ArticlePDF Available

Serine Transhydroxymethylase

Authors:

Abstract

Serine transhydroxymethylase was shown to catalyze the synthesis of l-serine from glycine and formaldehyde in the absence of tetrahydrofolate. Initial velocity, product inhibition, and exchange reaction studies showed that the addition of substrates was ordered, with glycine adding first. The effect of tetrahydrofolate on increasing the rate of serine synthesis was correlated with the effect of this coenzyme on the exchange of the α-hydrogen of glycine with protons of the solvent. The data were interpreted to mean that formaldehyde is transferred from serine to 5,10-methylene tetrahydrofolate through an intermediate enzyme-formaldehyde complex.
THE JOURNAL OF BIOLOGICAL CEIEMISTRY
Vol. 248, No. 10, Issue of May 25, pp. 3631-3635, 19i3
Printed in U.S.A.
Serine Transhydroxymethylase
A KITETIC STUDY OF THE SYn’THESIS OF SEILI?\‘E IS THI4: A13SE:SCI1: OF TETl~AHYDBOFOLATE*
(Received for publication, 1l)ecember 5, 1972)
RIING
S.
CHEN AND
LAVERXE
SCHIHCH
From the Department of Chemistry, Blqjjlox College, BluJlorr, Ohio -45817
SUMMARY
Serine transhydroxymethylase was shown to catalyze the
synthesis of L-serine from glycine and formaldehyde in the
absence of tetrahydrofolate. Initial velocity, product in-
hibition, and exchange reaction studies showed that the
addition of substrates was ordered, with glycine adding first.
The effect of tetrahydrofolate on increasing the rate of serine
synthesis was correlated with the effect of this coenzyme on
the exchange of the a-hydrogen of glycine with protons of the
solvent. The data were interpreted to mean that formalde-
hyde is transferred from serine to 5, IO-methylene tetrahy-
drofolate through an intermediate enzyme-formaldehyde
complex.
Serine transhydroxymethylase (EC 2.1.2.1.) catalyzes the
reversible interconversion of serine and glyciue as shown in Rexc-
tion 1.
Aerine + tetrahydrofolate ti glycine (1)
+ 5,10-methylene tetrahydrofolate
Serine transhydroxymethylase is a pyridoxal 5-phosphatc-cou-
taining enzyme, and most of our previous communications have
dealt with the role of this coenzyme in the enzymatic reaction
(l-4). We assumed that the role of tetrahydrofolatc w-as to
accept or donate a l-carbon moiety at the oxidation level of form-
aldehyde. During the past several years we have found that
serine transhydroxymethylase will cleave several /%hydrosy
amino acids to glycine and an aldehyde. Most of these reac-
tions showed no requirement for tetrahydrofolate. These reac-
tions include the cleavage of L-threonine and allothreoniue to
glycine and acetaldehyde, and erythro- and three-P-phenylserine
to glycine and benzaldehyde (5). The cleavage of Lu-methyl-
serine, however, did show a requirement for tetrahydrofolate (2).
Our understanding of the involvement of tetrahydrofolate in the
reaction could not explain the above observations.
We have now demonstrated that serine transhydrosymcthylase
also catalyzes the interconversion of serine and glycine in the
absence of tetrahydrofolate. In this paper we present evidence
for the synthesis of serine from glycine and formaldehyde. A
* This work was supported by Natioual Science Foundation
Grant GB-31780.
study of the kiuct,ic properties of this reaction has permitted us
to write a mcchxuism which partially explains the substrate
specificity of the enzyme and the role of tetrahydrofolate.
EXPERIMEXTAL PROCEDURE
Jlaterials-Gl?-tine, scriue, and tctrahgdrofolate were pur-
chased from Sigma Chemical Co. Allothreonine and N, N-
bis(2-hydrosyethyl)2-amiuoethane sulfouic acid were obtained
from Nutritional Riochcmicals Corp. [I-14C]Glycine and [2-3H]-
glycine were purchased from International Chemical and Nu-
clear Corp. Serine transhydroxymethylase was purified as
previously described (6).
Purification of [%311]G/gcine-Tritiated glycine was adjusted to
~1-1 2 with I-ICI aud placed ou a Dowex 50-W-X12 column (0.5 x
2.0 cm) which had beeu washed with 6
M
HCl and HZ0 until the
eluate was pII 2. The loaded column was washed with 0.01 M
HCl until no radioactivity could be detected in the eluate. The
glycine was then cluted from the column with 6 M IICl. The
HCl was removed by lyophilization and the resulting residue
was dissolved iu ILO. The lyophilization procedure was re-
peated two more times.
