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 31 毫秒
1.
Dmitriy N. Shevela 《BBA》2006,1757(4):253-261
It is shown that the hydrazine-induced transition of the water-oxidizing complex (WOC) to super-reduced S-states depends on the presence of bicarbonate in the medium so that after a 20 min treatment of isolated spinach thylakoids with 3 mM NH2NH2 at 20 °C in the CO2/HCO3-depleted buffer the S-state populations are: 42% of S−3, 42% of S−2, 16% of S−1 and even formal S−4 state is reached, while in the presence of 2 mM NaHCO3, the same treatment produces 30% of S−3, 38% of S−2, and 32% of S−1 and there is no indication of the S−4 state. Bicarbonate requirement for the oxygen-evolving activity, very low in untreated thylakoids, considerably increases upon the transition of the WOC to the super-reduced S-states, and the requirement becomes low again when the WOC returns back to the normal S-states using pre-illumination. The results are discussed as a possible indication of ligation of bicarbonate to manganese ions within the WOC.  相似文献   

2.
It is known that a 1,2,3-triazolato-bridged dinuclear platinum(II) complex, [{cis-Pt(NH3)2}2(μ-OH)(μ-1,2,3-ta-N 1,N 2)](NO3)2 (AMTA), shows high in vitro cytotoxicity against several human tumor cell lines and circumvents cross-resistance to cisplatin. In the present study, we examined a dose- and time-dependent effect of AMTA on the higher-order structure of a large DNA, T4 phage DNA (166 kbp), by adapting single-molecule observation with fluorescence microscopy. It was found that AMTA induces the shrinking of DNA into a compact state with a much higher potency than cisplatin. From a quantitative analysis of the Brownian motion of individual DNA molecules in solution, it became clear that the density of a DNA segment in the compact state is about 2,000 times greater than that in the absence of AMTA. Circular dichroism spectra suggested that AMTA causes a transition from the B to the C form in the secondary structure of DNA, which is characterized by fast and slow processes. Electrophoretic measurements indicated that the binding of AMTA to supercoiled DNA induces unwinding of the double helix. Our results indicate that AMTA acts on DNA through both electrostatic interaction and coordination binding; the former causes a fast change in the secondary structure from the B to the C form, whereas the latter promotes shrinking in the higher-order structure as a relatively slow kinetic process. The shrinking effect of AMTA on DNA is attributable to the possible increase in the number of bridges along a DNA molecule. It is concluded that AMTA interacts with DNA in a manner markedly different from that of cisplatin.  相似文献   

3.
Stress-induced arrest of ventilatory motor pattern generation is tightly correlated with an abrupt increase in extracellular potassium concentration ([K+]o) within the metathoracic neuropil of the locust, Locusta migratoria. Na+/K+-ATPase inhibition with ouabain elicits repetitive surges of [K+]o that coincide with arrest and recovery of motor activity. Here we show that ouabain induces repetitive [K+]o events in a concentration-dependent manner. 10−5 M, 10−4 M, and 10−3 M ouabain was bath-applied in semi-intact locust preparations. 10−4 M and 10−3 M ouabain reliably induced repetitive [K+]o events whereas 10−5 M ouabain had no significant effect. In comparison to 10−4 M ouabain, 10−3 M ouabain increased the number and hastened the time to onset of repetitive [K+]o waves, prolonged [K+]o event duration, increased resting [K+]o, and diminished the absolute value of [K+]o waves. Recovery of motor patterning following [K+]o events was less likely in 10−3 M ouabain. In addition, we show that K+ channel inhibition using TEA suppressed the onset and decreased the amplitude of ouabain-induced repetitive [K+]o waves. Our results demonstrate that ventilatory circuit function in the locust CNS is dependent on the balance between mechanisms of [K+] accumulation and [K+] clearance. We suggest that with an imbalance in favour of accumulation the system tends towards a bistable state with transitions mediated by positive feedback involving voltage-dependent K+ channels.  相似文献   

