首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Resonance Raman spectroscopy has been used to investigate the structure and environment of the heme group in bovine liver catalase compound II. Both Soret- and Q-band excitation have been employed to observe and assign the skeletal stretching frequencies of the porphyrin ring. The oxidation state marker band v4 increases in frequency from 1373 cm-1 in ferricatalase to 1375 cm-1 in compound II, consistent with oxidation of the iron atom to the Fe(IV) state. Oxidation of five-coordinate, high-spin ferricatalase to compound II is accompanied by a marked increase of the porphyrin core marker frequencies that is consistent with a six-coordinate low-spin state with a contracted core. An Fe(IV) = O stretching band is observed at 775 cm-1 for compound II at neutral pH, indicating that there is an oxo ligand at the sixth site. At alkaline pH, the Fe(IV) = O stretching band shifts to 786 cm-1 in response to a heme-linked ionization that is attributed to the distal His-74 residue. Experiments carried out in H218O show that the oxo ligand of compound II exchanges with bulk water at neutral pH, but not at alkaline pH. This is essentially the same behavior exhibited by horseradish peroxidase compound II and the exchange reaction at neutral pH for both enzymes is attributed to acid/base catalysis by a distal His residue that is believed to be hydrogen-bonded to the oxo ligand. Thus, the structure and environment of the heme group of the compound II species of catalase and horseradish peroxidase are very similar. This indicates that the marked differences in their reactivities as oxidants are probably due to the manner in which the protein controls access of substrates to the heme group.  相似文献   

2.
Fe(IV)=O resonance Raman stretching vibrations were recently identified by this laboratory for horseradish peroxidase compound II and ferryl myoglobin. In the present report it is shown that Fe(IV)=O stretching frequency for horseradish peroxidase compound II will switch between two values depending on pH, with pKa values corresponding to the previously reported compound II heme-linked ionizations of pKa = 6.9 for isoenzyme A-2 and pKa = 8.5 for isoenzyme C. Similar pH-dependent shifts of the Fe(IV)=O frequency of ferryl myoglobin were not detected above pH 6. The Fe(IV)=O stretching frequencies of compound II of the horseradish peroxidase isoenzymes at pH values above the transition points were at a high value approaching the Fe(IV)=O stretching frequency of ferryl myoglobin. Below the transition points the horseradish peroxidase frequencies were found to be 10 cm-1 lower. Frequencies of the Fe(IV)=O stretching vibrations of horseradish peroxidase compound II for one set of isoenzymes were found to be sensitive to deuterium exchange below the transition point but not above. These results were interpreted to be indicative of an alkaline deprotonation of a distal amino acid group, probably histidine, which is hydrogen bonded to the oxyferryl group below the transition point. Deprotonation of this group at pH values above the pKa disrupts hydrogen bonding, raising the Fe(IV)=O stretching frequency, and is proposed to account for the lowering of compound II reactivity at alkaline pH. The high value of the Fe(IV)=O vibration of compound II above the transition point appears to be identical in frequency to what is believed to be the Fe(IV)=O vibration of compound X.  相似文献   

3.
We report geometry optimization results on heme compound I (ferryl-oxo + porphyrin cation radical), compound II (ferryl-oxo) and ferric-hydroxo species with thiolate or imidazole axial ligands. We also examine protonated forms of compound I and compound II species, prompted by recent reports that, in at least two different hemoproteins, compound II may in fact contain a hydroxo rather than an oxo ligand. We propose that the stable compound I and compound II species of hemoproteins (e.g., peroxidases and myoglobin) most likely contain a hydroxo rather than the oxo ligand traditionally assumed, whereas the extremely transient compound I species of monooxygenase hemoproteins (P450) would contain an oxo atom. We show evidence impacting the previously accepted notion in hemoprotein computational chemistry that non-covalent interactions and medium polarization effects are essential in properly describing the electronic structure of heme-thiolate high-valent complexes. On a different note, we find that the charge density on the iron remains essentially the same throughout the catalytic cycles of heme-containing oxygenases and peroxidases, despite clear changes in bond lengths and spin densities suggestive of various iron oxidation states. The iron thus appears to simply relay the electron flux between the porphyrin and the axial dioxygen/superoxo/peroxo/oxo/hydroxo ligands.  相似文献   