Tritium Exchange Studies-The rate of enzymatic catalysis
of the exchange of the a-hydrogen of glycine with protons of the
solvent was determined as follows. To 80 ~1 of [2-3H]glycine
(220,000 cpm), 7.5
InM
in 0.01
M
N,N-bi@hydroxyethyl)2-
aminoethaue sulfonic acid, PI-I 7.2, were added 10 ~1 (56 pg) of
cuzyme. The reaction, which occurred at 37”, was stopped by
the addition of 10 ~1 of 12y0 trichloroacetic acid. The dena-
tured protein was removed by centrifugation and 80 ~1 of the
supernatant solution were placed on a Dowes SO-W12 column
(0.5
x
4.0 cm), pH 2.0. The loaded column was washed with
0.01 hr IICl, aud the eluate containing the exchanged tritium
was collected in three counting vials, 0.4 ml per vial. Each
vial was filled with 10 ml of Bray’s (7) solvent containing 0.5y0
2,5-bis-2-(5-tert-butylbeuzoxazolyl) thiophene and counted in a
Tri-Carb 3310 scintillation counter.
Xerine and Threonine Synthesis-The synthesis of serine was
measured by the rate of couversion of [l-*4C]glycirie to [PC]-
serine. h typical expcrimcnt was performed as follows. To
40 ~1 of 0.01 M N, N-bis(2-hydroxyethyl)2-aminoethane sulfonic
acid (p1-T 7.2, aud 37”) were added [lJ*C]glycine and formalde-
hyde to final concentrations ranging from 1 to 8 mhf and 1 x lo5
to 8
x
IO5 cpm and 28 pg of enzyme. The reaction was stopped
by the additiou of 5 ~1 of 12% trichloroacetic acid. The dena-
tured proteiu was removed by centrifugation. -1liquots (10 ~1)
3631
This is an Open Access article under the CC BY license.
3632
were removed, spotted on sheets of Whatman No. 4 paper, and
subjected to high voltage electrophoresis (200 volts per cm) in
1.6
M
formic acid for 2 hours. Mixtures of glycine and serine
were used as markers. After they were sprayed with ninhydrin,
the areas corresponding to the serine marker were cut out and
placed in a counting vial with 20 ml of toluene containing 0.5%
2,5-his-2-(5.tert-butylbenzoxazolyl)thiophene. The radioactiv-
ity was counted as described above.
In several experiments acetaldehyde replaced formaldehyde.
Under these conditions the enzyme synthesizes threonine (or
allothreonine) (5). The extent of this reaction was determined
as with serine, except that threonine markers were used to locate
the product on the Whatman No. 4 sheets.
RESULTS
Se&e synthesis-Incubation of serine transhydroxymethylase
with glycine and formaldehyde was observed to produce a radio-
active compound which behaved as serine in electrophoresis
and paper chromatography experiments. This was concluded
from the following experiment. A reaction mixture of enzyme,
[l-1*C]glycine, and formaldehyde was incubated as described
under “Experimental Procedure.” Aliquots were spotted on
Whatman No. 4 paper and chromatographed, with pyridine-
water, 80:20, serving as the solvent system. The area of the
chromatogram corresponding to serine was found to contain a
measurable amount of a radioactive compound. This com-
pound was eluted from the paper and rechromatographed in
three solvents which totally or partially resolve serine and gly-
tine. These solvents were: butanol-acetic acid-water, 100 : 22 :
50; tert-butyl alcohol-acetone&pentanone-water-diethylamine,
40 : 20 : 20 : 20 : 4; phenol-water-acetic acid, 50 : 10 : 1. In each
case the radioactive compound migrated with serine. Because
the paper chromatography experiments are time-consuming, the
procedure used for separating glycine and serine for the kinetic
studies was high voltage electrophoresis in 1.6
M
formic acid.
With the experimental conditions employed, the synthesis of
serine was linear with time and enzyme concentration. Reac-
tion times were 30 min. The counts per min for the no-enzyme
control compound did not exceed 10% of the reaction counts
per min. Duplicate experiments were consistently within the
range of &2%.
The use of formaldehyde resulted in two limitations on reaction
conditions. First, formaldehyde inhibited the reaction at con-
cent.rations above 15 mM, which limited the range of concentra-
tions which could be studied. Second, since amino acids react
with formaldehyde (8), we avoided experiments in which the
concentrations of amino acid was more than one order of mag-
nitude greater than the formaldehyde concentration.
Initial Velocity Studies-The initial rate of serine synthesis
was determined at several concentrations of glycine and form-
aldehyde. Duplicate reaction mixtures were used for each con-
centration of substrate. Three 10.~1 aliquots were used from
each reaction mixture in the electrophoresis system. In Fig. 1
is shown a double reciprocal plot with glycine as the variable
substrate and formaldehyde as the changing fixed substrate.
