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 15 毫秒
1.
The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5'-mono- and -triphosphates in D2O solution at pH' = 3 were measured. The paramagnetic probes were [Cr(III)(H2O)6]3+, [Cr(III)(H2O)3(HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III)(H2O)3(UTP)-, while the matrix nucleotides (0.1 M) were H2AMP, HIMP-, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]-/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(alpha,beta)], [Cr(III)(NH3)4(HPP)(alpha,beta)], [Co(III)(NH3)3(H2PPP)(alpha,beta,gamma)] and [Co(III)(NH3)4(HPP)(alpha,beta)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complexes as well as the structures of the outer sphere [Cr(III)(H2O)6]3(+)-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)- species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7)(O...N = 2.82 A). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2')H group and the adenine N(3) (O...N = 2.80 A) as well as phosphate oxygen atoms and a water molecule (O...O = 2.47 A). The metal center has an almost regular octahedral coordination geometry. The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar "anti" conformation around the N(9)-C(1') glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)- complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O...O = 2.63 A; O...N = 2.72, 2.70 A). For the H2AMP complex, the [Cr(III)(H2O)6]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s-1 (H(2)) for HIMP-.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Karle IL  Urry DW 《Biopolymers》2005,77(4):198-204
Tropoelastin is a complex polymeric protein composed primarily of repeating segments of Val-Pro-Gly-Gly, Val-Pro-Gly-Val-Gly, and Ala-Pro-Gly-Val-Gly-Val that occurs in connective tissue and arteries. It has rubber-like extensible properties. A synthetic cyclic dodecapeptide, with a double repeat of the hexapeptide sequence, has been shown to undergo a reversible inverse temperature transition; that is, crystals grow at 60 degrees C and dissolve in the mother liquor upon cooling. An x-ray crystal structure analysis established that the cyclic backbone formed an elongated loop with a Pro-Gly, type II beta turn at both ends. Six internal cross strand NH...OC hydrogen bonds form between six NH donors and four O=C acceptors where two of the carbonyl O atoms are bifurcated acceptors. As a result, the molecule is pulled up into a corrugated profile. The corrugated loops form extended beta-sheets by additional intermolecular hydrogen bonds. An analysis of the dome region in a corrugated sheet suggests a reversible mechanism for extending and contracting the length of the whole molecule, akin to the motion of opening and closing an umbrella, caused by the motion of a water molecule with its associated hydrogen bonds acting as spokes. Crystal parameters: C44H72N12O12.3H2O, sp. gr. P2(1)2(1)2(1), a = 9.212 angstroms, b = 19.055 angstroms, c = 32.247 angstroms, d = 1.157 g/cm3.  相似文献   

3.
The octahedral complex tetraammine(chloroaquo)cobalt(III) dichloride is shown to be the HCl hydrolysis product of both P1,2-bidentate tetraammine(pyrophosphato)cobalt(III) [Co(NH3)4HP2O7 or CoPP] and bidentate tetraammine(phosphato)cobalt(III) [Co(NH3)4PO4 or CoP]. The complex crystallizes in the orthorhombic space group Pna21 with cell dimensions a = 13.033(2)A, b = 6.710(1)A, and c = 10.318(2)A; the crystal structure was refined to a final disagreement index of 0.033. The average of the four Co-N distances is 1.944 +/- 6A. The Co-Cl distance is 2.257(2)A and the Co-O(W) distance is 1.971(4)A. Both protons of the coordinated water molecule are engaged in strong hydrogen bonds to the two nonbonded chloride counterions with O(W)-Cl distances of 3.087(6)A and 3.123(6)A. Each nonbonded chloride is engaged in seven hydrogen bonding interactions resulting from the high ratio of hydrogen bond donors to acceptors in the CoP structure. Cobalt bisphosphate (CoP2) is the final enzyme hydrolysis product when CoPP is used as substrate in the yeast inorganic pyrophosphatase reaction. The bridge oxygen atom is the site of initial CoPP cleavage both for HCl catalyzed hydrolysis as well as for enzyme catalyzed hydrolysis.  相似文献   

