling in the C=N group of the anti isomer as well as the

coupling observed in the syn isomer (5.3 Hz) are in line with the couplings reported for

other oximes.15 The 15N chemical shifts of the syn and anti-PLP-aldoximes are δ = -26.5

and -15.1 ppm, respectively (relative to neat nitromethane in an external capillary). The

shift of syn-PLP-aldoxime agrees well with the shift reported for syn-benzaldoxime [δ = -

26.3, (E)-benzaldehyde oxime];16 in aliphatic aldoximes the anti isomers are also less

shielded than their syn counterparts (though the differences are smaller than observed

here).16

S7
NMR analysis of the reaction of PLP and DHAs

Three products (A, B, and C) are formed from the reaction between PLP and DHA; the

ratios of their abundance depend on the nature of the acid and change during the course

of the reaction (other factors, such as concentrations of components, ionic strength,

temperature, pD, etc. might also affect the ratio, but a detailed study of their influence

was beyond the scope of this work). In the case of 2 the 1H NMR signals (Table 3S) of

products A and B appear first, but during the course of the reaction product A is

converted completely into B and C. The final reaction mixture consists primarily of B

with 11 – 25 % of product C. In the case of 3 the reaction goes to completion in less than

a minute, and only products B and C can be observed (C constitutes 11 % of the total). In

contrast, the reaction of 1 with PLP is very slow; even after 12 days roughly 20% of the

PLP starting material remains. An intermediate product A is formed, but it has different

chemical shifts than intermediate A observed from the reaction of PLP with 2. Each of

the products (A, B, and C) has the same number of signals as the parent PLP, and the

signals are in the vicinity of those of PLP except for the CHO signal. Characteristic

features of the upfield lines of products B and C are their splittings due to 15N couplings
15
observed in experiments with N-enriched DHAs. These observations suggest that the

products retain the carbon skeleton of PLP and changes occur only at the aldehyde group.

From comparison with the spectral properties of syn- and anti-PLP-aldoxime, we

identified product B as the syn aldoxime and product C as the anti aldoxime (see Table

3S). The structure(s) of intermediate A were not further investigated.

S8
Table 3S. 1H NMR chemical shifts of products A, B, and C of the reaction of PLP with
15
N enriched 2 (as disodium salt at pD = 7.99, 8 mM concentration of PLP) and of
15
aldoximes resulting from reaction of PLP with NH2OH (pD = 8.04, 81 mM PLP).
15
Values in parenthesis/brackets are N-1H/31P-1H coupling constants in Hz, respectively.

The chemical shifts are referenced to the line of 1,4-dioxane at δ(1H) = 3.530.

Product CH=N C6-H (arom) 5-CH2O 2-CH3

Aa 8.77 (0b) 7.58 4.78[5.2] 2.53

B 8.397 (2.6) 7.667 4.699 [5.2] 2.231

C 7.767 (14.1) 7.566 –c 2.246

syn-PLP-oxime 8.278 (2.4) 7.584 4.623 [5.3] 2.075

anti-PLP-oxime 7.772 (14.0) 7.502 4.553 [5.3] 2.184


a
The lines shift to some extent in the course of the reaction. bSinglet. cNot visible due to

overlap with the residual solvent signal.

Concentration-dependence of the reaction: spectrophotometric analysis

PLP (0.1 mM) was incubated with various concentrations of hydroxylamine, 2, or 3.

The reactions were performed in 100 mM HEPES buffer, pH 8.0, and were allowed to

proceed for 30 min at ambient temperature. The absorbance at 388 nm, which

corresponds to unmodified PLP, was then recorded. The decrease in absorbance at 388

nm was plotted as a function of inhibitor concentration (mM) divided by PLP

concentration. The points were fitted using the “IC50 full 4 parameter” function in the

GraFit software package 17.

S9
As can be seen from Figure 1S, hydroxylamine is the most reactive with PLP, followed

by 3 and then 2. A 1:1 ratio of hydroxylamine:PLP is sufficient to convert almost all of

the PLP to aldoxime form, and 3 is only slightly less effective than hydroxylamine. In

contrast, an excess of 2 over PLP must be employed to result in substantial conversion to

the aldoxime form.

Figure 1S. Concentration-dependence of the reaction of PLP and DHAs monitored by

UV/vis spectroscopy.

UV/Vis data for compounds 4-6

UV/vis measurements were conducted as in the Experimental Section. Non-inhibitory

compounds 5 and 6 do not interact with PLP in solution, while addition of 4, which is

inhibitory, to PLP results in a spectral shift consistent with aldoxime formation.

S10
Figure 2S. Absorbance spectra of PLP and compounds 4, 5, 6 recorded at pH 8.0.

Hydrolysis of DHAs at pH 8.0 and 7.4

We performed hydrolysis experiments at pH 8.0 and pH 7.4 by detection of


18
hydroxylamine in solutions of hydroxamic acids as previously described. The method

exploits the use of picrylsulfonic acid, which reacts with primary amines at alkaline pH to

produce a species that absorbs light at 494 nm. This method is very sensitive and can

detect low micromolar concentrations of hydroxylamine; the major disadvantage is that

picrylsulfonic acid takes 15 min to react with hydroxylamine, so detailed kinetic studies

were not possible with this method. Briefly, 1 mL of 100 mM phosphate buffer, pH 8.0

or 7.4, containing 0.5 mM (1, 5, 6) or 0.05 mM (2, 3, 4) DHA (diluted from freshly

prepared 10 mM stocks) was prepared. The compounds were incubated at room

temperature for 0, 30, or 60 min, and then 15 µL picrylsulfonic acid (5 % w/v, Sigma

Aldrich) was added. Absorbance at 494 nm was recorded 15 min after addition of

picrylsulfonic acid.