From the crossover point the dissociation constant for glycine is
calculated to be 7
mM
which is in good agreement with the 6
InM
value obtained from spectral measurements (5). Fig. 2 shows
slope and intercept replots of the data in Fig. 1 versus the recipro-
cal of the formaldehyde concentration. From the intercept
replot the limiting Michaelis constant for formaldehyde was
calculated to be 5.6 mM. The slope replot appears to go through
the origin, indicating that at infinite formaldehyde concentra-
iHCHOJ mM /
(IIGLY CINE) x lO-2 M-’
FIG.
1. Double reciprocal plots of the initial rate of serine
synthesis as a function of glycine concentration at the indicated
concentrations of formaldehyde. Each reaction mixture con-
tained 28 pg of enzyme. Details are described under “Experi-
mental Procedure.”
“0
-
x
w
n
0
d
6
-i
m
4
L
0 2 6 6
(IC-lkiO~ x 10-2M-’
IO 12
FIG. 2.
Slope and intercept replots from Fig. 1. Slope replot,
0-O;
intercept replot, A---A.
tion the requirement for glycine disappears. This initial velocity
pattern is characteristic of equilibrium-ordered mechanisms
(9)
Fig. 3 shows the double reciprocal plot of the initial velocity,
with formaldehyde as the variable substrate and glycine as the
changing fixed substrate. The observation that all lines in-
tersect at the vertical axis is in agreement with the interpretation
of the data in Fig. 2 that, at infinite formaldehyde conccntra-
tion, the reaction is independent of the glycine concentration.
Product Inhibition-Figs. 4 and 5 show that serine is a com-
petitive inhibitor of glycine and a noncompetitive inhibitor of
formaldehyde in the synthesis of serine. These experiments
indicate that the addition of the substrates is ordered, with gly-
tine combining with the enzyme before formaldehyde. This
mechanism is different from the one indicated in the initial ve-
locity studies in Figs. 2 and 3 which supported an equilibriunl-
ordered addition. This mechanism predicts compet,ition by
serine with both glycine and formaldehyde (9).
Formaldehyde was also tested as an inhibitor of allothreonine
degradation to glycine and acetaldchyde. This reaction was
8
0 2 3 6 8 13 12
(I/HCHO) x IO“ M-’
FIG.
3. Double reciprocal plots of the initial rate of scrine syn-
thesis as a function of formaldehyde concentration at the indi-
cated concentrations of glycine. Details are described under
“Experimental Procedure.”
I I
0 2 4 6 8
(I/GLYCINE) x IO-*M-I
FIG.
4. Inhibition of serine synthesis by serine as a function of
glycine concentration. Each reaction mixture contained 40 fig
of enzyme and G rnM formaldehyde.
-I Q I 2 3 4 5
II-HCHO) x 10-2M-’
FIG.
5. Inhibition of serine synthesis by serine as a function of
formaldehyde concentration. Each reaction mixture contained
40 fig of enzyme and 1.8 mM glycine. Details are described under
“Experimental Procedure.”
assayed by converting the acetaldehyde to ethanol with DPNH
and alcohol dehydrogenase (5). Formaldehyde is a poor sub-
strate for alcohol dehydrogenase, thus the concentration range
that could be used in the inhibition studies was limited. We
found no inhibition of the allothreonine reaction by 1 mM form-
aldehyde at allothreonine concentrations as low as 0.25 m&I.
This finding suggests that formaldehyde cannot bind to the free
enzyme and it is in agreement with the ordered addition stated
above.
From the data in Fig. 4 the Zii for L-scrine was calculated to
3633
TABLE 1
E$ect of tetrahydrofolate on serine and threonine synthesis and
exchange of or-hydrogen of glycine with protons of the solvent
Reaction’”
-
I
Serine synthesis.
...............
Serine svnthesis. ...............
Glycine a-hydrogen exchange. .
Glycine or-hydrogen exchange.
Threonine synthesis. .
Threonine synthesis.
ntoles/min/mole
+ 80
- 0.5
+ 120
1.2
+ 10
- 7.0
a See “Experimental Procedure” for reaction conditions.
Tetrs-
hydrofolate Turnover number
be 0.8 mM, which is the same value we found previously by dif-
ferent methods (5). From the inhibition studies we calculated a
Km of 7.7
mM for formaldehyde.
Tritium Exchange Studies-We
have previously shown that
serine transhydroxymethylase catalyzes the exchange of the
a-hydrogen of glycine with protons of the solvent (3) in the
absence of an aldehyde. The rate of this exchange reaction
was enhanced severalfold by the addition of tetrahydrofolate.