4.
The three dimensional molecular structure of the adduct formed between the anticancer drug cisplatin and a DNA dinucleotide d(pGpG) has been determined by x-ray diffraction analysis at 1.37 A resolution and refined to a final R-factor of 0.11. This structure, solved by using data from a previously reported crystal form in the space group C222(1), resembles that found in the space group P2(1)2(1)2 (Sherman, et al., Science, 230, 412-417, 1985; ibid, J. Amer. Chem. Soc. 110, 7368-7381, 1988). In both structures, four crystallographically independent cis-[Pt(NH3)2(d(pGpG]] molecules aggregate into a tetrameric cluster that is stabilized by a large number of intermolecular hydrogen bonds and base-base stacking interactions. In each molecule, the platinum atom is coordinated to the N7 atoms of two guanine bases arranged in a head-to-head orientation, resulting in a large dihedral angle between the guanines. Intermolecular guanine-guanine base pairings between different intrastrand crosslinked molecules are used extensively in the crystal lattice.  相似文献   

5.
The hydrogensquarate [LeuNH(2)] (HSq) of L-leucinamide has been synthesized and its structure has been determined by single crystal X-ray diffraction. A three dimensional network is formed by hydrogen bonds with participation of the O=C-NH(2) function, the hydrogensquarate ion and the N(+)H(3) group [NH(2)...O=C((Sq)) (2.840 and 2.749 A), ((Sq))OH...O=C(NH(2)) (2.618 A), NH(3) (+)...O=C((Sq)) (3.246, 2.804 and 2.823 A)], respectively. A theoretical approximation of the electronic structure was carried out by means of ab initio UMP2 and MP2 level of theory at the 6-311++G** basis set. The IR-spectroscopic assignment in the solid-phase was obtained by linear-polarized IR-spectroscopy of oriented samples as colloid suspensions in a nematic host and application of the reducing-difference procedure for the interpretation of polarized IR-spectra.  相似文献   

6.
The crystal structure of the DNA minor-groove DNA-binding drug berenil has been determined. Molecular-modelling techniques have been used to establish plausible binding modes of the structure to A-T sequences. These have shown that specific hydrogen bonds are possible between the amidine groups of the drug molecule and 02 atoms of thymine, although global energy minimisations tended to emphasise electrostatic interactions with phosphate groups rather than these hydrogen bonds with bases.  相似文献   

7.
The structure of the complex formed between d(CGTACG)(2) and the antitumor agent 9-amino-[N-(2-dimethylamino)ethyl]acridine-4-carboxamide has been solved to a resolution of 1.6 A using X-ray crystallography. The complex crystallized in space group P6(4) with unit cell dimensions a = b = 30.2 A and c = 39.7 A, alpha = beta = 90 degrees, gamma = 120 degrees. The asymmetric unit contains a single strand of DNA, 1. 5 drug molecules, and 29 water molecules. The final structure has an overall R factor of 19.3%. A drug molecule intercalates between each of the CpG dinucleotide steps with its side chain lying in the major groove, and the protonated dimethylamino group partially occupies positions close to ( approximately 3.0 A) the N7 and O6 atoms of guanine G2. A water molecule forms bridging hydrogen bonds between the 4-carboxamide NH and the phosphate group of the same guanine. Sugar rings adopt the C2'-endo conformation except for cytosine C1 which moves to C3'-endo, thereby preventing steric collision between its C2' methylene group and the intercalated acridine ring. The intercalation cavity is opened by rotations of the main chain torsion angles alpha and gamma at guanines G2 and G6. Intercalation perturbs helix winding throughout the hexanucleotide compared to B-DNA, steps 1 and 2 being unwound by 8 degrees and 12 degrees, respectively, whereas the central TpA step is overwound by 17 degrees. An additional drug molecule, lying with the 2-fold axis in the plane of the acridine ring, is located at the end of each DNA helix, linking it to the next duplex to form a continuously stacked structure. The protonated N,N-dimethylamino group of this "end-stacked" drug hydrogen bonds to the N7 atom of guanine G6. In both drug molecules, the 4-carboxamide group is internally hydrogen bonded to the protonated N-10 atom of the acridine ring. The structure of the intercalated complex enables a rationalization of the known structure-activity relationships for inhibition of topoisomerase II activity, cytotoxicity, and DNA-binding kinetics for 9-aminoacridine-4-carboxamides.  相似文献   

8.
The structure of the complex between the minor groove binder netropsin and d(GGCCAATTGG) was determined via single-crystal X-ray techniques. The structure was refined to completion using refmac5.1.24, resulting in a residual R-factor of 20.0% (including 68 water molecules). Using crystal engineering and cryocooling techniques, the resolution could be enhanced to 1.75 A, resulting in an unambiguous determination of the drug conformation and orientation. As previously noticed, bifurcated hydrogen bonds are formed between the amide nitrogen atoms of the drug and the N3 and O2 atoms of A and T base pairs, respectively, clearly cataloging the structure to class I. As the bulky NH2 group on guanine was believed to prevent binding of the drug in the minor groove, the detailed nature of several of the amidinium and guanidinium end contacts were further investigated by ab initio quantum chemical methods.  相似文献   