4.
The reaction of metmyoglobin with H2O2 was investigated in a pH range between 8.5 and 6.0 with the aid of stopped flow-rapid scan and rapid freezing-EPR techniques. Singular value decomposition analyses of the stopped flow data at pH 8.5 revealed that a spectral species previously unknown accumulated during the reaction and exhibited a Soret absorption maximum at >/=423 nm. In the EPR experiments, the new species exhibited a set of g values at 2.32, 2.19, and 1.94, indicating that the species was assignable to a ferric hydroperoxy (Fe(III)[O-O-H]-) compound. In contrast, the hydroperoxy compound scarcely accumulated in the reaction at pH 6.0, and the dominant intermediate species accumulated was compound I, which was derived from the oxygen-oxygen bond cleavage of the hydroperoxy compound. The accumulated amount of the hydroperoxy compound relative to compound I showed a pH dependence with an apparent pKa (pKaapp) from 6.95 to 7.27 depending on the metmyoglobins examined. This variation in pKaapp paralleled that in pKa of the acid-alkaline transition (pKaAB) of metmyoglobins, suggesting that the accumulation of hydroperoxy compound is controlled by the distal histidine. We propose that the H2O2 activation by metmyoglobin is promoted at the acidic condition due to the imidazolium form of the distal histidine, and we further propose that the controlled protonation state of the distal histidine is important for the facile O-O bond cleavage in heme peroxidases.  相似文献   

5.
Studies that elucidate the behavior of the hemoglobins (Hbs) and myoglobins upon reaction with hydrogen peroxide are essential to the development of oxygen carrier substitutes. Stopped-flow kinetics and resonance Raman data show that the reaction between hydrogen peroxide and oxygenated and deoxygenated ferric Hb I (oxy- and deoxy-HbI) from Lucina pectinata produce compound I and compound II ferryl species. The rate constants ratio (k23/k41) between the formation of compound II from compound I (k23) and the oxidation of the ferrous HbI (k41, i.e., 25 M(-1) s(-1)) of 12 x 10(-4) M suggests that HbI has a peroxidative capacity for removing H2O2 from solution. Resonance Raman presents the formation of both, met-aquo-HbI and compound II ferryl species in the cyclic reaction of HbI with H2O2. The ferric HbI species is maintained by the presence of H2O2; it can produce HbI compound I, or it can be reduced to a deoxy-HbI derivative to form HbI compound II upon reaction with H2O2. The compound II ferryl vibration frequency appears at 805 and 769 cm(-1) for HbIFe(IV)=(16)O and HbIFe(IV)=(18)O species, respectively. This ferryl mode indicates the absence of hydrogen bonding between the carbonyl group of the distal Q64 and the HbIFe(IV)=O ferryl moiety. The observation suggests that both the trans-ligand effect and the polarizabilty of the HbI heme pocket are responsible for the observed ferryl oxo vibrational energy. The vibrational mode also suggests that the carbonyl group of the distal Q64 is oriented toward the iron of the heme group, increasing the distal pocket electron density.  相似文献   

6.
Bromoperoxidase Compound I has been formed in reactions between bromoperoxidase and organic peroxide substrates. The absorbance spectrum of bromoperoxidase Compound I closely resembles the Compound I spectra of other peroxidases. The pH dependence of the second order rate constant for the formation of Compound I with hydrogen peroxide demonstrates the presence of an ionizable group at the enzyme active site having a pKa of 5.3. Protonation of this acidic group inhibits the rate of Compound I formation. This pKa value is higher than that determined for other peroxidases but the overall pH rate profiles for Compound I formation are similar. The one-electron reduction of bromoperoxidase Compound I yields Compound II and a second reduction yields native enzyme. Bromoperoxidase Compound II readily forms Compound III in the presence of an excess of hydrogen peroxide. Compound III passes through an as yet uncharacterized intermediate (III) in its decay to native enzyme. Compound III is produced and accumulates in enzymatic bromination reactions to become the predominate steady state form of the enzyme. Since Compound III is inactive as catalyst for enzymatic bromination, its accumulation leads to an idling reaction pathway which displays an unusual kinetic pattern for the bromination of monochlorodimedone.  相似文献   