These studies showed that the rate of exchange could be cor-
related with the presence of an enzyme-glycine complex which
absorbs at 495 nm (3, 4). Since tetrahydrofolate greatly ac-
celerates both the rate of synthesis of serine and the rate of ex-
change of the a-hydrogen of glycinc, it was of interest to com-
pare the effect of this compound on the two reactions. The
reactions were initiated as described under “Experimental Pro-
cedure.” The concentrations of reactants were the same for
both studies: glycine, 7.5 mnf; formaldehyde, 7.0 rnnf; and tet-
rahydrofolate, 0.2 mM. Since the formaldehyde concentration
in the serine synthesis reaction in the absence of tetrahydrofolate
is at its
K,
value, we multiplied the observed serine synthesis
rate by 2 so that we could compare the maximum rate with the
tritium exchange reaction rate which does not involve form-
aldehyde. The results are recorded in Table I and show that
the rate of exchange of the a-hydrogen of glycine is faster than
the rate of serine synthesis. This shows that the breaking of
the a-hydrogen-carbon bond of glycinc is not the rate-limiting
step in serine synthesis. Tetrahydrofolate enhanced the tritium
exchange and serine synthesis reactions by 100- and 160-fold,
respectively. The a-hydrogen exchange reaction probably ex-
hibits an isotope effect, and thcrc is no assurance that it is the
same in the presence and absence of tctrahydrofolate. However,
the enhancement of the rates of the two reactions is sim-
ilar enough to indicate that the role of tetrahydrofolate in the
two reactions may be due to the same effect.
The effect of tetrahydrofolate on the cleavage and synthesis
of allothreonine was also tested. We found that tetrahydrofolate
had no effect on the rate of cleavage of allothreonine but did
increase the rate of synthesis from glycine and acetaldehyde by a
factor of 1.4 (Table I). The rate of allothreonine synthesis in
the absence of tetrahydrofolate is 14 times greater than the rate
of serine synthesis (Table I).
Glycine-Serine Exchange-Even though L-serine is cleaved to
glycine and formaldehyde very slowly in the absence of tetra-
hydrofolate, it would be possible that the enzyme could catalyze
a rapid exchange between carbon atoms 1 and 2 of glycine and
serine. Reaction mixtures containing L-serine and glycine, at
concentrations equal to their
Km
values, and enzyme were incu-
bated at 37” for periods up to 4 hours in the presence of [VC]-
3634
glycine. So exchange between glycinc and serine could be
c
detected at enzyme levels which would show serine synthesis
from glycine and formaldehyde in a 5-min incubat,ion. These
data are consistent with an ordered mechanism.
DISCUSSIOS
The role of t,etrahydrofolate as a carrier of l-carbon groups is
well established. However, the very stable formation of the
tetrahydrofolate-formaldehyde complex, 5,10-metlrylene tct-
rahydrofolate, does not explain its apparent absolute rcquirc-
meut in the serine transl~ydroxymetl~ylase reaction (10). Pre-
vious studies which showed that tetrahydrofolate greatly accel-
erated the enzymic catalyses of the exchange of the a-protons
of glycine with the solvent produced the first evidence that this
compound may have another role in addition to the one of acting
as the formaldehyde carrier (4). The observation that serine
transhydrosymethylase would cleave several P-hydroxy amino
acids in the absence of tetrahydrofo1at.e added to the mystery
of the role of this coenzyme in the serine-glycine reaction (5).
We have now shown that serine transhydroxymethylase also
catalyzes the reversible interconversion of serine and glycine in
the absence of tetrahydrofolate. The reaction rates are very
slow, but the synthesis of serine was sufficiently fast to permit a
kinetic study of this reaction. We feel that the data help to
explain the role of tetrahydrofolate in the serine transhydrosy-
methylase reaction.
One of the key questions concerning tetrahydrofolnte
is whether it act,s as the direct l-carbon acceptor from serine or
whether formaldehyde is transferred through an intermediate
enzyme-formaldehyde complex. These two mechanisms are
shown schematically in Scheme 1 (S, serine; G, glgcine; F, form-
aldehyde; THF, tetrahydrofolate; 5,10-THF, 5, lo-methylene
tetrahydrofolatc). Scheme la shows a direct transfer of the
l-carbon moiety between serine and tetrahgdrofolate, while
Scheme lb shows the transfer through an intermediate enzyme-
formaldehyde complex.
1 THF S j + l&GiG - -iq (la)
hiF I
ii1
+ I THF k --cl * 1 5, IO-THF I /
G1 (I@
ScrreME 1
The observation that tetrahydrofolate is not absolutely required
in the reactions catalyzed by serine transhydroxymetl~ylase
favors Scheme lb as the mechanism. Scheme lb also is more
consistent with data presented in this paper.
In Scheme 2 is depicted a mechanism which we feel best fits
the experimental data published in this and previous manuscripts
(l-6). In t’his mechanism, L-serine forms a Schiff base complex
with the enzyme (Structure 1) which is cleaved to an enzyme-
formaldehyde-glycine complex (Structure 2). The next step
in the reaction is the dissociation of formaldehyde, which we
now know occurs at an extremely slow rate, i.e. turnover number
SCHEME 2
of about 0.004. Since glycine cannot dissociate prior to form-
nldehydc dissociation it cannot equilibrate with glycine in solu-
tion, thus explaining the failure to observe exchange of L-serine
with [‘“C]glycinc. When allothreonine or threonine replaces
L-serine, lnt~ermediate 2 would have enzyme-bound acetaldehyde
which, because of either its difference in chemical reactivity or
its shape, can dissociate from the enzyme, leaving a dissociable
enzymeglycine complex. fl-Phenylserine would generate an
enzyme-benzaldehyde complex in Structure 2. This large bulky
structure apparently dissociates from the enzyme rapidly, since
this compound is cleaved by the enzyme even more rapidly than
serine (5).