4.
Four-coordinate 1:2 gold(I) complex salts with cis-bis(diphenylphosphino)ethene, [Au(dppey)2]X have been synthesized for X = PF6, CF3SO3, BF4, Cl, Br and BPh4 and characterized by NMR spectroscopy and electrospray mass spectrometry. Single crystal X-ray structure determinations show the BF4, Cl and Br complexes to be isostructural, although with different degrees of hydration, while the BPh4 complex crystallizes as an acetone solvate with two molecules in the asymmetric unit. The Au(P-P)2 core for the BF4, Cl and Br complexes adopts D2 symmetry with Au-P bond lengths 2.3980(7)-2.4009(7) Å and inter-ligand P-Au-P angles 114.78(2)-127.82(2))°. The Au(P-P)2 core in the BPh4 complex is unsymmetrical with Au-P bond lengths 2.364(1)-2.420(1) Å and inter-ligand P-Au-P angles 104.76(5)-137.50(4)°. In vitro cytotoxicity studies show the PF6, CF3SO3, BF4, Cl, Br and I complexes to be potent and selective growth inhibitors of the human cell lines MCF7 (hormone-dependent breast cancer), MDA-MB-231 (hormone-independent breast cancer), MM96L (melanoma), CI80-13S (cisplatin resistant ovarian cancer) and a normal cell line NFF (neonatal foreskin fibroblasts), achieving IC50 values between 13 and 196 nM. The halogen and triflate salts were approximately twice as potent towards the MCF7 and MDA-MB-231 cell lines compared to the PF6 and BF4 derivatives; while the cytotoxicity of all complexes towards the sensitive CI80-13S and MM96L cancer cell lines was approximately 10-fold greater than that displayed towards the normal human cell line (NFF).  相似文献   

5.
A macrocyclic ligand possessing a donor set of {N3S2} synthesised via Cs+-templation, 4-(pyridin-2-ylmethyl)-1,7-dithia-4,10-diazacyclododecane (L) and its Cu(II) complex, [CuL(NCMe)]2+ (6), are described. This Cu(II) complex interacts with a range of anions, F, Cl, Br, I, HCOO, AcO, CO32−, NO3, C2O42−, H2PO4, SCN, CN, BF4. Of the investigated anions, I, SCN, and CN, show strong interaction with the Cu(II) centre as indicated by their spectral variations. The iodide particularly demonstrates distinct change in colour. This change originates from a newly appeared band at 471 nm upon iodide binding, which arises from the ligand (I) to Cu(II) charge transfer (LMCT) in the iodide-substituted Cu(II) complex, [CuLI]+ (7). All organic compounds are characterised by NMR spectroscopy and/or microanalysis. The identities of the two Cu(II) complexes are confirmed by using microanalysis and the complex 6 is crystallographically analysed.  相似文献   

6.
The complex of xylan and iodine and its formation in a solution of xylan, CaCl2, and I2 + KI was investigated by UV/Vis, second-derivative UV/Vis, and Raman spectroscopy. The complex forms only at very high concentrations of CaCl2, suggesting that when the water available in the solution is not sufficient to fully hydrate the calcium cation the chelation with the hydroxyl groups of the xylan can occur. The electronic spectra indicate that iodine is present in the form of three linear polyiodides I93−, I113−, and I133− structures, which the Raman spectra show to be linear aggregates of the I3 and I5 substructures. Iodide concentration has a significant influence on the relative population of I93−, I113−, and I133−, as well as I3 and I5, which lead to changes in both the UV/Vis absorption maxima shifts and changes in the Raman spectra. The key difference between this system of complexes with the linear polyiodide aggregates and that of amylose is that the longest aggregate observed with the amylose system, the I153− polyanion, is not observed with the xylans. This indicates that the ordered arrays in the xylan-iodine complex do not exceed 4 nm in length. It is not possible to conclude at this time whether the ordered segment of the xylan molecule is linear or helical. If it is linear the length of the longest ordered arrays would be eight xylose residues. The number would exceed eight if the xylan molecule were helically wound.  相似文献   