7.
In all mammalian peroxidases, the heme is covalently attached to the protein via two ester linkages between conserved aspartate (Asp94) and glutamate residues (Glu242) and modified methyl groups on pyrrole rings A and C. Only myeloperoxidase has an additional sulfonium ion linkage between the sulfur atom of the conserved methionine 243 and the beta-carbon of the vinyl group on pyrrole ring A. Upon reduction from Fe(III) to Fe(II), lactoperoxidase (LPO) but not myeloperoxidase (MPO) is shown to adopt three distinct active site conformations which depend on pH and time. Comparative spectroscopic analysis (UV-Vis absorption and resonance Raman) of the ferrous forms of LPO, wild-type MPO and the variants Asp94Val, Glu242Gln, Met243Thr and Met243Val clearly demonstrate that a single, stable ferrous form of MPO is present only in those proteins which retain an intact sulfonium linkage. By contrast, both ferrous Met243Thr and Met243Val can assume two conformations. They resemble ferrous LPO, being five-coordinated high-spin species that are distinguished by the strength of the proximal Fe-histidine bond. This bond weakens with time or decreasing pH, as indicated by the Fe-histidine stretching bands.  相似文献   

8.
New class of 19F pH indicators: fluoroanilines   总被引:1,自引:0,他引:1       下载免费PDF全文
The pH dependence of the 19F chemical shift has been characterized for a number of fluorine-substituted aniline derivatives. These compounds constitute a new class of 19F nuclear magnetic resonance (NMR) pH indicators, characterized by single 19F resonance lines with sensitivities ranging from 2 to 7 ppm/pH unit near the aniline pKa; total shifts between conjugate acid and base of 5-15 ppm; and pKas ranging from 1 to 7. One compound, N,N-(methyl-2-carboxyisopropyl)-4-fluoroaniline, has a pKa of 6.8 and a sensitivity of 5 ppm/pH unit. This compound displays significant broadening of its 19F resonance near the aniline pKa (6.8), due to a decreased rate of exchange between conjugate acid and base species. Our results are consistent with slow dissociation of an intramolecular hydrogen bond in the zwitterionic species that limits the exchange rate between protonated and unprotonated forms for N,N-(methyl-2-carboxyisopropyl)-4-fluoroaniline.  相似文献   

9.
Salicylhydroxamic acid inhibits myeloperoxidase activity.   总被引:3,自引:0,他引:3  
Salicylhydroxamic and benzohydroxamic acids were found to bind to the resting state of myeloperoxidase and inhibit ligand binding to the heme iron. An ionizable group on the enzyme with pKa = 4 affects salicylhydroxamic acid binding; binding occurs when this group is not protonated. The binding of the heme iron ligands (e.g. cyanide, nitrite, and chloride) is probably controlled by the same ionizable group. The equilibrium dissociation constant of the salicylhydroxamic acid-myeloperoxidase complex is about 2 x 10(-6) M, and the association rate constant is 7.4 x 10(6) M-1.s-1. Salicylhydroxamic acid serves as a donor to the higher oxidation state of myeloperoxidase and thereby inhibits guaiacol oxidation. Salicylhydroxamic acid was also found to bind to intestinal peroxidase and lactoperoxidase. Salicylhydroxamic acid binding to all three mammalian peroxidases was about 3 orders of magnitude stronger than benzohydroxamic acid binding. We conclude that the salicylhydroxamic and benzohydroxamic acids bind in the distal heme cavity of these peroxidases and interact with the heme ligand binding site.  相似文献   

10.
The active site amino acid residues of lignin peroxidase are homologous to those of other peroxidases; however, in contrast to other peroxidases, no pH dependence is observed for the reaction of ferric lignin peroxidase with H2O2 to form compound I (Andrawis, A., Johnson, K.A., and Tien, M. (1988) J. Biol. Chem. 263, 1195-1198). Chloride binding is used in the present study to investigate this reaction further. Chloride binds to lignin peroxidase at the same site as cyanide and hydrogen peroxide. This is indicated by the following. 1) Chloride competes with cyanide in binding to lignin peroxidase. 2) Chloride is a competitive inhibitor of lignin peroxidase with respect to H2O2. The inhibition constant (Ki) is equal to the dissociation constant (Kd) of chloride at all pH values studied. Chloride binding is pH dependent: chloride binds only to the protonated form of lignin peroxidase. Transient-state kinetic studies demonstrate that chloride inhibits lignin peroxidase compound I formation in a pH-dependent manner with maximum inhibition at low pH. An apparent pKa was calculated at each chloride concentration; the pKa increased as the chloride concentration increased. Extrapolation to zero chloride concentration allowed us to estimate the intrinsic pKa for the ionization in the lignin peroxidase active site. The results reported here provide evidence that an acidic ionizable group (pKa approximately 1) at the active site controls both lignin peroxidase compound I formation and chloride binding. We propose that the mechanism for lignin peroxidase compound I formation is similar to that of other peroxidases in that it requires the deprotonated form of an ionizable group near the active site.  相似文献   

