scholarly journals Hyperoxidation of Peroxiredoxins 2 and 3

2013 ◽  
Vol 288 (20) ◽  
pp. 14170-14177 ◽  
Author(s):  
Alexander V. Peskin ◽  
Nina Dickerhof ◽  
Rebecca A. Poynton ◽  
Louise N. Paton ◽  
Paul E. Pace ◽  
...  

Typical 2-Cys peroxiredoxins (Prxs) react rapidly with H2O2 to form a sulfenic acid, which then condenses with the resolving cysteine of the adjacent Prx in the homodimer or reacts with another H2O2 to become hyperoxidized. Hyperoxidation inactivates the Prx and is implicated in cell signaling. Prxs vary in susceptibility to hyperoxidation. We determined rate constants for disulfide formation and hyperoxidation for human recombinant Prx2 and Prx3 by analyzing the relative proportions of hyperoxidized and dimeric products using mass spectrometry as a function of H2O2 concentration (in the absence of reductive cycling) and in competition with catalase at a fixed concentration of H2O2. This gave a second order rate constant for hyperoxidation of 12,000 m−1 s−1 and a rate constant for disulfide formation of 2 s−1 for Prx2. A similar hyperoxidation rate constant for Prx3 was measured, but its rate of disulfide formation was ∼10-fold higher, making it is more resistant than Prx2 to hyperoxidation. There are two active sites within the homodimer, and at low H2O2 concentrations one site was hyperoxidized and the other present as a disulfide. Prx with two hyperoxidized sites formed progressively at higher H2O2 concentrations. Although the sulfenic acid forms of Prx2 and Prx3 are ∼1000-fold less reactive with H2O2 than their active site thiols, they react several orders of magnitude faster than most reduced thiol proteins. This observation has important implications for understanding the mechanism of peroxide sensing in cells.

Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 2029-2029
Author(s):  
S. Paul Bajaj ◽  
Amanda Sutton ◽  
Sreejesh Shanker ◽  
Amy E Schmidt ◽  
Sayeh Agah ◽  
...  

Abstract Factor VII (FVII) consists of an N-terminal γ-carboxyglutamic acid (Gla) domain followed by two epidermal growth factor-like (EGF1 and EGF2) domains and the C-terminal protease domain. Activation of FVII results in a two-chain FVIIa molecule consisting of a light chain (Gla-EGF1-EGF2 domains) and a heavy chain (protease domain) held together by a single disulfide bond. The complex of tissue factor (TF) and FVIIa activates FIX and FX during coagulation. FVIIa on its own is structurally more “zymogen-like” and when bound to TF it is more “active enzyme-like.” We obtained crystal structures of EGR-VIIa/soluble (s) TF (2.9 Å resolution), dansyl-EGR-VIIa/sTF (1.9 Å resolution) and benzamidine-VIIa/sTF (1.6 Å resolution). We also investigated the effect of TF binding on the S1, S2, and S3/S4 subsites (Schechter and Berger, BBRC, 27:157-162, 1967) in FVIIa. The affinity of variously inhibited FVIIa to sTF was also measured using Biacore technology. For obtaining second order inhibition rate constants, FVIIa ± soluble (s) TF was incubated with each inhibitor for various times, diluted several fold and assayed for the residual FVIIa activity. The second order rate constants were obtained by plotting the first order rate constants versus the inhibitor concentrations. These data are summarized in the table below. From these data it appears that all subsites are affected upon FVIIa binding to sTF. Since in the crystal structure of EGR-VIIa/sTF the P1 Arg residue is the only residue that makes contact with FVIIa, it follows that the S1 site is affected ~10-fold upon binding to sTF. Adding a dansyl group that partially occupies the S3/S4 position (1.9 Å structure) increases the second order rate constant 7-fold (2.41 versus 0.35) over that of EGR-ck. Moreover the addition of Pro (DFPR-ck) or Phe (DFFR-ck) residue occupying the S2 position increases the second order rate constant 357-fold and 1500- fold, respectively (125 and 525 versus 0.35). Thus, comparison of dEGR, DFPR, DFFR inhibition suggests that FVIIa prefers Phe at S2 and at S3/S4 positions, and that TF opens up the S1/S2/S3/S4 sites for substrate or inhibitor occupancy. These data are consistent with LTR (P3/P2/P1) residues in FX at its activation cleavage site as well as with LTR (P3/P2/P1) residues and FTR (P3/P2/P1) residues at the 145-146 and 180-181 FIX activation cleavage sites, respectively. Thus, these studies with chloromethylketone inhibitors have biologic relevance. For Biacore studies, sTF was amine coupled to a CM5 chip. The binding of unoccupied active site FVIIa in 5 mM calcium to sTF was characterized by a KD of 7 nM. Benzamidine (10 mM)-VIIa, p-aminobenzamidine (pAB, 1 mM)-VIIa, EGR-VIIa, dEGR-VIIa, DFPR-VIIa and DFFR-VIIa each bound to sTF with KD values ranging from 1- 2 nM. These affinity measurements indicate that the S1 site occupied FVIIa molecule (benzamidine-FVIIa, pAB-VIIa) has essentially the same conformation as the S1/S2/S3/S4 occupied FVIIa. This conclusion is consistent with similar crystal structures of variously inhibited FVIIa molecules complexed with sTF. The differential rates of incorporation of various chloromethylketone inhibitors could be due to the interaction of various residues (P1, P2, P3, P4) with the corresponding active subsites (S1/S2/S3/S4) of FVIIa. Additionally, the rate of incorporation of chloromethylketone inhibitors into FVIIa also involves the irreversible alkylation step, which could be faster for DFFR-ck and DFPR-ck. Once these inhibitors are incorporated, it appears that they induce the same conformation in FVIIa as achieved by S1 site occupancy alone. Thus S1 site occupancy in FVIIa induces the required conformation to modestly increase the affinity for TF. Second Order Rate Constants for Inhibition of FVIIa ± sTF with Various Chloromethylketone (ck) Inhibitors Inhibitor Minus sTF k (min−1 mM−1) Plus sTF k (min−1 mM−1) Fold Difference EGR-ck 0.04 0.35 8.8 dansyl EGR-ck 0.07 2.41 34.4 (D)FPR-ck 2.3 125 54.3 D)FFR-ck 5.6 525 93.8