Scrinc synthesis was detected because k, is large enough that
the reaction of the enzyme-glycine complex with formaldehyde
occurs at a measurable rate. The mechanism is more complex
than that shown in Scheme 2 since at least four enzyme-glycine
compleses have been identified spectrally (6) and kinetically
(II). In Scheme 2 we show only two of the complexes with
formaldehyde binding to the enzyme-glycine complex in which
the glycine has lost its a-proton. There is no evidence that
formaldehyde binds to this particular intermediate, however.
The breaking of the oc-proton-carbon bond of glycine might be
expected to be the rate-limiting step, but the data in Table I
indicate that this is not so since serine synthesis is slower than
tritium exchange. This suggests that the rate-limiting step
somehow involves formaldehyde.
Cheng and Haslam (11) have recently shown that the forma-
tion of the enzyme-glycine complex has a rate constant of about
104. This would be several orders of magnitude faster than
the rate of synthesis of serine in the absence of tetrahydrofolate.
It is not surprising, then, that the initial velocity patterns sug-
gest that the enzyme-glycine complex is at thermodynamic equi-
librium. As Cleland (9) points out, under these conditions
the product serine should behave as a competitive inhibitor of
both glycine and formaldehyde even though the reaction is or-
dered. In this system we find that serine is a noncompetitive
inhibitor of formaldehyde. This means that at saturating levels
of formaldehyde t,he concentration of free enzyme is not zero.
This would be true in Scheme 2 if the addition of formaldehyde
to the enzyme-glycine complex was slow and the enzyme was
not saturated with glycine. The data for Figure 5 were obtained
at 1.8 IBM glycine, which is well below the
K,
value of 6 mM for
glycine.
Wllen
we repeated this experiment at 20 mM glycine
and 10 mnr serine, the reciprocal plots deviated from linearity.
The deviation could be explained by assuming a nonenzymatic
reaction of glycine and serine with formaldehyde to the extent
that it changed the formaldehyde concentrations. We could
not minimize this effect by raising the formaldehyde concentra-
tions since this would have led to formaldehyde inhibition.
The observation that tetrahydrofolate accelerates only the
serine-glycine reaction suggests that its role is to catalyze the
removal or addition of formaldehyde from the active site. Since
the enzyme in the absence of tetrahydrofolate degrades allo-
threonine and P-phenylserine at rates comparable to serine in the
presence of tetrahydrofolate, it is apparent that the role of this
coenzymc is not crucial to the breaking of the LY-/3 carbon bond
of the @-hydroxy amino acid (5). Formaldehyde is known to
react readily with primary and secondary amines to form stable
hydroxymethylamines (8). If in the cleavage of serine such
an intermediate was generated, it would tend not to break down
to free formaldehyde and enzyme. One way of removing the
formaldehyde from the enzyme would be to transfer the form-
aldehyde to an even more stable form. The formation of 5, lo-
3635
methylene tetrahydrofolate with an equilibrium constant of
3.2 x lo4 would serve this function very well (10). One unan-
swered question concerning tetrahydrofolate is the relationship
of the effects of this coenzyme on the tritium exchange reaction
and the serine synthesis reaction. It is intriguing to speculate
that both effects are attributable to a single chemical interaction
between tetrahydrofolate and the enzyme. Also, it should be
possible to determine which of the four known enzyme-glycine
complexes acts as acceptor of formaldehyde.
REFERENCES
1. SCHIRCH,
L.,
AND MASON,
M. (19G2)
J. Biol. Chem. 237, 2578
2. SCHIRCH,
L. G.,
AND MASON,
M. (1963)
J. Biol. Chem. 238, 1032
3.
SCI~IRCH,
L.,
AND JENIUXS,
W. T. (1964) J.
Biol.
Chem. 239,
3797
4. SCHIRCEI,
L.,
AND JENIUNS,
W. T. (1964) J.
Biol. Chem. 239,
3801
5. SCHIRCH,
L.,
AND GROSS,
T. (1968)
J. Biol. Chem. 243,
5651
6. SCHIRCH, L., AND DILLER,
A. (1971) J.
Biol. Chem. 246,
3961
7.
BRAY, G.
A. (1960)
Anal. Biochem. 1,
279
8. KALLXN,
R. G.,
AND JENCICS,
W. P. (1966) J. Biol.
Chem.
