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.
In the recently proposed local-access model for proton transfers in the bacteriorhodopsin transport cycle (Brown et al. 1998. Biochemistry. 37:3982-3993), connection between the retinal Schiff base and Asp85 (in the extracellular direction) and Asp96 (in the cytoplasmic direction)is maintained as long as the retinal is in its photoisomerized state. The directionality of the proton translocation is determined by influences in the protein that make Asp85 a proton acceptor and, subsequently, Asp96 a proton donor. The idea of concurrent local access of the Schiff base in the two directions is now put to a test in the photocycle of the D115N/D96N mutant. The kinetics had suggested that there is a single sequence of intermediates, L<-->M1<-->M2<-->N, and the M2-->M1 reaction depends on whether a proton is released to the extracellular surface. This is now confirmed. We find that at pH 5, where proton release does not occur, but not at higher pH, the photostationary state created by illumination with yellow light contains not only the M1 and M2 states, but also the L and the N intermediates. Because the L and M1 states decay rapidly, they can be present only if they are in equilibrium with later intermediates of the photocycle. Perturbation of this mixture with a blue flash caused depletion of the M intermediate, followed by its partial recovery at the expense of the L state. The change in the amplitude of the C=O stretch band at 1759 cm-1 demonstrated protonation of Asp85 in this process. Thus, during the reequilibration the Schiff base lost its proton to Asp85. Because the N state, also present in the mixture, arises by protonation of the Schiff base from the cytoplasmic surface, these results fulfill the expectation that under the conditions tested the extracellular access of the Schiff base would not be lost at the time when there is access in the cytoplasmic direction. Instead, the connectivity of the Schiff base flickers rapidly (with the time constant of the M1<-->M2 equilibration) between the two directions during the entire L-to-N segment of the photocycle.  相似文献   

2.
J Heberle  D Oesterhelt    N A Dencher 《The EMBO journal》1993,12(10):3721-3727
Surface bound pH indicators were applied to study the proton transfer reactions in the mutant Asp85-->Glu of bacteriorhodopsin in the native membrane. The amino acid replacement induces a drastic acceleration of the overall rise of the M intermediate. Instead of following this acceleration, proton ejection to the extracellular membrane surface is not only two orders of magnitude slower than M formation, it is also delayed as compared with the wild-type. This demonstrates that Asp85 not only accepts the proton released by the Schiff's base but also regulates very efficiently proton transfer within the proton release chain. Furthermore, Asp85 might be the primary but is not the only proton acceptor/donor group in the release pathway. The Asp85-->Glu substitution also affects the proton reuptake reaction at the cytoplasmic side, although Asp85 is located in the proton release pathway. Proton uptake is slower in the mutant than in the wild-type and occurs during the lifetime of the O intermediate. This demonstrates a feed-back mechanism between Asp85 and the proton uptake pathway in bacteriorhodopsin.  相似文献   

3.
Bacteriorhodopsin (BR), a membrane protein found in Halobacterium salinarum, functions as a light-driven proton pump. The Schiff base region has a quadrupolar structure with positive charges located at the protonated Schiff base and Arg82, and the counterbalancing negative charges located at Asp85 and Asp212. The quadropole inside the protein is stabilized by three water molecules, forming a roughly planar pentagonal cluster composed of these waters and two oxygens of Asp85 and Asp212 (one from each carboxylate side chain). It is known that BR lacks proton-pumping activity if Asp85 or Asp212 is neutralized by mutation, but binding of Cl- has different functional effects in mutants at these positions. Binding of Cl- to D85T converts into a chloride ion pump (Sasaki, J., Brown, L. S., Chon, Y.-S., Kandori, H., Maeda, A., Needleman, R., and Lanyi, J. K. (1995) Science 269, 73-75). On the other hand, photovoltage measurements suggested that binding of Cl- to D212N restores the proton-pumping activity at low pH (Moltke, S., Krebs, M. P., Mollaaghababa, R., Khorana, H. G., and Heyn, M. P. (1995) Biophys. J. 69, 2074-2083). In this paper, we studied halide-bound D212N mutant BR in detail. Light-induced pH changes in a suspension of proteoliposomes containing D212N(Cl-) at pH 5 clearly showed that Cl- restores the proton-pumping activity. Spectral blue-shift induced by halide binding to D212N indicates that halides affect the counterion of the protonated Schiff base, whereas much smaller halide dependence of the lambdamax than in D85T suggests that the binding site is distant from the chromophore. In fact, the K minus BR difference Fourier-transform infrared (FTIR) spectra of D212N at 77 K exhibit little halide dependence for vibrational bands of retinal and protein. The only halide-dependent bands were the C=N stretch of Arg82 and some water O-D stretches, suggesting that these groups constitute a halide-binding pocket. A strongly hydrogen-bonded water molecule is observed for halide-bound D212N, but not for halide-free D212N, which is consistent with our hypothesis that such a water molecule is a prerequisite for proton-pumping activity of rhodopsins. We concluded that halide binding near Arg82 in D212N restores the water-containing hydrogen-bonding network in the Schiff base region. In particular, the ion pair formed by the Schiff base and Asp85 through a strongly hydrogen-bonded water is essential for the proton-pumping activity of this mutant and may be controlled by the halide binding to the distant site.  相似文献   