1991 ◽  
Vol 69 (2) ◽  
pp. 363-367
Author(s):  
Guoying Xu ◽  
Jan A. Herman

Ion/molecule reactions in mixtures of ethyl chloride with C1–C4 alkylamines were studied by ICR mass spectrometry. Ethyl cation transfer to C1–C4 alkylamines proceeds mainly through diethylchloronium ions with rate constants ~3 × 10−10cm3 s−1. In the case of s-butylamine the corresponding rate constant is 0.5 × 10−10 cm3 s−1. Key words: ICR mass spectrometry, ion/molecule reactions, ethylchloride, methylamine, ethylamine, propylamines, butylamines


2018 ◽  
Vol 62 (6) ◽  
Author(s):  
Krisztina M. Papp-Wallace ◽  
Melissa D. Barnes ◽  
Jim Alsop ◽  
Magdalena A. Taracila ◽  
Christopher R. Bethel ◽  
...  

ABSTRACT The imipenem-relebactam combination is in development as a potential treatment regimen for infections caused by Enterobacteriaceae possessing complex β-lactamase backgrounds. Relebactam is a β-lactamase inhibitor that possesses the diazabicyclooctane core, as in avibactam; however, the R1 side chain of relebactam also includes a piperidine ring, whereas that of avibactam is a carboxyamide. Here, we investigated the inactivation of the Klebsiella pneumoniae carbapenemase KPC-2, the most widespread class A carbapenemase, by relebactam and performed susceptibility testing with imipenem-relebactam using KPC-producing clinical isolates of Enterobacteriaceae . MIC measurements using agar dilution methods revealed that all 101 clinical isolates of KPC-producing Enterobacteriaceae ( K. pneumoniae , Klebsiella oxytoca , Enterobacter cloacae , Enterobacter aerogenes , Citrobacter freundii , Citrobacter koseri , and Escherichia coli ) were highly susceptible to imipenem-relebactam (MICs ≤ 2 mg/liter). Relebactam inhibited KPC-2 with a second-order onset of acylation rate constant ( k 2 / K ) value of 24,750 M −1 s −1 and demonstrated a slow off-rate constant ( k off ) of 0.0002 s −1 . Biochemical analysis using time-based mass spectrometry to map intermediates revealed that the KPC-2–relebactam acyl-enzyme complex was stable for up to 24 h. Importantly, desulfation of relebactam was not observed using mass spectrometry. Desulfation and subsequent deacylation have been observed during the reaction of KPC-2 with avibactam. Upon molecular dynamics simulations of relebactam in the KPC-2 active site, we found that the positioning of active-site water molecules is less favorable for desulfation in the KPC-2 active site than it is in the KPC-2–avibactam complex. In the acyl complexes, the water molecules are within 2.5 to 3 Å of the avibactam sulfate; however, they are more than 5 to 6 Å from the relebactam sulfate. As a result, we propose that the KPC-2–relebactam acyl complex is more stable than the KPC-2–avibactam complex. The clinical implications of this difference are not currently known.