241,
5864
9.
CLELAND,
W. W. (1970) in
The Enzymes (BOYER,
P., ed) Vol 2,
3rd Ed, p. 1, Academic Press, New York
10.
KALLEN,
R.
G., AND JENCICS,
W. P. (1966)
J. BioZ. Chem.
241.
5851
11. CIIENG,
C. H.,
AND HASLIM,
J. L. (1972)
Biochemistry 2, 3413
... GS is a key assimilatory enzyme for ammonia 59 , primarily responsible for scavenging ammonia, a highly reactive and cytotoxic metabolite, and converts it to glutamine 60 . SHMTs, along with GDC (glycine decarboxylase complex), being involved in the transfer of hydroxymethyl of serine to typically polyglutamylated tetrahydrofolate producing Gly and 5,10-methylene THF 61 , while amino methyltransferase is involved in the glycine catabolism. ...
Article
Full-text available
In plants, during growth and development, photoreceptors monitor fluctuations in their environment and adjust their metabolism as a strategy of surveillance. Phytochromes (Phys) play an essential role in plant growth and development, from germination to fruit development. FR-light (FR) insensitive mutant (fri) carries a recessive mutation in Phytochrome A and is characterized by the failure to de-etiolate in continuous FR. Here we used iTRAQ-based quantitative proteomics along with metabolomics to unravel the role of Phytochrome A in regulating central metabolism in tomato seedlings grown under FR. Our results indicate that Phytochrome A has a predominant role in FR-mediated establishment of the mature seedling proteome. Further, we observed temporal regulation in the expression of several of the late response proteins associated with central metabolism. The proteomics investigations identified a decreased abundance of enzymes involved in photosynthesis and carbon fixation in the mutant. Profound accumulation of storage proteins in the mutant ascertained the possible conversion of sugars into storage material instead of being used or the retention of an earlier profile associated with the mature embryo. The enhanced accumulation of organic sugars in the seedlings indicates the absence of photomorphogenesis in the mutant.
Article
Hydrogen stable isotope analyses (δ2H) of plant derived organic compounds are a useful tool for ecological, environmental, and palaeoclimatological research. However, during organic compound synthesis, variable biosynthetic 2H-fractionation has been suggested to occur as a result of changes in plant carbon fluxes. So far, inference has been based on examining the δ2H patterns of plant compounds along environmental gradients, among plant species, and between plant organs. In an alternative approach, we used four plant species with four different types of mutations that cause impaired starch synthesis to assess whether variability in carbon metabolism affects the biosynthetic 2H-fractionation during cellulose, phytol, and acetogenic lipid synthesis. We found that mutants with impaired starch synthesis always had higher cellulose and phytol δ2H values compared to the wild type. By contrast, 2H-fractionation during acetogenic lipid biosynthesis generally did not show strong metabolic sensitivity. We rationalise these differences by considering the biosynthetic pathway of each compound and the likely source of the variable isotope effects. In different organic compounds, the sensitivity of variable biosynthetic 2H-fractionation to changes in C-metabolism depends on incorporation of specific H atoms from precursor molecules. As such, we determined that the similar increase in cellulose and phytol δ2H values as an effect of impaired starch synthesis most likely originates in triose-phosphates.
Article
Full-text available
2-Keto-4-hydroxyglutarate aldolase (2-oxo-4-hydroxyglutarate glyoxylate-lyase:2-oxo-4-hydroxyglutarate ⇌ pyruvate + glyoxylate), which catalyzes a terminal step in the catabolism of l-hydroxyproline by mammals, selectively interacts with formaldehyde; a Schiff-base complex is formed with apparently the active site lysyl residue of the enzyme. This was established by showing that (a) aldolase activity is destroyed when the enzyme is treated with NaBH4 in the presence of formaldehyde; (b) the stoichiometry of binding (formaldehyde to enzyme) is 1:1 on a molar basis; (c) formaldehyde competes with the binding of pyruvate or glyoxylate in the presence of NaBH4; and (d) the isolated reducedenzyme-formaldehyde adduct is N⁵-methyllysine. Advantage was taken of this finding to show that the irreversible inactivation of 2-keto-4-hydroxyglutarate aldolase by cyanide in the presence of formaldehyde (or, by inference, with any one of a number of other aldehydes) is due to the formation of an aminonitrile. This conclusion is based upon the demonstration that the mole per mole stoichiometry of the complex formed with the enzyme is 1:1:1 (aldolaseformaldehyde-cyanide) and that N⁶-carboxymethyllysine can be isolated after strong acid hydrolysis of the enzyme-[¹⁴C]-formaldehyde-cyanide adduct.