4.
Time-resolved photovoltage measurements were performed with the acid-blue (bR605A) and acid-purple (bR565A) forms of bacteriorhodopsin (bR) in the time range from 25 ns to 100 s. The bR605A and bR565A pigments were formed by titration with H2SO4 in the absence and presence of 150 mM KCI, respectively. Qualitatively the kinetics of the charge displacement in these two states are similar and consist of two fast phases in one direction (100 ns bandwidth limited and approximately 1 microsecond) followed by a decay in the opposite direction via one component for bR605A (4.4 +/- 0.6 ms) or two components for bR565A (33 +/- 8 microseconds and 3.6 +/- 0.5 ms). The transient photovoltage signal returns exactly to the initial value after several milliseconds, well before the passive discharge of the electrical measuring system at 2 s. We conclude that no net charge transfer occurs in either bR605A or bR565A. The direction of the fast components is opposite that of net proton translocation in bR at pH 7. So, if the charge that moves back and forth is due to a proton, it moves first in the direction of the cytoplasmic side of the membrane (< 1 microsecond) and returns to its initial position via the 4.4 ms (bR605A) or the 33 microseconds and 3.6 ms (bR565A) decay components. The amplitude of the charge motion in both low pH forms is too large to be due to isomerization alone and is comparable to one of the major components in bR at pH 7.2  相似文献   

5.
Recombinant human phenylalanine hydroxylase (hPAH) expressed in Escherichia coli for 24 h at 28 degrees C has been found by two-dimensional electrophoresis to exist as a mixture of four to five molecular forms as a result of nonenzymatic deamidation of labile Asn residues. The multiple deamidations alter the functional properties of the enzyme including its affinity for l-phenylalanine and tetrahydrobiopterin, catalytic efficiency, and substrate inhibition and also result in enzyme forms more susceptible to limited tryptic proteolysis. Asn(32) in the regulatory domain deamidates very rapidly because of its nearest neighbor amino acid Gly(33) (Solstad, T., Carvalho, R. N., Andersen, O. A., Waidelich, D., and Flatmark, T. (2003) Eur. J. Biochem., in press). Matrix-assisted laser desorption/ionization time of flight-mass spectrometry of the tryptic peptides in the catalytic domain of a 24-h (28 degrees C) expressed enzyme has shown Asn(376) and Asn(133) to be labile residues. Site-directed mutagenesis of nine Asn residues revealed that the deamidations of Asn(32) and Asn(376) are the main determinants for the functional and regulatory differences observed between the 2- and 24-h-induced wild-type (wt) enzyme. The Asn(32) --> Asp, Asn(376) --> Asp, and the double mutant forms expressed for 2 h at 28 degrees C revealed qualitatively similar regulatory properties as the highly deamidated 24-h expressed wt-hPAH. Moreover, deamidation of Asn(32) in the wt-hPAH (24 h expression at 28 degrees C) and the Asn(32) --> Asp mutation both increase the initial rate of phosphorylation of Ser(16) by cAMP-dependent protein kinase (p < 0.005). By contrast, the substitution of Gly(33) with Ala or Val, both preventing the deamidation of Asn(32), resulted in enzyme forms that were phosphorylated at a similar rate as nondeamidated wt-hPAH, even on 24-h expression. The other Asn --> Asp substitutions (in the catalytic domain) revealed that Asn(207) and Asn(223) have an important stabilizing structural function. Finally, two recently reported phenylketonuria mutations at Asn residues in the catalytic domain were studied, i.e. Asn(167) --> Ile and Asn(207) --> Asp, and their phenotypes were characterized.  相似文献   