Compounds 1, 5, 6 (non-inhibitory) did not show any significant absorbance even after 1

hour incubation at pH 8.0 or pH 7.4, indicating that no release of hydroxylamine

occurred. Incubation of 3 in buffer shows immediate evolution of hydroxylamine, and

S11
the absorbance does not substantially increase at subsequent time points. In contrast, 2

and 4 show a slow, steady increase in hydroxylamine evolution throughout the hour at

both pH 8.0 (Figure 3S) and 7.4.

Figure 3S. Evolution of hydroxylamine in buffered solutions of hydroxamic acids 2, 3,

and 4. Times are cumulative (0, 30, or 60 min in buffer plus 15 min reaction with

picrylsulfonic acid)

In order to support the conclusions of the picrylsulfonic acid experiments, we prepared

solutions of selected DHAs in pH 8.0 buffer and subjected them to ESI-MS after 1-2

days. In all cases, the DHA was the predominant species observed. However, in the case

of inhibitory compounds 2, 3, and 4, significant signals corresponding to the respective

monohydroxamic acids were visible, while we did not observe the corresponding

dicarboxylic acids. We also observed signals corresponding to the monohydroxamic

acids minus a water molecule, which could correspond to cyclic intermediates resulting

from intramolecular condensation (or could be ionization artifacts).

S12
References

1. Schraml, J.; Tkadlecova, M.; Pataridis, S.; Soukupova, L.; Blechta, V.; Roithova,

J.; Exner, O., Ring-substituted benzohydroxamic acids: 1H, 13C and 15N NMR spectra

and NH-OH proton exchange. Magn. Reson. Chem. 2005, 43 (7), 535-542.

2. Hauser, C.; Renfrow, W. J., Benzohydroxamic acid. Org. Synth., Colle. Vol. 1959,

2, 67.

3. Stolberg, M. A.; Mosher, W. A.; Wagner-Jauregg, T., Synthesis of a Series of

Vicinally Substituted Hydroxamic Acids. J. Am. Chem. Soc. 1957, 79 (10), 2615-2617.

4. Fritz, H. P.; von Stetten, O. E., 1H-NMR und IR-Spektren der

Formhydroxamsaure und ihres Natriumsalzes. Z. Naturforsch., B: Chem. Sci. 1969, 24b,

947-949.

5. Witherup, T. H.; Abbott, E. H., Carbon-13 nuclear magnetic resonance spectra of

the vitamin B-6 group. J. Org. Chem. 1975, 40 (15), 2229-2233.

6. Harruff, R. C.; Jenkins, W. T., A 13C n.m.r. study of the B6 vitamins and of their

aldimine derivatives. Organic Magnetic Resonance 1976, 8 (11), 548-557.

7. O'Leary, M. H.; Payne, J. R., 13C NMR spectroscopy of labeled pyridoxal 5'-

phosphate. Model studies, D-serine dehydratase, and L-glutamate decarboxylase. J. Biol.

Chem. 1976, 251 (8), 2248-2254.

8. Korytnyk, W.; Singh, R. P., Proton Magnetic Resonance Spectra of Compounds

in the Vitamin B6 Group. J. Am. Chem. Soc. 1963, 85 (18), 2813-2817.

9. Korytnyk, W.; Ahrens, H.; Donald, B. M.; Lemuel, D. W., Nuclear magnetic

resonance spectroscopy of vitamin B6. In Methods Enzymol., Academic Press: 1970; Vol.

Volume 18, Part 1, pp 475-483.

S13
10. Crépaux, D.; Lehn, J. M., Application des mesures de couplages 15N-H à la

détermination de configurations d'oximes. Organic Magnetic Resonance 1975, 7 (10),

524-526.

11. Pejkovic-Tadic, I.; Hranisavljevi-Jakovljevic, M.; Nescaroni, S.; Pascual, C.;

Simon, W., Protonenresonanzspektren von Oximen aromatischer Aldehyde. Helv. Chim.

Acta 1965, 48 (5), 1157-1160.

12. Jerslev, B.; Larsen, S., Hydrogen bonding and stereochemistry of ring-

hydroxylated aromatic aldehyde oximes. Crystal structures of three 4-OH-substituted

benzaldehyde oximes. Acta Chem. Scand. 1991, 45 (3), 285-291.

13. Phillips, W. D., Studies of hindered internal rotation in organic molecules by

nuclear magnetic resonance. Ann. N. Y. Acad. Sci. 1958, 70 (Nuclear Magnetic

Resonance), 817-832.

14. Sharif, S.; Huot, M. C.; Tolstoy, P. M.; Toney, M. D.; Jonsson, K. H. M.;

Limbach, H.-H., 15N Nuclear Magnetic Resonance Studies of Acid-Base Properties of

Pyridoxal-5'-Phosphate Aldimines in Aqueous Solution. The Journal of Physical

Chemistry B 2007, 111 (15), 3869-3876.


13
15. Kalinowski, H.-O.; S.Berger; Braun, S., C-NMR-Spektroskopie. Thieme:

Stuttgart, 1984.

16. Berger, S.; Braun, S.; H.-O.Kalinowski, NMR-Spektroskopie von Nichtmetallen 2.


15
N-NMR-Spektroskopie. Thieme: Stuttgart, 1992.

17. Leatherbarrow, R. J. GraFit Version 5, Erithacus Software Ltd.: Horley, UK,

2001.

S14
18. Pietta, P.; Calatroni, A.; Zio, C., Spectrophotometric determination of residual

hydroxylamine in hydroxamic acids. Analyst 1982, 107, 341-343.

S15

You might also like