11.
The catalytic cycle intermediates of heme peroxidases, known as compounds I and II, have been of long standing interest as models for intermediates of heme proteins, such as the terminal oxidases and cytochrome P450 enzymes, and for non-heme iron enzymes as well. Reports of resonance Raman signals for compound I intermediates of the oxo-iron(IV) porphyrin pi-cation radical type have been sometimes contradictory due to complications arising from photolability, causing compound I signals to appear similar to those of compound II or other forms. However, studies of synthetic systems indicated that protein based compound I intermediates of the oxoiron(IV) porphyrin pi-cation radical type should exhibit vibrational signatures that are different from the non-radical forms. The compound I intermediates of horseradish peroxidase (HRP), and chloroperoxidase (CPO) from Caldariomyces fumago do in fact exhibit unique and characteristic vibrational spectra. The nature of the putative oxoiron(IV) bond in peroxidase intermediates has been under discussion in the recent literature, with suggestions that the Fe(IV)O unit might be better described as Fe(IV)-OH. The generally low Fe(IV)O stretching frequencies observed for proteins have been difficult to mimic in synthetic ferryl porphyrins via electron donation from trans axial ligands alone. Resonance Raman studies of iron-oxygen vibrations within protein species that are sensitive to pH, deuteration, and solvent oxygen exchange, indicate that hydrogen bonding to the oxoiron(IV) group within the protein environment contributes to substantial lowering of Fe(IV)O frequencies relative to those of synthetic model compounds.  相似文献   

12.
The reaction between native myeloperoxidase and hydrogen peroxide, yielding Compound II, was investigated using the stopped-flow technique. The pH dependence of the apparent second-order rate constant showed the existence of a protonatable group on the enzyme with a pKa of 4.9. This group is ascribed to the distal histidine imidazole, which must be deprotonated to enable the reaction of Compound I with hydrogen peroxidase to take place. The rate constant for the formation of Compound II by hydrogen peroxide was 3.5.10(4) M-1.s-1. During the reaction of myeloperoxidase with H2O2, rapid reduction of added cytochrome c was observed. This reduction was inhibitable by superoxide dismutase, and this demonstrates that superoxide anion radicals are generated. When potassium ferrocyanide was used as an electron donor to generate Compound II from Compound I, the pH dependence of the apparent second-order rate constant indicated involvement of a group with a pKa of 4.5. However, with ferrocyanide as an electron donor, protonation of the group was necessary to enable the reaction to take place. The rate constant for the generation of Compound II by ferrocyanide was 1.6.10(7) M-1.s-1. We also investigated the reaction of Compound II with hydrogen peroxide, yielding Compound III. Formation of Compound III (k = 50 M-1.s-1) proceeded via two different pathways, one of which was inhibitable by tetranitromethane. We further investigated the stability of Compound II and Compound III as a function of pH, ionic strength and enzyme concentration. The half-life values of both Compound II and Compound III were independent of the enzyme concentration and ionic strength. The half-life value of Compound III was pH-dependent, showing a decreasing stability with increasing pH, whereas the stability of Compound II was independent of pH over the range 3-11.  相似文献   

13.
The standard reduction potential of the redox couple compound I/native enzyme has been determined for human myeloperoxidase (MPO) and eosinophil peroxidase (EPO) at pH 7.0 and 25 degrees C. This was achieved by rapid mixing of peroxidases with either hydrogen peroxide or hypochlorous acid and measuring spectrophotometrically concentrations of the reacting species and products at equilibrium. By using hydrogen peroxide, the standard reduction potential at pH 7.0 and 25 degrees C was 1.16 +/- 0.01 V for MPO and 1.10 +/- 0.01 V for EPO, independently of the concentration of hydrogen peroxide and peroxidases. In the case of hypochlorous acid, standard reduction potentials were dependent on the hypochlorous acid concentration used. They ranged from 1.16 V at low hypochlorous acid to 1.09 V at higher hypochlorous acid for MPO and from 1.10 V to 1.03 V for EPO. Thus, consistent results for the standard reduction potentials of redox couple compound I/native enzyme of both peroxidases were obtained with all hydrogen peroxide and at low hypochlorous acid concentrations: possible reasons for the deviation at higher concentrations of hypochlorous acid are discussed. They include instability of hypochlorous acid, reactions of hypochlorous acid with different amino-acid side chains in peroxidases as well as the appearance of a compound I-chloride complex.  相似文献   