9.
In the equilibrium between B-DNA and Z-DNA in poly(dC-dG), the [Co(NH3)6]3+ ion stabilizes the Z form 4 orders of magnitude more effectively than the Mg2+ ion. The structural basis of this difference is revealed in Z-DNA crystal structures of d(CpGpCpGpCpG) stabilized by either Na+/Mg2+ or Na+/Mg2+ plus [Co(NH3)6]3+. The crystals diffract X-rays to high resolution, and the structures were refined at 1.25 A. The [Co(NH3)6]3+ ion forms five hydrogen bonds onto the surface of Z-DNA, bonding to a guanine O6 and N7 as well as to a phosphate group in the ZII conformation. The Mg2+ ion binds through its hydration shell with up to three hydrogen bonds to guanine N7 and O6. Higher charge, specific fitting of more hydrogen bonds, and a more stable complex all contribute to the great effectiveness of [Co(NH3)6]3+ in stabilizing Z-DNA.  相似文献   

10.
Abstract

Sarma et al. (J. Biomol. Str. and Dynam. 2, 1085 (1985) have proposed, on the basis of nuclear magnetic resonance experiments on the complex of netropsin with poly(dA)·poly(dT), that the drug molecule lies asymmetrically along the dA side of the minor groove and makes hydrogen bonds only with the dA strand. If the crystal structure analyses of B-DNA (Fratini et al., J. Biol. Chem. 257, 14686 (1982)) and of its complex with netropsin (Kopka et al., J. Mol. Biol. 183, 553 (1985)) are any guide, this off-center, wide-groove model is stereochemically unlikely. More to the point, the off-center model is unnecessary to explain the observed nmr data. All of the nuclear Overhauser and other observations are fully explained by the structure seen in the x-ray crystal analysis, in which netropsin sits squarely centered within the minor groove, making bifurcated hydrogen bonds with both strands.  相似文献   

11.
Binding of an antitumor drug to DNA, Netropsin and C-G-C-G-A-A-T-T-BrC-G-C-G   总被引:27,自引:0,他引:27  
The antitumor antibiotic netropsin has been co-crystallized with a double-helical B-DNA dodecanucleotide of sequence: C-G-C-G-A-A-T-T-BrC-G-C-G, and the structure of the complex has been solved by X-ray diffraction at a resolution of 2.2 A. The structure has been refined independently by Jack-Levitt and Hendrickson-Konnert least-squares methods, leading to a final residual error of 0.257 by the Jack-Levitt approach (0.211 for two-sigma data) or 0.248 by the Hendrickson-Konnert approach, with no significant difference between refined structures. The netropsin molecule displaces the spine of hydration and fits snugly within the minor groove in the A-A-T-T center. It widens the groove slightly and bends the helix axis back by 8 degrees, but neither unwinds nor elongates the double helix. The drug molecule is held in place by amide NH hydrogen bonds that bridge adenine N-3 and thymine O-2 atoms, exactly as with the spine of hydration. The requirement of A X T base-pairs in the binding site arises because the N-2 amino group of guanine would demand impermissibly close contacts with netropsin. It is proposed that substitution of imidazole for pyrrole in netropsin should create a family of "lexitropsins" capable of reading G X C-containing base sequences.  相似文献   