1976 ◽  
Vol 153 (2) ◽  
pp. 495-497 ◽  
Author(s):  
D C Wilton

The enzyme deoxyribose 5-phosphate aldolase was irreversibly inactivated by the substrate analogue acrolein with a pseudo-first-order rate constant of 0.324 min-1 and a Ki (apparent) of 2.7 × 10(-4) m. No inactivation was observed after prolonged incubation with the epoxide analogues glycidol phosphate and glycidaldehyde. It is suggested that the acrolein is first activated by forming a Schiff base with the enzyme active-site lysine residue and it is the activated inhibitor that reacts with a suitable-active-site nucleophile.


1979 ◽  
Author(s):  
J.H. Griffin ◽  
G. Beretta

Previous 3H-DFP studies showed that HF binding to kaolin does not result in formation of new active sites. New studies show that single chain zymogen HF, in solution or surface bound, reacts slowly with DFP causing loss of coagulant activity and uptake of 1 mol DIP/mol HF with a 2nd order rate constant, k2=0.4 M-1 min-1. High MW kininogen (HMWK) did not affect the rate constant for this reaction. PK reacts with DFP causing loss of coagulant activity with k2=0.6 M-1min-1. These values are ~103lower than for the activated enzymes (160 and 500 M-1 min-1 for HFa and kallikrein respectively) and are similar to values for trypsinogen that exhibits weak enzymatic activity. HF, PK, HMWK, and kaolin were separately preincubated with 40mM DFP for 5 min to inhibit traces of active enzymes. Then, mixing of these reagents in the presence of 40mM DFP caused a burst of cleavage of 131I-HF and 125I-PK, the extent of which depended on the amount of HMWK. To avoid reciprocal proteolysis of HF and PK, “killed” zymogens (DIP-HF or DIP-PK formed by 48 hr incubation with 40mM DFP) were used in place of HF or PK. Controls showed that PK and DIP-PK were similarly cleaved by purified HFa. If DIP-131I-HF or DIP-125I-PK was substituted for HF or PK in mixtures of HF, PK, HMWK, and kaolin, no detectable cleavage occurred. These data allow that inherent activity of single chain forms of HF or PK may trigger-surface dependent reactions, but the vast majority of molecules are activated by reciprocal proteolysis between HF and PK with HMWK as a cofactor.


1986 ◽  
Vol 39 (8) ◽  
pp. 1257 ◽  
Author(s):  
NL Arthur ◽  
JC Biordi

Rate constants for the recombination of CH3 radicals have been measured by means of the rotating sector technique in the temperature range 373- 463 K, and at a pressure of 30 Torr . CH3 radicals were produced by the photolysis of acetone, and the experimental data were fitted to sector curves generated from Shepp's theory. The results give kb = (2.81�0.22)×1013 cm3 mol-1 s-1, which, under the chosen experimental conditions, is close to its high-pressure limiting value. A comparison is made with the other values of the rate constant reported in the literature, and a best value is suggested.


1984 ◽  
Vol 62 (3) ◽  
pp. 580-585 ◽  
Author(s):  
Parminder S. Surdhar ◽  
Rizwan Ahmad ◽  
David A. Armstrong

Spectral changes and rates of reaction of flavins and several radical species have been investigated at pH 7, 9, and 10 in the presence and absence of β-mercaptoethanol. The radicals •CO2−, eaq−, and [Formula: see text] reacted with FAD at pH 10 to give a spectrum of FAD •Fl− with rate constants of 7 ± 1 × 108 and 4 ± 1 × 108 M−1 s−1 for •CO2− and [Formula: see text] respectively. At pH 7 only •FlH was observed and at pH 9 a mixture of •FIH and •Fl−.Interactions between flavin radicals and sulphydryl at 10−4 M concentration did not cause perturbations in the uv–visible spectra until either the radical and/or the sulphydryl were ionized. With FAD at pH 9 or 10 and LFl at pH 10 the 370 nm peak of •Fl− was enhanced by about 15% and a second larger growth occurred near 450 nm in the presence of 10−4 to 10−2 M sulphydryl. We attribute this to the formation of labile intermediate RSHFl•−, which must also be involved in the reduction of Fl by [Formula: see text] at pH 9 or 10.The second order rate constant k13 for reaction of [Formula: see text] with FAD at pH 9 and 10 was found to be 4.2 ± 0.5 × 108 M−1 s−1 and 2.0 ± 0.4 × 108 M−1 s−1 respectively. The rate constant for the reaction between [Formula: see text] and LFl at pH 10 was slightly faster, 7 ± 1 × 108 M−1 s−1, probably reflecting the fact that LFl lacks the bulky negatively charged adenine dinucleotide group of FAD.