14.
We examine the issue of ferryl protonation in heme proteins. An analysis of the results obtained from X-ray crystallography, resonance Raman spectroscopy, and extended X-ray absorption spectroscopy (EXAFS) is presented. Fe-O bond distances obtained from all three techniques are compared using Badger's rule. The long Fe-O bond lengths found in the ferryl crystal structures of myoglobin, cytochrome c peroxidase, horseradish peroxidase, and catalase deviate substantially from the values predict by Badger's rule, while the oxo-like distances obtained from EXAFS measurements are in good agreement with the empirical formula. Density functional calculations, which suggest that M?ssbauer spectroscopy can be used to determine ferryl protonation states, are presented. Our calculations indicate that the quadrupole splitting (DeltaE(Q)) changes significantly upon ferryl protonation. New resonance Raman data for horse-heart myoglobin compound II (Mb-II, pH 4.5) are also presented. An Fe-O stretching frequency of 790cm(-1) (shifting to 754cm(-1) with (18)O substitution) was obtained. This frequency provides a Badger distance of r(Fe-O)=1.66A. This distance is in agreement with the 1.69A Fe-O bond distance obtained from EXAFS measurements but is significantly shorter than the 1.93A bond found in the crystal structure of Mb-II (pH 5.2). In light of the available evidence, we conclude that the ferryl forms of myoglobin (pKa4), horseradish peroxidase (pKa4), cytochrome c peroxidase (pKa4), and catalase (pKa7) are not basic. They are authentic Fe(IV)oxos with Fe-O bonds on the order of 1.65A.  相似文献   

15.
We report the optical and resonance Raman spectral characterization of ferrous recombinant Chlamydomonas LI637 hemoglobin. We show that it is present in three pH-dependent equilibrium forms including a 4-coordinate species at acid pH, a 5-coordinate high spin species at neutral pH, and a 6-coordinate low spin species at alkaline pH. The proximal ligand to the heme is the imidazole group of a histidine. Kinetics of the reactions with ligands were determined by stopped-flow spectroscopy. At alkaline pH, combination with oxygen, nitric oxide, and carbon monoxide displays a kinetic behavior that is interpreted as being rate-limited by conversion of the 6-coordinate form to a reactive 5-coordinate form. At neutral pH, combination rates of the 5-coordinate form with oxygen and carbon monoxide were much faster (>10(7) microM-1 s-1). The dissociation rate constant measured for oxygen is among the slowest known, 0.014 s-1, and is independent of pH. Replacement of the tyrosine 63 (B10) by leucine or of the putative distal glutamine by glycine increases the dissociation rate constant 70- and 30-fold and increases the rate of autoxidation 20- and 90-fold, respectively. These results are consistent with at least two hydrogen bonds stabilizing the bound oxygen molecule, one from tyrosine B10 and the other from the distal glutamine. In addition, the high frequency (232 cm-1) of the iron-histidine bond suggests a structure that lacks any proximal strain thus contributing to high ligand affinity.  相似文献   

16.
Redox thermodynamic data provide a detailed insight into control of the reduction potential E degrees' of the [Fe(S-Cys)4] site in rubredoxin. Mutant forms were studied in which specific structural changes were made in both the primary and secondary coordination spheres. Those changes have been probed by resonance Raman spectroscopy. The decrease of approximately 200 mV in E degrees' observed for the [Fe(S-Cys)3(O-Ser)]-/2- couples in the surface ligand mutants C9S and C42S is essentially enthalpic in origin and associated with the substitution of ligand thiolate by ligand olate. However, the pH dependence of the potentials below characteristic pKa(red) approximately equals 7 is an entropic contribution, plausibly associated with increased conformational flexibility induced by a longer Fe(II)-O(H)-Ser bond in the reduced form. The presence of a second surface Ser ligand in the new double mutant protein C9S/C42S affects the enthalpic term primarily for pH>pKa(red) > or = 9.3, but for pHpKa approximately 9: [Fe(III)(S-Cys)3(OH)]- + e- --> [Fe(II)(S-Cys)3(OH)]2-. pH [Fe(II)(S-Cys)3(OH2)]-.  相似文献   