12.
Several lines of functional evidence have shown that anion binding to a nonsynergistic site is a prerequisite for the anion-mediated iron release mechanism of transferrins. We report here structural evidence of the location of sulfate anion binding sites of the ovotransferrin N-lobe via the 1.90 A resolution apo crystal structure. The crystals were grown in an ammonium sulfate solution and belonged to space group P6(3)22 with the following unit cell dimensions: a = b = 125.17 A and c = 87.26 A. The structural determination was performed by isomorphous replacement, using Pt and Au derivatives. The structure refinement gave an R-factor of 0.187 in the resolution range of 7.0-1.90 A for the final model. From the electron density map, the existence of four bound SO(4)(2)(-) anions was detected. Three of them that exhibited reasonably low B-factors were all located in the opened interdomain cleft (sites 1-3). In site 1, the bound anion directly interacts with an Fe(3+)-coordinating ligand; SO(4) O1 and SO(4) O3 form hydrogen bonds with His250 NE2. Oxygen atom O3 of the same sulfate anion makes a hydrogen bond with Ser91 OG in a hinge strand. The sulfate anion in site 2 partially occupies the synergistic anion binding sites; SO(4) O2 and SO(4) O3 are hydrogen bonded to Arg121 NE and NH2, respectively, both of which are consensus anchor groups for CO(3)(2)(-) anion in holotransferrins. The former oxygen atom of SO(4)(2)(-) is also hydrogen bonded to Ser122 N, which forms a hydrogen bond with Fe(3+)-coordinating ligand Asp60 OD2 in holotransferrins. Some of the SO(4)(2)(-) oxygen atoms in sites 1 and 2 interact indirectly through H(2)O molecules with functionally important protein groups, such as the other Fe(3+)-coordinating ligands, Tyr92 OH and Tyr191 OH, and a dilysine trigger group, Lys209 NZ. In site 3, SO(4) O1 and SO(4) O4 form hydrogen bonds with Ser192 OG and Tyr191 N, respectively, and SO(4) O2 forms hydrogen bonds with Ser192 N and Ser192 OG. These structural data are consistent with the view that the anion bindings to the interdomain cleft, especially to sites 1 and 2, play crucial roles in the domain opening and synergistic carbonate anion release in the iron release mechanism of the ovotransferrin N-lobe.  相似文献   

13.
Poly(L-histidine)-phosphate (H2PO4-, HPO4(2-)) and poly(L-glutamate)-phosphate systems (residue/phosphate, 1:1) in the presence of Ca2+ are studied by infrared spectroscopy. In the poly(L-histidine)-phosphate systems N...HOP in equilibrium NH+...O-P hydrogen bonds are formed where most phosphate protons are found at the histidine ring. With an increase in the degree of hydration the proportion of the proton limiting structure NH+...O-P increases. In the poly(L-glutamate)-dihydrogen phosphate system most phosphate protons are found at the carboxylate groups. Different behavior is observed for poly(L-glutamate)-hydrogen phosphate mixtures, where the residence time of the phosphate proton at the hydrogen acceptor carboxylate group is very short. This residence time increases, however, with increasing humidity. All these results support the triphasic theory of biological calcification involving a tripartite protein-calcium-phosphate complex where these hydrogen bonds can be present. The behavior of these hydrogen bonds can also explain the formation of a nidus of calcium phosphate salts in calcium oxalate-containing urinary calculi.  相似文献   

14.
The crystal structures of the 2:1 complex of the self-complementary DNA octamer d(GAAGCTTC) with actinomycin D has been determined at 3.0 A resolution. This is the first example of a crystal structure of a DNA-drug complex in which the drug intercalates into the middle of a relatively long DNA segment. The results finally confirmed the DNA-actinomycin intercalation model proposed by Sobell & co-workers in 1971. The DNA molecule adopts a severely distorted and slightly kinked B-DNA-like structure with an actinomycin D molecule intercalated in the middle sequence, GC. The two cyclic depsipeptides, which differ from each other in overall conformation, lie in the minor groove. The complex is further stabilized by forming base-peptide and chromophore-backbone hydrogen bonds. The DNA helix appears to be unwound by rotating one of the base-pairs at the intercalation site. This single base-pair unwinding motion generates a unique asymmetrically wound helix at the binding site of the drug, i.e. the helix is loosened at one end of the intercalation site and tightened at the other end. The large unwinding of the DNA by the drug intercalation is absorbed mostly in a few residues adjacent to the intercalation site. The asymmetrical twist of the DNA helix, the overall conformation of the two cyclic depsipeptides and their interaction mode with DNA are correlated to each other and rationally explained.  相似文献   

15.
The interaction energy was calculated, by the ab initio FMO method, for complexes between LCK protein and four inhibitors (staurosporine, BMS compound 2, and our compounds 3 and 4). In every case a number of CH/pi hydrogen bonds have been disclosed in the so-called adenine pocket. In complexes of 2, 3, and 4, CH/pi and NH/pi hydrogen bonds have been observed in another pocket. In view of the above results, the aniline ring of 3 was replaced by 2,6-dimethyl aniline to increase the potency for LCK kinase. A 10-fold increase in the potency has been achieved for 4 over 3. We suggest that the concept of weak hydrogen bonds is useful in the rational design of drugs.  相似文献   