2020 ◽  
Author(s):  
Chunhong Long ◽  
Liqiang Dai ◽  
Chao E ◽  
Lin-Tai Da ◽  
Jin Yu

ABSTRACTCas1 and Cas2 are highly conserved proteins across CRISPR-Cas systems and play a significant role in protospacer acquisition. Here we study the protospacer (or ps) DNA binding, recognition, and response to cleavage on the protospacer-adjacent-motif complementary sequence or PAMc by Cas1-Cas2, implementing all-atom molecular dynamics simulations. First, we noticed that two active sites of Cas1&1’ bind asymmetrically to two identical PAMc in the simulation. For psDNA containing only one PAMc to be recognized, it is then found that the non-PAMc association site remains destabilized until after the bound PAMc being cleaved. Thus, correlation appears to exist between the two active sites, which can be allosterically mediated by psDNA and Cas2&2’ in bridging. To substantiate such findings, we further simulated Cas1-Cas2 in complex with synthesized psDNA sequences psL and psH, which have been measured with low and high efficiency in acquisition, respectively. Notably, such inter-site correlation becomes largely enhanced for Cas1-Cas2 in complex with psH, and remains low with psL. Hence, our studies demonstrate that PAMc recognition and cleavage in one active site of Cas1-Cas2 allosterically regulates non-PAMc association/reaction in the other site, and such allosteric regulation is mediated by non-catalytic Cas 2 and DNA protospacer in acquisition.


1985 ◽  
Vol 63 (10) ◽  
pp. 2673-2678 ◽  
Author(s):  
Robert A. McClelland ◽  
Claude Moreau

Hydrolysis kinetics are reported for four spiro ortho esters: 3,4-dihydro-6-methoxy-1H-2-benzopyran-1-spiro-2′-1′,3′-dioxolane (13), its 1′,3′-dioxane analog (14), and the 6-unsubstituted versions of each (11 and 12). For comparison, also included are the diethoxy analogs: 1,1-diethoxy-3,4-dihydro-6-methoxy-1H-2-benzopyran (10) and the 6-unsubstituted compound (9). Product analysis implicates an initial opening of the dioxolane or dioxane ring in the spiro ortho esters, as expected on the basis of stereoelectronic considerations. The intermediate dialkoxycarbocations can be observed in HCl solutions. A detailed analysis has been carried out for the 6-methoxy systems to provide the rate constants k1, the second-order rate constant for H+-catalyzed formation of the cation from the ortho ester, k2, the first-order rate constant for water addition to the cation, and k−1, the first-order rate constant for ring closing of the cation to reform the ortho ester. The two spiro ortho esters are shown in this analysis to undergo reversible ring opening in their hydrolysis, in that values of k−1, are greater than k2. The differences, however, are not large, k−1/k2 being 1.2 (dioxolane, 13) and 3.8 (dioxane, 14). Comparison with the diethoxy ortho ester also reveals that the ring opening process (k1, rate constants) is inherently more difficult with the dioxolane, although not with the dioxane. An argument involving lone pair orientation is advanced to explain this.


1987 ◽  
Vol 241 (3) ◽  
pp. 893-897 ◽  
Author(s):  
M Leyh-Bouille ◽  
M Nguyen-Distèche ◽  
C Bellefroid-Bourguignon ◽  
J M Ghuysen

The 26,000-Mr DD-peptidase of Streptomyces K15 binds one equivalent of thiol reagents as 5,5′-dithiobis-(2-nitrobenzoate) or p-chloromercuribenzoate (pCMB). Derivatization of the DD-peptidase by pCMB decreases the efficacy of the initial binding of the ester carbonyl donor Ac2-L-Lys-D-Ala-D-lactate to the enzyme (K), the rate of enzyme acylation by the donor (K+2) and the rate of enzyme deacylation (k+3). However, the value of the k+2/k+3 ratio, and therefore the percentage of total enzyme which, at saturating concentrations of the donor, is present as acyl-enzyme at the steady state of the reaction, are not modified. The enzyme's binding sites for pCMB and benzylpenicillin are not mutually exclusive. But, when compared with the native enzyme, the pCMB-derivatized enzyme undergoes acylation by benzylpenicillin with a decreased second-order-rate constant (k+2/K) value and gives rise to a penicilloyl adduct of increased stability. Since the acyl-enzyme mechanism is not annihilated by pCMB derivatization, it is proposed that basically, and like all the other DD-peptidases/penicillin-binding proteins so far characterized, the Streptomyces K15 DD-peptidase is an active-site-serine enzyme.


Sign in / Sign up

Export Citation Format

Share Document