17.
The pH- and temperature-dependent changes in the coordination sphere of the heme c group of N alpha-acetyl microperoxidase-8 (Ac-MP-8) have been studied by examining its optical, resonance Raman, electron paramagnetic resonance, and magnetic circular dichroism spectra. An optical titration indicates that Ac-MP-8 exists in three major ionization forms over the pH 1-12 range that are linked by pK alpha values of approximately 3 and 9. The acid form that is present at pH 1.5 exists as a mixture of five- and six-coordinate high-spin species and most likely has water or buffer ions as axial ligand(s). On titration to pH 7, the His18 residue is deprotonated and becomes the proximal ligand to the iron to give a six-coordinate neutral form that has water as the sixth ligand. This form exists in a thermal high-spin intermediate-spin state equilibrium. On raising the pH to 10, an alkaline form is generated which is predominantly a five-coordinate high-spin species. It is formed by ionization of the proximal His18 residue to its imidazolate form with concomitant dissociation of the water ligand at the sixth site. At concentrations of Ac-MP-8 greater than 10 microM, some six-coordinate low-spin species are formed that are attributed to a dimer in which a His18 residue from a second molecule of Ac-MP-8 coordinates to the sixth site of another to give a bis-His complex. Raising the pH to 11.5 does not produce an appreciable amount of the six-coordinate complex with hydroxide as the sixth ligand. These studies show that Ac-MP-8 is a good water-soluble model for the peroxidases that exhibits minimal aggregation at concentrations below 10 microM in the neutral and alkaline pH regions.  相似文献   

18.
Catalase-peroxidases (KatGs) are unique peroxidases exhibiting a high catalase activity and a peroxidase activity with a wide range of artificial electron donors. Exchange of tyrosine 249 in Synechocystis KatG, a distal side residue found in all as yet sequenced KatGs, had dramatic consequences on the bifunctional activity and the spectral features of the redox intermediate compound II. The Y249F variant lost catalase activity but retained a peroxidase activity (substrates o-dianisidine, pyrogallol, guaiacol, tyrosine, and ascorbate) similar to the wild-type protein. In contrast to wild-type KatG and similar to monofunctional peroxidases, the formation of the redox intermediate compound I could be followed spectroscopically even by addition of equimolar hydrogen peroxide to ferric Y249F. The corresponding bimolecular rate constant was determined to be (1.1 +/- 0.1) x 107 m-1 s-1 (pH 7 and 15 degrees C), which is typical for most peroxidases. Additionally, for the first time a clear transition of compound I to an oxoferryl-like compound II with peaks at 418, 530, and 558 nm was monitored when one-electron donors were added to compound I. Rate constants of reaction of compound I and compound II with tyrosine ((5.0 +/- 0.3) x 104 m-1 s-1 and (1.7 +/- 0.4) x 102 m-1 s-1) and ascorbate ((1.3 +/- 0.2) x 104 m-1 s-1 and (8.8 +/- 0.1) x 101 m-1 s-1 at pH 7 and 15 degrees C) were determined by using the sequential stopped-flow technique. The relevance of these findings is discussed with respect to the bifunctional activity of KatGs and the recently published first crystal structure.  相似文献   

19.
Two crystals forms of bovine pancreatic trypsin inhibitor are produced between pH 8.39 and 10.13 when crystals are grown at room temperature from solutions of 1.5 M potassium phosphate. Lower pH values favor the form II crystals, whereas higher pH values favor the form III. The transition from one crystal form to the other occurs at pH 9.35. We examined the crystal lattice contacts in both crystal forms and identified an unusual interaction we believe explains these observations. Spanning the crystallographic 2-fold axis in form III crystals, the Lys 41 side-chain amino nitrogens from 2 symmetry-related molecules are only 2.72 A apart, implying they are hydrogen bonded to one another. In form II crystals, the Lys 41 side-chain amino group is protonated and forms a salt bridge with a solvent-derived phosphate group. For the Lys 41 side-chain amino groups to hydrogen bond in form III crystals, at least 1 member of the pair must be deprotonated. The transition that occurs at pH 9.35 marks the pKa for deprotonation. In solution, the pKa for the Lys 41 side chain is around 10.8. The pKa for one of the interacting Lys 41 side chains in form III crystals is therefore shifted downward by about 1.5 pH units. The energy for lowering the pKa value comes from the many additional intermolecular hydrogen bonds that are present in form III crystals: 19 compared to only 8 in form II crystals.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号