16.
A crystal of d(CGCGCG) in the Z-DNA lattice was soaked with ruthenium(III) hexaammine and its structure refined at 1.2 A resolution. Three unique metal complexes were found absorbed to each hexamer duplex. In addition, two symmetry-related binding sites were located, yielding a total of five ruthenium complexes bound to each d(CGCGCG) duplex. One unique site and its symmetry related site are nearly identical to the binding site of cobalt(III) hexaammine on Z-DNA. At that position, the metal complex bridges the convex surfaces of two adjacent Z-DNA strands by hydrogen bonds to the N7 and O6 functional groups of the guanine bases. The remaining three ruthenium three ruthenium(III) hexaammine binding sites are not present in the cobalt(III) hexaammine Z-DNA structure. Of these, two are related by symmetry and span the gap between the convex outer surface of one Z-DNA strand and the helical groove crevice of a neighboring strand. The third ruthenium site has no symmetry mate and involves interactions with only the deep groove. In this interaction, the metal complex hydrogen bonds to both the phosphate backbone and to a set of primary shell water molecules that extend the hydrogen bonding potential of the deep groove crevice out to the surface of the molecule. Solution studies comparing the circular dichroism spectra of low salt poly(dG-dC).poly(dG-dC) samples in the presence of ruthenium(III) and cobalt(III) hexammine show that the ruthenium complex does stabilize Z-DNA in solution, but not as effectively as the cobalt analogue. This suggests that some of the interactions available for the larger ruthenium complex may not be important for stabilization of the left-handed DNA conformation.  相似文献   

17.
In the course of our work relating to the design of a bihelical structure (I) from diphenic anhydride by tethering with cystine di-OMe, stable, hard, and rigid crystals, mp 215-218 degrees C were isolated in low yields ( approximately 2%). The crystal structure established that it was a bis amide (II) arising from diphenic acid and cystine di-OMe [(II), C(22)H(22)N(2)O(6)S(2) (a = 9.897 (1) A, b = 12.210 (1) A, c = 18.192 (1) A, sp. gr. P2(1)2(1)2(1))]. An authentic sample of (II) was subsequently prepared in 47% yields by condensation of diphenic acid dichloride with cystine di-OMe. A most surprising feature of II was, despite its high density, rigidity, and hardness, it did not exhibit any normal hydrogen bonds. The nearest approximation to a "usual" hydrogen bond was the single NH...OC linkage that occurred between molecules along a twofold screw axis. In this linkage, N...O = 3.265 A and H...O = 2.43 A, values that are at least 10% longer than those usually observed in peptides. The rigidity of the crystals appears to depend upon many weak hydrogen bonds of the type CH...O, CH...pi, CH...S, and NH...S working in concert. Even these attractions have separations that are at the high end of the range of previously observed values, although some of the weak hydrogen bonds have been rarely reported and have poorly defined ranges. The attractive effect of each of these weak bonds may be enhanced by the occurrence of a number of them in a parallel fashion like rungs in a ladder.  相似文献   

18.
Lucina pectinata ctenidia harbor three heme proteins: sulfide-reactive hemoglobin I (HbI(Lp)) and the oxygen transporting hemoglobins II and III (HbII(Lp) and HbIII(Lp)) that remain unaffected by the presence of H(2)S. The mechanisms used by these three proteins for their function, including ligand control, remain unknown. The crystal structure of oxygen-bound HbII(Lp) shows a dimeric oxyHbII(Lp) where oxygen is tightly anchored to the heme through hydrogen bonds with Tyr(30)(B10) and Gln(65)(E7). The heme group is buried farther within HbII(Lp) than in HbI(Lp). The proximal His(97)(F8) is hydrogen bonded to a water molecule, which interacts electrostatically with a propionate group, resulting in a Fe-His vibration at 211 cm(-1). The combined effects of the HbII(Lp) small heme pocket, the hydrogen bonding network, the His(97) trans-effect, and the orientation of the oxygen molecule confer stability to the oxy-HbII(Lp) complex. Oxidation of HbI(Lp) Phe(B10) --> Tyr and HbII(Lp) only occurs when the pH is decreased from pH 7.5 to 5.0. Structural and resonance Raman spectroscopy studies suggest that HbII(Lp) oxygen binding and transport to the host bacteria may be regulated by the dynamic displacements of the Gln(65)(E7) and Tyr(30)(B10) pair toward the heme to protect it from changes in the heme oxidation state from Fe(II) to Fe(III).  相似文献   