7.
The SS bond-activation of diorganyl disulfide by the anionic metal carbonyl fragment [Mn(CO)5] gives rise to an extensive chemistry. Oxidative decarbonylation addition of 2,2′-dithiobis(pyridine-N-oxide) to [Mn(CO)5], followed by chelation and metal-center oxidation, led to the formation of [MnII(SC5H4NO)3] (1). The effective magnetic moment in solid state by SQUID magnetometer was 5.88 μB for complex 1, which is consistent with the MnII having a high-spin d5 electronic configuration in an octahedral ligand field. The average Mn(II)S, SC and NO bond lengths of 2.581(1), 1.692(4) and 1.326(4) Å, respectively, indicate that the negative charge of the bidentate 1-oxo-2-thiopyridinato [SC5H4NO] ligand in complex 1 is mainly localized on the oxygen atom. The results are consistent with thiolate-donor [SC5H4NO] stabilization of the lower oxidation state of manganese (Mn(I)), while the O,S-chelating [SC5H4NO] ligand enhances the stability of manganese in the higher oxidation state (Mn(II)). Activation of SS bond as well as OH bond of 2,2′-dithiosalicylic acid by [Mn(CO)5] yielded [(CO)3Mn(μ-SC6H4C(O)O)2Mn(CO)3]2− (4). Oxidative addition of bis(o-benzamidophenyl) disulfide to [Mn(CO)5] resulted in the formation of cis-[Mn(CO)4(SR)2] (R=C6H4NHCOPh) which was employed as a chelating metallo ligand to synthesize heterotrinuclear [(CO)3Mn(μ-SR)3Co(μ-SR)3Mn(CO)3] (8) possessing a homoleptic hexathiolatocobalt(III) core.  相似文献   

8.
Sublevel structure of the 8S7/2 electronic ground state of anionic bis(phthalocyaninato)gadolinium(III) has been determined by simulation analysis of an ESR spectrum in frozen solution. The simplex multidimensional minimization algorithm was employed to find the zero-field-splitting parameter set giving the minimum RMS error from the observed spectrum. The parameter set {B20,B40,B60} for the potential of D4d symmetry has been determined to be ±{(1.54 ± 0.01)× 10−2 cm−1, (0.9 ± 0.1)× 10−4 cm−1, (−0.6 ± 0.9)× 10−6 cm−1}. The energy difference between the lowest and highest sublevels has been found to be about 0.5 cm−1.  相似文献   

9.
In this study we assessed the growth, morphological responses, and N uptake kinetics of Salvinia natans when supplied with nitrogen as NO3, NH4+, or both at equimolar concentrations (500 μM). Plants supplied with only NO3 had lower growth rates (0.17 ± 0.01 g g−1 d−1), shorter roots, smaller leaves with less chlorophyll than plants supplied with NH4+ alone or in combination with NO3 (RGR = 0.28 ± 0.01 g g−1 d−1). Ammonium was the preferred form of N taken up. The maximal rate of NH4+ uptake (Vmax) was 6–14 times higher than the maximal uptake rate of NO3 and the minimum concentration for uptake (Cmin) was lower for NH4+ than for NO3. Plants supplied with NO3 had elevated nitrate reductase activity (NRA) particularly in the roots showing that NO3 was primarily reduced in the roots, but NRA levels were generally low (<4 μmol NO2 g−1 DW h−1). Under natural growth conditions NH4+ is probably the main N source for S. natans, but plants probably also exploit NO3 when NH4+ concentrations are low. This is suggested based on the observation that the plants maintain high NRA in the roots at relatively high NH4+ levels in the water, even though the uptake capacity for NO3 is reduced under these conditions.  相似文献   