6.
Role of the calcium-binding residues Asp231, Asp233, and Asp438 of Bacillus amyloliquefaciens α-amylase (BAA) on the enzyme properties was investigated by site-directed mutagenesis. The calcium-binding residues Asp231, Asp233, and Asp438 were replaced with Asn, Asn, and Gly to produce the mutants D231N, D233N, and D438G, respectively. The mutant amylases were purified to homogeneity and the purified enzymes was estimated to be approximately 58 kDa. The specific activity for the mutant enzyme D233N was decreased by 84.8%, while D231N and D438G showed a decrease of 6.3% and 3.5% to that of the wild-type enzyme, respectively. No significant changes in the K m value, thermo-stability, optimum temperature, and optimum pH were observed in the mutations of D231N and D438G, while substitution of Asp233 with Asn resulted in a dramatic reduction in the value of catalytic efficiency (K cat/K m) and thermo-stability at 60°C. The ranges of optimum temperature and optimum pH for D233N were also reduced to about 10°C and 3–4 units, respectively.  相似文献   

7.
Because mutations of the ionizable Asp at position 55 of the phosphatidylcholine preferring phospholipase C from Bacillus cereus (PLC(Bc)) to a non-ionizable Asn generate a mutant enzyme (D55N) with 10(4)-fold lower catalytic activity than the wild-type enzyme, we tentatively identified Asp55 as the general base for the enzymatic reaction. To eliminate the alternate possibility that Asp55 is a structurally important amino acid, the X-ray structures of unbound D55N and complexes of D55N with two non-hydrolyzable substrate analogues have been solved and refined to 2.0, 2.0, and 2.3A, respectively. The structures of unbound wild-type PLC(Bc) and a wild-type PLC(Bc)-complex with a non-hydrolyzable substrate analogue do not change significantly as a result of replacing Asp55 with Asn. These observations demonstrate that Asp55 is not critical for the structural integrity of the enzyme and support the hypothesis that Asp55 is the general base in the PLC(Bc)-catalyzed hydrolysis of phospholipids.  相似文献   

8.
The structure of the A24D/D61N substituted subunit c of Escherichia coli ATP synthase, in which the essential carboxylate has been switched from residue 61 of the second transmembrane helix (TMH) to residue 24 of the first TMH, has been determined by heteronuclear multidimensional NMR in a monophasic chloroform/methanol/water (4:4:1) solvent mixture. As in the case of the wild-type protein, A24D/D61N substituted subunit c forms a hairpin of two extended alpha-helices (residues 5-39 and 46-78), with residues 40-45 forming a connecting loop at the center of the protein. The structure was determined at pH 5, where Asp24 is fully protonated. The relative orientation of the two extended helices in the A24D/D61N structure is different from that in the protonated form of the wild-type protein, also determined at pH 5. The C-terminal helix is rotated by 150 degrees relative to the wild-type structure, and the N-terminal helix is rotated such that the essential Asp24 carboxyl group packs on the same side of the molecule as Asp61 in the wild-type protein. The changes in helix-helix orientation lead to a structure that is quite similar to that of the deprotonated form of wild-type subunit c, determined at pH 8. When a decameric ring of c subunits was modeled from the new structure, the Asp24 carboxyl group was found to pack in a cavity at the interface between two subunits that is similar to the cavity in which Asp61 of the wild-type protein is predicted to pack. The interacting faces of the packed subunits in this model are also similar to those in the wild-type model. The results provide further evidence that subunit c is likely to fold in at least two conformational states differing most notably in the orientation of the C-terminal helix. Based upon the structure, a mechanistic model is discussed that indicates how the wild-type and A24D/D61N subunits could utilize similar helical movements during H(+) transport-coupled rotation of the decameric c ring.  相似文献   