Chapter
It was earlier asserted that all enzymes can rightly be regarded as transferases, in the sense that they all catalyze the transfer of some fragment—an electron, a proton, a group, or even the whole—of the donor substrate to an acceptor (1). For convenience of classification, however, the Enzyme Commission of the International Union of Biochemistry has sorted the enzymes into six major classes, only one of which is officially termed “transferases” (2). These (official) transferases are group transferases. The present chapter presents a sampling of such transferases, with examples drawn from nearly all of the subsubclasses (categories) of these enzymes. The presentations are concise, and highlight mainly those aspects of each enzyme which have to do with covalent catalysis. It will be seen that group transfer is never a direct one between donor and acceptor. Always the enzyme acts to receive a group from the donor and pass it on to the acceptor. Some of the enzymes chosen for exposition have a long-familiar and accepted mode of action, and are included for the sake of completeness. But others are, at the time of writing, frankly controversial, there being no consensus as to their chemical mechanism. They are designedly included here in order to stress the issues which are in dispute.
Article
Pyridoxal 5′-phosphate (PLP) is the biologically active form of B6 vitamers. PLP plays the role as coenzyme in more than 160 different enzymes acting on a variety of substrates and inhibitors predominantly possessing amine and carboxylate moieties. The key feature of PLP-dependent enzymes is the catalytic versatility, carrying out transamination, decarboxylation, and beta and gamma elimination and substitution via several catalytic intermediates. Catalysis frequently accompanies a conformational transition between an open, less active state and a closed, fully active state. Recent investigations on functional genomics of PLP-dependent enzymes and the search of novel targets for therapeutic activity indicate that this class of enzymes has not been fully explored in the past and calls for more intensive and focussed studies.
Chapter
In devising its synthetic strategies Nature has developed a number of molecules, enzyme substrates or cofactors, which are remarkable in the diversity of different reaction paths open to each of them. Probably the most versatile of these is pyridoxal phosphate (PLP), the essential cofactor of amino acid metabolism. This compound can initiate reactions that may lead to the cleavage of any of the four bonds at the a-carbon, to electrophilic or nucleophilic reactions at the P-carbon and even to bond cleavage and formation at the y-carbon of a-amino acids. Thanks to this chemical versatility, PLP plays a pivotal role in connecting carbon and nitrogen metabolism, in the formation of biogenic amines, in providing an entry into the ‘one-carbon pool’ and in a number of other important processes. Underlying this multitude of different reactions is a simple, common mechanistic principle. The cofactor forms a Schiff‘s base with the amino group of the substrate and then acts as an electron sink transiently storing electrons which are freed in the cleavage of bonds until ,they are claimed again in a new bond-forming step.
Article
1-(N,N-Dimethylaminomethyl)-2-formylcymantrene (AFCMT) has been resolved into enantiomers through an intermediate formation of diastereomeric complexes with (S)-Ala-(S)-Ala, (S)-Ala-Gly and Gly-(S)-Ala. By the X-ray anomalous dispersion method the absolute configuration of its enantiomers has been determined: (-)436 AFCMT-(S), (+)436 AFCMT-(R).Alkylation of enantiomeric complexes (R)-and (S)-AFCMT-(GlyGly) Cu(II) with acetaldehyde gives, respectively, (R)-and (S)-Thr with an asymmetric yield of 92–98% and (R)- and (S)-allo-Thr with an asymmetric yield of 95–100%, only the N-terminal glycine being alkylated.The AFCMT enantiomers were also employed for retroracemisation of (R,S)-Ala-(R,S)-Nva; in this case an excess of (S)-Ala and (R)-Nva is obtained for (S)-AFCMT. (R)- and (S)-AFCMT are not liable to racemisation in the course of the threonine synthesis and retroracemisation of depeptides and can be repeatedly employed for these transformations.
Chapter
Introduction Properties of the Mammalian Enzyme Mechanism Studies Control Mechanisms
Article
Four types of threonine aldolases (TAs) with different stereospecificities were tested on the aldol synthesis of phenylserine (PS) starting from benzaldehyde and glycine under kinetic and thermodynamic control. At short time of reaction the chosen enzymes show different stereoselectivities (l-syn, l-anti, d-syn) compared to the reaction at equilibrium where syn-PS is obtained as the major product (d.e. ∼20%) for all types of TAs. A new aspect of the catalytic mechanism involved and a sight on the relative energy barriers for the possible rate-determining steps were obtained, based on data from the 13C-label distribution between components at the conditions close to equilibrium.
Article
Full-text available
A modified method for the purification of serine transhydroxymethylase from rabbit liver is presented. The purified enzyme is not only shown to catalyze the cleavage of serine but in addition, the cleavage of L-threonine and DL-allothreonine to glycine and acetaldehyde. The last two reactions had previously been thought to have been catalyzed by two separate enzymes, i.e. threonine aldolase and allothreonine aldolase. The evidence in support of a single enzyme catalyzing the three reactions includes the constant ratio for the specific activities during purification, similarity in affinity constants for substrates and inhibitors, and the ability of D-alanine to remove pyridoxal phosphate from the active site by transamination for both the allothreonine and serine reactions.