10.
Low temperature NMR spectroscopy has been used to characterize the mixtures formed in the oligomerization reactions of the unsaturated complex [Re2(μ-H)2(CO)8] (1), promoted by hydrido-carbonyl rhenates. Three families of chain clusters, constituted by Re(CO)4 units connected through Re(μ-H)Re interactions, have been obtained. The first one, of general formula [(CO)5Re-{HRe(CO)4}2n+1], was formed using [HRe2(CO)9] as promoter. The nature of the products was confirmed by 13C NMR of 13CO enriched samples. The formation of Re6 and Re8 chain clusters was recognized. The other two families have general formula [H-{HRe(CO)4}2n] and [H-{HRe(CO)4}2n+1] and were obtained using as initiators [HB(sBu)3] or [H2Re(CO)4], respectively. Mixtures of oligomers with mean chain lengths higher than 10 have been observed. The addition of a strong acid caused H2 evolution, leading back to the “monomer” 1. For all the three families, each oligomerization step was reversible, with the longer oligomers favoured at the lowest temperatures, where, however, the reactions were very slow, usually preventing the attainment of the equilibrium. Variable temperature NMR spectra revealed a dynamic process involving the terminal H2Re(CO)4 moiety(ies), that simultaneously exchanges terminal/bridging hydrides and the carbonyls trans to them (ΔG# 41-44 kJ mol−1). At room temperature, the more hydrogen-rich chain clusters also underwent dehydrogenation/cyclization reactions.  相似文献   

11.
We have studied the protonation equilibria of a dicopper(II) complex [Cu2(μ-OH)(C21H33ON6)](ClO4)2·H2O, (1), in aqueous solution, its interactions with DNA, its cytotoxic activity, and its uptake in tumoral cells. C21H33ON6 corresponds to the ligand 4-methyl-2,6-bis[(6-methyl-1,4-diazepan-6-yl)iminomethyl]phenol. From spectrophotometric data the following pKa values were calculated 3.27, 4.80 and 6.10. Complex 1 effectively promotes the hydrolytic cleavage of double-strand plasmid DNA under anaerobic and aerobic conditions. The following kinetic parameters were calculated kcat of 2.73 × 10−4 s−1, KM of 1.36 × 10−4 M and catalytic efficiency of 2.01 s−1 M−1, a 2.73 × 107 fold increase in the rate of the reaction compared to the uncatalyzed hydrolysis rate of DNA. Competition assays with distamycin reveal minor groove binding. Complex 1 inhibited the growth of two tumoral cell lines, GLC4 and K562, with the IC50 values of 14.83 μM and 34.21 μM, respectively. There is a good correlation between cell growth inhibition and intracellular copper content. When treated with 1, cells accumulate approximately twice as much copper as with CuCl2. Copper-DNA adducts are formed inside cells when they are exposed to the complex. In addition, at concentrations that compound 1 inhibits tumoral cell growth it does not affect macrophage viability. These results show that complex 1 has a good therapeutic prospect.  相似文献   

12.
A new one-dimensional copper(II) polymer, [Cu4(dmapox)2(SCN)4(CH3CH2OH)2]n · 2nCH3CH2OH, where dmapox is the dianion of N,N′-bis[3-(dimethylamino)propyl]oxamide, was synthesized and characterized by elemental analysis, conductivity measurement, IR and electronic spectral studies. The crystal structure of the copper(II) complex has been determined by X-ray single-crystal diffraction. The complex crystallizes in triclinic, space group and exhibits infinite one-dimensional copper(II) polymeric chain bridged both by the bis-tridentate μ-trans-dmapox and μ-1,3-thiocyanate ligands. The environment around the copper(II) atom can be described as distorted square-pyramid. The Cu···Cu separations through μ-trans-oxamidate and thiocyanate bridges are 5.245(5) Å (Cu1-Cu1i)(i = −x+1, −y, −z+1), 5.262(4) Å (Cu2-Cu2ii)(ii = −x,−y, −z+1) and 6.022(3) Å (Cu1-Cu2), respectively. The interaction of the copper(II) complex with herring sperm DNA (HS-DNA) has been investigated by using absorption and emission spectral and electrochemical techniques and viscometry. The results reveal that the copper(II) complex interacts with the DNA in the mode of groove binding with the intrinsic binding constant of 2.38 × 105 M−1.  相似文献   