9.
The dynamics and the spectra of the excited state of the retinal in bacteriorhodopsin (bR) and its K-intermediate at pH 0 was compared with that of bR and halorhodopsin at pH 6.5. The quantum yield of photoisomerization in acid purple bR was estimated to be at least 0.5. The change of pH from 6.5 to 2 causes a shift of the absorption maximum from 568 to 600 nm (acid blue bR) and decreases the rate of photoisomerization. A further decrease in pH from 2 to 0 shifts the absorption maximum back to 575 nm when HCl is used (acid purple bR). We found that the rate of photoisomerization increases when the pH decreases from 2 to 0. The effect of chloride anions on the dynamics of the retinal photoisomerization of acid bR (pH 2 and 0) and some mutants (D85N, D212N, and R82Q) was also studied. The addition of 1 M HCl (to make acid purple bR, pH 0) or 1 M NaCl to acid blue bR (pH 2) was found to catalyze the rate of the retinal photoisomerization process. Similarly, the addition of 1 M NaCl to the solution of some bR mutants that have a reduced rate of retinal photoisomerization (D85N, D212N, and R82Q) was found to catalyze the rate of their retinal photoisomerization process up to the value observed in wild-type bR. These results are explained by proposing that the bound Cl- compensates for the loss of the negative charges of the COO- groups of Asp85 and/or Asp212 either by neutralization at low pH or by residue replacement in D85N and D212N mutants.  相似文献   

10.
Sequence-specific 2D methodology has been used to assign the 1H NMR signals for all active site residues in the paramagnetic cyano-met complexes of sperm whale synthetic double mutant His64[E7]-->Val/Thr67[E10]-->Arg (VR-met-MbCN) and triple mutant His64[E7]-->Val/Thr67[E10]-->Arg/Arg45[CD3]-->Asn (VRN-metMbCN). The resulting dipolar shifts for noncoordinated proximal side residues were used to quantitatively determine the orientation of the paramagnetic susceptibility tensor in the molecular framework for the two mutants, which were found indistinguishable but distinct from those of both wild-type and the His64[E7]-->Val single point mutant (V-metMbCN). The observed dipolar shifts for the E helix backbone protons and Phe43[CD1], together with steady-state nuclear Overhauser effect between the E helix and the heme, were analyzed to show that both the E helix and Phe43[CD1] move slightly closer to the iron to minimize the vacancy resulting from the His64[E7]-->Val substitution, as found in V-metMbCN (Rajarathnam, K., J. Qin, G.N. LaMar, M. L. Chiu, and S. G. Sligar. 1993. Biochemistry. 32:5670-5680). The dipolar shifts of the mutated Val64[E7] and Arg67[E10] allow the determination of their orientations relative to the heme, and the latter residue is shown to insert into the pocket and provide a hydrogen bond to the coordinated ligand, as found in the naturally occurring ValE7/ArgE10 genetic variant, Aplysia limacina Mb. The oxy-complex of both A. limacina Mb and VR-Mb, VRN-Mb have been proposed to be stabilized by this hydrogen bonding interaction (Travaglini Allocatelli, C. et al. 1993. Biochemistry. 32:6041-6049). The magnitude of the tilt of the major magnetic axes from the heme normal in VR-metMbCN and VRN-metMbCN, which is related to the tilt of the ligand, is the same as in wild-type or V-metMbCN, but the direction of tilt is altered from that in V-metMbCN. It is concluded that the change in the direction of the ligand tilt in both the double and triple mutants, as compared to WT metMbCN and V-metMbCN single mutant, is due to the attractive hydrogen-bonding between ArgE10 and the bound cyanide.  相似文献   

11.
The difference Fourier transform infrared spectrum for the N intermediate in the photoreaction of the light-adapted form of bacteriorhodopsin can be recorded at pH 10 at 274 K (Pfefferlé, J.-M., Maeda, A., Sasaki, J., and Yoshizawa, T. (1991) Biochemistry 30, 6548-6556). Under these conditions, Asp96-->Asn bacteriorhodopsin gives a photoproduct which shows changes in protein structure similar to those observed in N of wild-type bacteriorhodopsin. However, decreased intensity of the chromophore bands and the single absorbance maximum at about 400 nm indicate that the Schiff base is unprotonated, as in the M intermediate. This photoproduct was named MN. At pH 7, where the supply of proton is not as restricted as at pH 10, Asp96-->Asn bacteriorhodopsin yields N with a protonated Schiff base. The Asn96 residue, which cannot deprotonate as Asp96 in wild-type bacteriorhodopsin, is perturbed upon formation of both MN at pH 10 and N at pH 7. We suggest that the reprotonation of the Schiff base is preceded by a large change in the protein structure including perturbation of the residue at position 96.  相似文献   