Article
Full-text available
A number of substrate analogues of serine transhydroxymethylase were tested as competitive inhibitors and substrates. Our previous results show that l-serine, l-threonine, allothreonine, and d-alanine serve as substrates. In this study, we found that threo- and erythro-β-phenylserine were cleaved to benzaldehyde and glycine. Tetrahydrofolate was not required for this reaction. Substrate analogues which contained an amino group adjacent to a carboxyl, phosphonate, or sulfonate group were bound by the enzyme. Analogues which lacked either the amino group or a carboxyl group of the amino acid had little if any affinity for the active site. The only exceptions to this were several sulfhydryl compounds. Several d-amino acids were competitive inhibitors with d-cysteine being the most effective. This shows that the configuration around the α-carbon of the amino acid is not critical for substrate binding. The variation of Vmax and Km with pH for the substrate allothreonine was determined. Both Vmax and Km were found not to change from pH 5.8 to 7.6. However, an anion binding site on the enzyme with a pK of 6.2 was found. This site showed little specificity for anions with all anions tested serving as competitive inhibitors. The enzyme-glycine complex exhibits absorption maxima at 343, 425, and 495 mµ. The equilibrium between the complexes absorbing at 343 and 425 mµ was found to be sensitive to both temperature and pH. A group on the enzyme-glycine complex with a pK of 6.9 was found to be involved. The temperature sensitivity was observed to be due in part to a large increase in entropy, which suggests that conformational changes in the protein are also involved in the interconversion of the enzyme-glycine complexes.
Article
Full-text available
The existence of protonated hydroxymethylamines of secondary amines, R2(+)/NHCH2OH, in equilibrium with hydroxy-methylamines and protonated amines, has been demonstrated by titration at equilibrium and by rapid titration of the protonated hydroxymethylamines. Analogous compounds, R3(+)/NCH2OH, are formed from formaldehyde and the sterically favorable tertiary amines: pyridine, N-methylimidazole, and triethylenediamine. The equilibrium constants for the formation of these cationic hydroxymethylamines are two to three orders of magnitude less favorable than for the formation of neutral hydroxymethylamines; conversely, the basicity of hydroxymethylamines is two to three orders of magnitude less than that of the parent amines. These differences are ascribed primarily to solvation and polar effects. Protonation of the dihydroxymethyl adducts of primary amines was not detected. An analysis of the complex equilibria for the addition of protons and hydroxymethyl groups to imidazole suggests that protonation and hydroxymethylation do not occur on the same nitrogen atom. The equilibrium constant Kd = [R2NCH2NR2]/[R2NH]2[H2C(OH)2] for the formation of the methylenediamine adduct of morpholine has been shown to be 15,500 ± 600 by two titrimetric methods. An attempt is made to evaluate the effects on the formol titration of changes in water concentration, formaldehyde polymerization, the methanol introduced with commercial formalin, and nonspecific medium effects in concentrated formaldehyde solutions. It is concluded that these factors either have little significant effect or largely cancel each other out.
Article
Full-text available
The bell-shaped pH rate profile for the reaction of tetrahydrofolic acid with formaldehyde to form 5,10-methylene tetrahydrofolic acid has been shown to result from a change in rate-determining step with changing acidity. In alkaline solution, the rate-determining step is the acid-catalyzed dehydration of hydroxymethyl tetrahydrofolic acid and its conjugate base. This step is subject to general acid catalysis with a Brönsted slope, α, of approximately 0.75. In acid solution, attack of tetrahydrofolic acid on formaldehyde is rate-determining. This step is subject to catalysis by general acids, including the solvated proton, with a Brönsted slope of 0.20. At pH values above 4, the reaction is inhibited by thiols, which combine with formaldehyde in a base-catalyzed reaction to form hemithioacetals. Depending on the reaction conditions, this inhibition can result from a lowering of the equilibrium concentration of formaldehyde or from a trapping by thiol anion of unhydrated formaldehyde, as it is formed from formaldehyde hydrate. The equilibrium constant for hemithioacetal formation from formaldehyde and mercaptoethanol is approximately 620 m⁻¹. Secondary amines such as morpholine and imidazole catalyze the reaction of tetrahydrofolic acid with formaldehyde by nucleophilic catalysis by a pathway which must involve the intermediate formation of cationic imines, [see PDF for structure]
Article
A modification of the naphthalene-dioxane-PPO liquid scintillator has been described which will allow up to 3.0 ml of an aqueous solution to be counted. The efficiency of this method in the presence of 1.0 ml water is 65.6% for C14 and 11.7% for tritiated water. This liquid scintillator has been used to count carbon-14 and tritium in urine, plasma, and liver homogenate. The effect of isotope concentration, solute concentration, and the presence of acids and bases on the count rate has been investigated.