19.
A series of new iron(III) and copper(II) complexes of bovine serum transferrin (BTf), with carbonate and/or oxalate as the synergistic anion, are presented. The complexes [Fe(2)(CO(3))(2)BTf], [Fe(2)(C(2)O(4))(2)BTf], [Cu(2)(CO(3))(2)BTf] and [Cu(C(2)O(4))BTf] were prepared by standard titrimetric techniques. The oxalate derivatives were also obtained from the corresponding carbonate complexes by anion-displacement. The site-preference of the transition metal-oxalate synergism has facilitated the preparation and isolation of the mononuclear complex [Cu(C(2)O(4))BTf], the mixed-anion complexes [Cu(2)(CO(3))(C(2)O(4))BTf] and [Fe(2)(CO(3))(C(2)O(4))BTf] and the mixed-metal complex [FeCu(C(2)O(4))(2)BTf]. The sensitivity of electron paramagnetic resonance (EPR) spectroscopy to the nature of the synergistic anions at the specific-binding sites of the transferrins has made this physical technique particularly indispensable to this study. None of the other members of the transferrin family of proteins has ever been demonstrated to bind the ferric and cupric ions one after the other, each occupying a separate specific-binding site of the same transferrin molecule, as a response to the coordination restrictions imposed by the oxalate ion. The bathochromic shift of the visible p(pi)-d(pi*) CT band for iron(III)-BTf and the hypsochromic shift of the p(pi)-d(sigma*) CT band for copper(II)-BTf, on replacing carbonate by oxalate as the associated anion, are consistent with the relative positions of these anionic ligands in the spectrochemical series and the nature of the d-type acceptor orbitals involved in the CT transitions. The binding and spectroscopic properties of bovine serum transferrin--a serum transferrin--very nearly mirror those of human serum transferrin, but differ significantly from those of human lactoferrin.  相似文献   

20.
The NADP-dependent non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase from Streptococcus mutans (abbreviated Sm-ALDH) belongs to the aldehyde dehydrogenase (ALDH) family. Its catalytic mechanism proceeds via two steps, acylation and deacylation. Its high catalytic efficiency at neutral pH implies prerequisites relative to the chemical mechanism. First, the catalytic Cys284 should be accessible and in a thiolate form at physiological pH to attack efficiently the aldehydic group of the glyceraldehyde-3-phosphate (G3P). Second, the hydride transfer from the hemithioacetal intermediate toward the nicotinamide ring of NADP should be efficient. Third, the nucleophilic character of the water molecule involved in the deacylation should be strongly increased. Moreover, the different complexes formed during the catalytic process should be stabilised.The crystal structures presented here (an apoenzyme named Apo2 with two sulphate ions bound to the catalytic site, the C284S mutant holoenzyme and the ternary complex composed of the C284S holoenzyme and G3P) together with biochemical results and previously published apo and holo crystal structures (named Apo1 and Holo1, respectively) contribute to the understanding of the ALDH catalytic mechanism.Comparison of Apo1 and Holo1 crystal structures shows a Cys284 side-chain rotation of 110 degrees, upon cofactor binding, which is probably responsible for its pK(a) decrease. In the Apo2 structure, an oxygen atom of a sulphate anion interacts by hydrogen bonds with the NH2 group of a conserved asparagine residue (Asn154 in Sm-ALDH) and the Cys284 NH group. In the ternary complex, the oxygen atom of the aldehydic carbonyl group of the substrate interacts with the Ser284 NH group and the Asn154 NH2 group. A substrate isotope effect on acylation is observed for both the wild-type and the N154A and N154T mutants. The rate of the acylation step strongly decreases for the mutants and becomes limiting. All these results suggest the involvement of Asn154 in an oxyanion hole in order to stabilise the tetrahedral intermediate and likely the other intermediates of the reaction. In the ternary complex, the cofactor conformation is shifted in comparison with its conformation in the C284S holoenzyme structure, likely resulting from its peculiar binding mode to the Rossmann fold (i.e. non-perpendicular to the plane of the beta-sheet). This change is likely favoured by a characteristic loop of the Rossmann fold, longer in ALDHs than in other dehydrogenases, whose orientation could be constrained by a conserved proline residue. In the ternary and C284S holenzyme structures, as well as in the Apo2 structure, the Glu250 side-chain is situated less than 4 A from Cys284 or Ser284 instead of 7 A in the crystal structure of the wild-type holoenzyme. It is now positioned in a hydrophobic environment. This supports the pK(a) assignment of 7.6 to Glu250 as recently proposed from enzymatic studies.  相似文献   

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