12.
The gene coding for bacteriorhodopsin was modified in vitro to replace Asp212 with asparagine and expressed in Halobacterium halobium. X-ray diffraction measurements showed that the major lattice dimension of purple membrane containing the mutated bacteriorhodopsin was the same as wild type. At pH greater than 7, the Asp212----Asn chromophore was blue (absorption maximum at 585 nm) and exhibited a photocycle containing only the intermediates K and L, i.e. a reaction sequence very similar to that of wild-type bacteriorhodopsin at pH less than 3 and the blue form of the Asp85----Glu protein at pH less than 9. Since in the latter cases these effects are attributed to protonation of residue 85, it now appears that removal of the carboxylate of Asp212 has similar consequences as removing the carboxylate of Asp85. However, an important difference is that only Asp85 affects the pKa of the Schiff base. At pH less than 7, the Asp212----Asn protein was purple (absorption maximum at 569 nm) but photoexcitation produced only 15% of the normal amount of M and the transport activity was partial. The reactions of the blue and purple forms after photoexcitation are both quantitatively accounted for by a proposed scheme, K in equilibrium with L1 in equilibrium with L2----BR, but with the addition of an L1 in equilibrium with M reaction with unfavorable pKa for Schiff base deprotonation in the purple form. The latter hinders the transient accumulation of M, and the consequent branching at L1 allows only partial proton transport activity. The results are consistent with the existence of a complex counterion for the Schiff base proposed earlier (De Groot, H. J. M., Harbison, G. S., Herzfeld, J., and Griffin, R. G. (1989) Biochemistry 28, 3346-3353) and suggest that Asp85, Asp212, and at least one other protonable residue participate in it.  相似文献   

13.
Previous mutagenesis studies with bacteriorhodopsin have shown that reprotonation of the Schiff's base is the rate-limiting step in the photocycle of the D96N mutant, whereas retinal re-isomerization and return of the protein to the initial state constitute the rate-limiting events in the photocycle of the L93A mutant. Thus, in the D96N mutant, decay of the M intermediate is slowed down by more than 100-fold at pH 7. In the L93A mutant, decay of the O intermediate is slowed down by 250-fold. We report here that in the L93A, D96N double mutant, decay of the M intermediate, as well as the formation and decay of the O intermediate, are slowed down dramatically. The photocycle is completed by the decay of a long-lived O intermediate, as in the L93A mutant. The decay of the M and O intermediates in the double mutant parallels the behavior seen in the single mutants over a wide temperature and pH range, arguing that the observed independence is an intrinsic property of the mutant. The slow decay of the M and O intermediates can be selectively and independently reversed under conditions identical to those used for the corresponding intermediates in the D96N and L93A single mutants. Because the effects of the two individual mutations are preserved in the double mutant and can be independently reversed, we conclude that residues Asp 96 and Leu 93 act independently and at different stages of the bacteriorhodopsin photocycle. These results also show that formation of the O intermediate only requires protonation of the Schiff's base and is independent of the protonation of Asp 96 from the aqueous medium.  相似文献   

14.
The role of the extracellular Glu side chains of bacteriorhodopsin in the proton transport mechanism has been studied using the single mutants E9Q, E74Q, E194Q, and E204Q; the triple mutant E9Q/E194Q/E204Q; and the quadruple mutant E9Q/E74Q/E194Q/E204Q. Steady-state difference and deconvoluted Fourier transform infrared spectroscopy has been applied to analyze the M- and N-like intermediates in membrane films maintained at a controlled humidity, at 243 and 277 K at alkaline pH. The mutants E9Q and E74Q gave spectra similar to that of wild type, whereas E194Q, E9Q/E194Q/E204Q, and E9Q/E74Q/E194Q/E204Q showed at 277 K a N-like intermediate with a single negative peak at 1742 cm(-1), indicating that Asp(85) and Asp(96) are deprotonated. Under the same conditions E204Q showed a positive peak at 1762 cm(-1) and a negative peak at 1742 cm(-1), revealing the presence of protonated Asp(85) (in an M intermediate environment) and deprotonated Asp(96). These results indicate that in E194Q-containing mutants, the second increase in the Asp(85) pK(a) is inhibited because of lack of deprotonation of the proton release group. Our data suggest that Glu(194) is the group that controls the pK(a) of Asp(85).  相似文献   