13.
The first excited singlet state (S1) of carotenoids (also termed 2Ag) plays a key role in photosynthetic excitation energy transfer due to its close proximity to the S1 (Qy) level of chlorophylls. The determination of carotenoid 2Ag energies by optical techniques is difficult; transitions from the ground state (S0, 1Ag) to the 2Ag state are forbidden (“optically dark”) due to parity (g ← //→ g) as well as pseudo-parity selection rules (− ← //→ −). Of particular interest are S1 energies of the so-called xanthophyll-cycle pigments (violaxanthin, antheraxanthin and zeaxanthin) due to their involvement in photoprotection in plants. Previous determinations of S1 energies of violaxanthin and zeaxanthin by different spectroscopic techniques vary considerably. Here we present an alternative approach towards elucidation of the optically dark states of xanthophylls by near-edge X-ray absorption fine structure spectroscopy (NEXAFS). The indication of at least one π* energy level (about 0.5 eV below the lowest 1Bu+ vibronic sublevel) has been found for zeaxanthin. Present limitations and future improvements of NEXAFS to study optically dark states of carotenoids are discussed. NEXAFS combined with simultaneous optical pumping will further aid the investigation of these otherwise hardly accessible states.  相似文献   

14.
Phototropins, major blue-light receptors in plants, are sensitive to blue light through a pair of flavin mononucleotide (FMN)-binding light oxygen and voltage (LOV) domains, LOV1 and LOV2. LOV2 undergoes a photocycle involving light-driven covalent adduct formation between a conserved cysteine and the FMN C(4a) atom. Here, the primary reactions of Avena sativa phototropin 1 LOV2 (AsLOV2) were studied using ultrafast mid-infrared spectroscopy and quantum chemistry. The singlet excited state (S1) evolves into the triplet state (T1) with a lifetime of 1.5 ns at a yield of ∼50%. The infrared signature of S1 is characterized by absorption bands at 1657 cm−1, 1495-1415 cm−1, and 1375 cm−1. The T1 state shows infrared bands at 1657 cm−1, 1645 cm−1, 1491-1438 cm−1, and 1390 cm−1. For both electronic states, these bands are assigned principally to C=O, C=N, C-C, and C-N stretch modes. The overall downshifting of C=O and C=N bond stretch modes is consistent with an overall bond-order decrease of the conjugated isoalloxazine system upon a π-π transition. The configuration interaction singles (CIS) method was used to calculate the vibrational spectra of the S1 and T1 excited ππ states, as well as respective electronic energies, structural parameters, electronic dipole moments, and intrinsic force constants. The harmonic frequencies of S1 and T1, as calculated by the CIS method, are in satisfactory agreement with the evident band positions and intensities. On the other hand, CIS calculations of a T1 cation that was protonated at the N(5) site did not reproduce the experimental FMN T1 spectrum. We conclude that the FMN T1 state remains nonprotonated on a nanosecond timescale, which rules out an ionic mechanism for covalent adduct formation involving cysteine-N(5) proton transfer on this timescale. Finally, we observed a heterogeneous population of singly and doubly H-bonded FMN C(4)=O conformers in the dark state, with stretch frequencies at 1714 cm−1 and 1694 cm−1, respectively.  相似文献   

15.
The folding mechanism and stability of dimeric formate dehydrogenase from Candida methylica was analysed by exposure to denaturing agents and to heat. Equilibrium denaturation data yielded a dissociation constant of about 10−13 M for assembly of the protein from unfolded chains and the kinetics of refolding and unfolding revealed that the overall process comprises two steps. In the first step a marginally stable folded monomeric state is formed at a rate (k1) of about 2 × 10−3 s−1 (by deduction k−1 is about10−4 s−1) and assembles into the active dimeric state with a bimolecular rate constant (k2) of about 2 × 104 M−1 s−1. The rate of dissociation of the dimeric state in physiological conditions is extremely slow (k−2 ∼ 3 × 10−7 s−1).  相似文献   

16.