15.
Asn112 is located at the active site of thermolysin, 5-8 A from the catalytic Zn2+ and catalytic residues Glu143 and His231. When Asn112 was replaced with Ala, Asp, Glu, Lys, His, and Arg by site-directed mutagenesis, the mutant enzymes N112D and N112E, in which Asn112 is replaced with Asp and Glu, respectively, were secreted as an active form into Escherichia coli culture medium, while the other four were not. In the hydrolysis of a neutral substrate N-[3-(2-furyl)acryloyl]-Gly-L-Leu amide, the kcat/Km values of N112D and N112E exhibited bell-shaped pH-dependence, as did the wild-type thermolysin (WT). The acidic pKa of N112D was 5.7 +/- 0.1, higher by 0.4 +/- 0.2 units than that of WT, suggesting that the introduced negative charge suppressed the protonation of Glu143 or Zn2+-OH. In the hydrolysis of a negatively charged substrate, N-carbobenzoxy-l-Asp-l-Phe methyl ester (ZDFM), the pH-dependence of kcat/Km of the mutants decreased with increase in pH from 5.5 to 8.5, while that of WT was bell-shaped. This difference might be explained by the electrostatic repulsion between the introduced Asp/Glu and ZDFM, suggesting that introducing ionizing residues into the active site of thermolysin might be an effective means of modifying its pH-activity profile.  相似文献   

16.
The two most buried carboxyl groups in ribonuclease Sa (RNase Sa) are Asp33 (99% buried; pK 2.4) and Asp79 (85% buried; pK 7.4). Above these pK values, the stability of the D33A variant is 6kcal/mol less than wild-type RNase Sa, and the stability of the D79A variant is 3.3kcal/mol greater than wild-type RNase Sa. The key structural difference between the carboxyl groups is that Asp33 forms three intramolecular hydrogen bonds, and Asp79 forms no intramolecular hydrogen bond. Here, we focus on Asp79 and describe studies of 11 Asp79 variants. Most of the variants were at least 2kcal/mol more stable than wild-type RNase Sa, and the most interesting was D79F. At pH 3, below the pK of Asp79, RNase Sa is 0.3kcal/mol more stable than the D79F variant. At pH 8.5, above the pK of Asp79, RNase Sa is 3.7kcal/mol less stable than the D79F variant. The unfavorable contribution of Asp79 to the stability appears to result from the Born self-energy of burying the charge and, more importantly, from unfavorable charge-charge interactions. To counteract the effect of the negative charge on Asp79, we prepared the Q94K variant and the crystal structure showed that the amino group of the Lys formed a hydrogen-bonded ion pair (distance, 2.71A; angle, 100 degrees ) with the carboxyl group of Asp79. The stability of the Q94K variant was about the same as the wild-type at pH 3, where Asp79 is uncharged, but 1kcal/mol greater than that of wild-type RNase Sa at pH 8.5, where Asp79 is charged. Differences in hydrophobicity, steric strain, Born self-energy, and electrostatic interactions all appear to contribute to the range of stabilities observed in the variants. When it is possible, replacing buried, non-hydrogen bonded, ionizable side-chains with non-polar side-chains is an excellent means of increasing protein stability.  相似文献   

17.
We have introduced amino acid substitutions into two regions of the extracellular domain of the vesicular stomatitis virus (VSV) glycoprotein (G protein) and examined the effect of these mutations on protein transport, low-pH-induced stability of G protein oligomers, and membrane fusion activity. We suggested previously that the region between amino acids 118 and 139 may be important for the membrane fusion activity of G protein, on the basis of the characterization of a fusion-defective G protein mutant (M. A. Whitt, P. Zagouras, B. Crise, and J. K. Rose, J. Virol. 64:4907-4913, 1990). It has also been postulated by others that this region as well as the region between amino acids 181 and 212 may constitute putative internal fusion domains of VSV G protein. In this report, we show that three different amino acids substitutions between residues 118 and 139 (G-124-->E, P-127-->D, and A-133-->K) either altered or abolished low-pH-dependent membrane fusion activity. In contrast, substitutions between residues 192 and 212 resulted either in G proteins that had wild-type fusion activity or in mutant proteins in which the mutation prevented transport of G protein to the cell surface. Two of the substitutions between residues 118 and 139 (G-124-->E and P-127-->D) resulted in G proteins that were fusion defective at pH 5.7, although syncytia were observed after cells were treated with fusion buffer at pH 5.5, albeit at levels significantly less than that induced by wild-type G protein. Interestingly, when either G-124-->E or P-127-->D was incorporated into tsO45 virions, the resulting particles were not infectious, presumably because the viral envelope was not able to fuse with the proper intracellular membrane. These results support the hypothesis that the region between amino acids 118 and 139 is important for the membrane fusion activity of VSV G protein and may constitute an internal fusion domain.  相似文献   