Background

DNase antibodies can play an important role in the pathogenesis of different autoimmune pathologies.

Methods

An immunoglobulin light chain phagemid library derived from peripheral blood lymphocytes of patients with systemic lupus erythematosus (SLE) was used. The small pools of phage particles displaying DNA binding light chains with different for DNA were isolated by affinity chromatography on DNA-cellulose and the fraction eluted with 0.5 M NaCl was used for preparation of individual monoclonal light chains (MLChs, 28 kDa). Forty-five of 451 individual colonies were randomly chosen for a study of MLChs with DNase activity. The clones were expressed in Escherichia coli in a soluble form, and MLChs were purified by metal chelating chromatography followed by gel filtration, and studied in detail.

Results

Fifteen of 45 MLChs efficiently hydrolyzed DNA, and fourteen of them demonstrated various optimal concentrations of KCl or NaCl in a 1–100 mM range and showed one or two pH optima in a 4.8–9.1 range. All MLChs were dependent on divalent metal cations: the ratio of relative DNase activity in the presence of Mn2 +, Ca2 +, Mg2 +, Ni2 +, Zn2 +, Cu2 +, and Co2 + was individual for each MLCh preparation. Fourteen MLChs demonstrated a comparable affinity for DNA (260–320 nM), but different kcat values (0.02–0.7 min− 1).

Conclusions

These observations suggest an extreme diversity of DNase abzymes from SLE patients.

General significance

SLE light chain repertoire can serve as a source of new types of DNases.  相似文献   

17.
Biodiversity and ecosystem functioning experiments have demonstrated that plant biomass of species grown in mixtures is often greater than plant biomass of monocultures (i.e., mixtures over yield). While we understand that plant species utilize resources differently, how a combination of species increases resource use and productivity is not well known, especially in wetland ecosystems. Here, we used a mesocosm experiment to explore diversity effects on plant biomass production and to examine the role of N partitioning as a mechanism for overyielding in wetland ecosystems. Plant functional groups (FGs) represented the unit of diversity, and we included five levels of diversity (0-4 FGs). To test for N partitioning, we used a stable isotope technique to determine niche breadth and proportion similarity of inorganic N use (NO3 and NH4+) for individual FGs as well as mixtures containing 3 and 4 FGs. We found that total plant biomass increased in the first season from an average of 290 ± 60 SE g ash-free dry mass (AFDM) m−2 at the 1 FG level to 490 ± 70 g AFDM m−2 at the 4 FG level and in the second season from an average of 560 ± 80 g AFDM m−2 at the 1 FG level to 1000 ± 90 g AFDM m−2 at the 4 FG level indicating overyielding. Plant species comprising the majority of mesocosm biomass demonstrated preferential uptake of 15NO3, while species with relatively less biomass (e.g., Acorus calamus and Carex crinita) preferred 15NH4+. Concentrations of 15N in biomass increased with FG richness, but only in the 15NO3 treatment. Niche breadth did not vary among levels of FG richness. We observed a greater niche overlap with an increase of FGs, with species taking up greater proportion of 15NO3 than 15NH4+. Our results indicate that plant overyielding in wetland mesocosms is not the result of niche partitioning of N chemical forms, but is associated with greater uptake of NO3.  相似文献   

18.
The peripheral light-harvesting complex of photosystem I contains red chlorophylls (Chls) that, unlike the typical antenna Chls, absorb at lower energy than the primary electron donor P700. It has been shown that the red-most absorption band arises from two excitonically coupled Chls, although this interaction alone cannot explain the extreme red-shifted emission (25 nm, ∼480 cm−1 for Lhca4 at 4 K) that the red Chls present. Here, we report the electric field-induced absorption changes (Stark effect) on the Qy region of the Lhca4 complex. Two spectral forms, centered around 690 nm and 710 nm, were necessary to describe the absorption and Stark spectra. The analysis of the lowest energy transition yields a high value for the change in dipole moment, Δμ710nm ≈ 8 Df−1, between the ground and excited states as compared with monomeric, Δμ = 1 D, or dimeric, Δμ = 5 D, Chl a in solution. The high value of the Δμ demonstrates that the origin of the red-shifted emission is the mixing of the lowest exciton state with a charge-transfer state of the dimer. This energetic configuration, an excited state with charge-transfer character, is very favorable for the trapping and dissipation of excitations and could be involved in the photoprotective mechanism(s) of the photosystem I complex.  相似文献   

19.
Treatment of 3-(4-carboxyphenylhydrazono)pentane-2,4-dione (HL) with transition metal ions afforded four novel complexes, [Zn(L)(μ2-OOCCH3)(H2O)]n (1), [Zn(L)2(MeOH)4] (2), {[Cd4(η2-L)4(μ2-η2-L)4(H2O)4(MeOH)2]·MeOH} (3) and [Cd(η2-L)(μ2-η2-OOCCH3)(H2O)2]n (4). Their crystal structures have been characterized by single-crystal X-ray crystallography. In polymer 1, the acetate anions bridge the Zn(II) ions forming an infinite one-dimensional (1-D) chain with L units acting as monodentate ligands in the side chain. In mononuclear complex 2, two L ligands act as monodentate fashion to coordinate to the Zn(II) ion. In its solid-state structure, [Zn(L)2(MeOH)4] groups are joined together by hydrogen bonds forming a three-dimensional (3-D) supramolecular network. In tetranuclear complex 3, four Cd(II) ions are linked by four μ2-η2-L ligands, and chelated by another four L ligands, respectively. In polymer 4, the acetate anions bridge the Cd(II) ions leading to a 1-D chain containing chelating L units in the side chain.  相似文献   

20.
To characterize driving forces and driven processes in formation of a large-interface, wrapped protein-DNA complex analogous to the nucleosome, we have investigated the thermodynamics of binding the 34-base pair (bp) H′ DNA sequence to the Escherichia coli DNA-remodeling protein integration host factor (IHF). Isothermal titration calorimetry and fluorescence resonance energy transfer are applied to determine effects of salt concentration [KCl, KF, K glutamate (KGlu)] and of the excluded solute glycine betaine (GB) on the binding thermodynamics at 20 °C. Both the binding constant Kobs and enthalpy ΔH°obs depend strongly on [salt] and anion identity. Formation of the wrapped complex is enthalpy driven, especially at low [salt] (e.g., ΔHoobs = − 20.2 kcal·mol− 1 in 0.04 M KCl). ΔH°obs increases linearly with [salt] with a slope (dΔH°obs/d[salt]), which is much larger in KCl (38 ± 3 kcal·mol− 1 M− 1) than in KF or KGlu (11 ± 2 kcal·mol− 1 M− 1). At 0.33 M [salt], Kobs is approximately 30-fold larger in KGlu or KF than in KCl, and the [salt] derivative SKobs = dlnKobs/dln[salt] is almost twice as large in magnitude in KCl (− 8.8 ± 0.7) as in KF or KGlu (− 4.7 ± 0.6).A novel analysis of the large effects of anion identity on Kobs, SKobs and on ΔH°obs dissects coulombic, Hofmeister, and osmotic contributions to these quantities. This analysis attributes anion-specific differences in Kobs, SKobs, and ΔH°obs to (i) displacement of a large number of water molecules of hydration [estimated to be 1.0(± 0.2) × 103] from the 5340 Å2 of IHF and H′ DNA surface buried in complex formation, and (ii) significant local exclusion of F and Glu from this hydration water, relative to the situation with Cl, which we propose is randomly distributed. To quantify net water release from anionic surface (22% of the surface buried in complexation, mostly from DNA phosphates), we determined the stabilizing effect of GB on Kobs: dlnKobs/d[GB]  = 2.7 ± 0.4 at constant KCl activity, indicating the net release of ca. 150 H2O molecules from anionic surface.  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