18.
Lin FP  Chen HC  Lin CS 《IUBMB life》1999,48(2):199-204
Site-directed mutagenesis was used to explore the roles of amino acid residues involved in the activity of chitinase from Aeromonas caviae. Kinetic parameters for 4-methylumbelliferyl-N,N'-diacetyl-chitobiose or 4-methylumbelliferyl-N,N',N"-triacetylchitotriose hydrolysis were determined with wild-type and mutant chitinases. Chitinases with the mutations E315D (or Q) and D391E (or N) were severely impaired and had dramatically decreased kcat. However, the effect of the these mutations on the Km values were different. The function of the carboxyl group of Asp313 was partially replaced by the amide of Asn when the 4-methylumbelliferyl-N,N',N"-triacetylchitotriose substrate was used. Results indicated that Asp313, Glu315, and Asp391 might be the best candidates for the catalytic residues of chitinase A from Aeromonas caviae.  相似文献   

19.
Arg(82) is one of the four buried charged residues in the retinal binding pocket of bacteriorhodopsin (bR). Previous studies show that Arg(82) controls the pK(a)s of Asp(85) and the proton release group and is essential for fast light-induced proton release. To further investigate the role of Arg(82) in light-induced proton pumping, we replaced Arg(82) with histidine and studied the resulting pigment and its photochemical properties. The main pK(a) of the purple-to-blue transition (pK(a) of Asp(85)) is unusually low in R82H: 1.0 versus 2.6 in wild type (WT). At pH 3, the pigment is purple and shows light and dark adaptation, but almost no light-induced Schiff base deprotonation (formation of the M intermediate) is observed. As the pH is increased from 3 to 7 the M yield increases with pK(a) 4.5 to a value approximately 40% of that in the WT. A transition with a similar pK(a) is observed in the pH dependence of the rate constant of dark adaptation, k(da). These data can be explained, assuming that some group deprotonates with pK(a) 4.5, causing an increase in the pK(a) of Asp(85) and thus affecting k(da) and the yield of M. As the pH is increased from 7 to 10.5 there is a further 2.5-fold increase in the yield of M and a decrease in its rise time from 200 &mgr;s to 75 &mgr;s with pK(a) 9. 4. The chromophore absorption band undergoes a 4-nm red shift with a similar pK(a). We assume that at high pH, the proton release group deprotonates in the unphotolyzed pigment, causing a transformation of the pigment into a red-shifted "alkaline" form which has a faster rate of light-induced Schiff base deprotonation. The pH dependence of proton release shows that coupling between Asp(85) and the proton release group is weakened in R82H. The pK(a) of the proton release group in M is 7.2 (versus 5.8 in the WT). At pH < 7, most of the proton release occurs during O --> bR transition with tau approximately 45 ms. This transition is slowed in R82H, indicating that Arg(82) is important for the proton transfer from Asp(85) to the proton release group. A model describing the interaction of Asp(85) with two ionizable residues is proposed to describe the pH dependence of light-induced Schiff base deprotonation and proton release.  相似文献   

20.
A Giletto  C N Pace 《Biochemistry》1999,38(40):13379-13384
The side-chain carboxyl of Asp 76 in ribonuclease T1 (RNase T1) is buried, charged, non-ion-paired, and forms three good intramolecular hydrogen bonds (2.63, 2.69, and 2.89 A) and a 2.66 A hydrogen bond to a buried, conserved water molecule. When Asp 76 was replaced by Asn, Ser, and Ala, the conformational stability of the protein decreased by 3.1, 3.2, and 3.7 kcal/mol, respectively. The stability was measured as a function of pH for wild-type RNase T1 and the D76N mutant and showed that the pH dependence below pH 3 was almost entirely due to Asp 76. The pK of Asp 76 is 0.5 in the native state and 3.7 in the denatured state. Thus, the hydrogen bonding of the carboxyl group of Asp 76 contributes more than half of the net stability of RNase T1 at pH 7. In addition, the charged carboxyl of Asp 76 stabilizes structure in the denatured states of RNase T1 that is not present in D76N, D76S, and D76A.  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