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.
To understand the ligand binding properties of the human GnRH receptor (hGnRH-R), 24 site-specific mutants within transmembrane helices (TMH) 1, 2, and 5 and the extracellular loop 2 (E2) were generated. These mutants were analyzed by using a functional reporter gene assay, monitoring receptor signaling via adenylate cyclase to a cAMP-responsive element fused to Photinus pyralis luciferase. The functional behavior of 14 receptor mutants, capable of G-protein coupling and signaling, was studied in detail with different well described agonistic and antagonistic peptide ligands. Furthermore, the binding constants were determined in displacement binding experiments with the antagonist [125I]Cetrorelix. The substitution of residues K36, Q204, W205, H207, Q208, F20, F213, F216, and S217 for alanine had no or only a marginal effect on ligand binding and signaling. In contrast, substitution of N87, Eg9, D9, R179, W206, Y211, F214, and T215 for alanine resulted in receptor proteins neither capable of ligand binding nor signal transduction. Within those mutants affecting ligand binding and signaling to various degrees, W101A, N102A, and N212Q differentiate between agonists and antagonists. Thus, in addition to N102 already described, the residues W101 in TMH2 and N212 in TMH5 are important for the architecture of the ligand-binding pocket. Based on the experimental data, three-dimensional models for binding of the superagonist D-Trp6-GnRH (Triptorelin) and the antagonist Cetrorelix to the hGnRH-R are proposed. Both decapeptidic ligands are bound to the receptor in a bent conformation with distinct interactions within the binding pocket formed by all TMHs, E2, and E3. The antagonist Cetrorelix with bulky hydrophobic N-terminal amino acids interacts with quite different receptor residues, a hint at the failure to induce an active, G protein-coupling receptor conformation.  相似文献   

2.
Human P-glycoprotein (P-gp) transports a wide variety of structurally diverse compounds out of the cell. Knowledge about the packing of the transmembrane (TM) segments is essential for understanding the mechanism of drug recognition and transport. We used cysteine-scanning mutagenesis and disulfide cross-linking analysis to determine which TM segment in the COOH half of P-gp was close to TMs 5 and 6 since these segments in the NH(2) half are important for drug binding. An active Cys-less P-gp mutant cDNA was used to generate 240 double cysteine mutants that contained 1 cysteine in TMs 5 or 6 and another in TMs 7 or 8. The mutants were subjected to oxidative cross-linking analysis. No disulfide cross-linking was observed in the 140 TM6/TM7 or TM6/TM8 mutants. By contrast, cross-linking was detected in several P-gp TM5/TM8 mutants. At 4 degrees C, when thermal motion is low, P-gp mutants N296C(TM5)/G774C(TM8), I299C(TM5)/F770C(TM8), I299C(TM5)/G774C(TM8), and G300C(TM5)/F770C(TM8) showed extensive cross-linking with oxidant. These mutants retained drug-stimulated ATPase activity, but their activities were inhibited after treatment with oxidant. Similarly, disulfide cross-linking was inhibited by vanadate trapping of nucleotide. These results indicate that significant conformational changes must occur between TMs 5 and 8 during ATP hydrolysis. We revised the rotational symmetry model for TM packing based on our results and by comparison to the crystal structure of MsbA (Chang, G. (2003) J. Mol. Biol. 330, 419-430) such that TM5 is adjacent to TM8, TM2 is adjacent to TM11, and TMs 1 and 7 are next to TMs 6 and 12, respectively.  相似文献   

3.
Subunit a plays a key role in promoting H+ transport and the coupled rotary motion of the subunit c ring in F1F0-ATP synthase. H+ binding and release occur at Asp-61 in the middle of the second transmembrane helix (TMH) of F0 subunit c. H+ are thought to reach Asp-61 via aqueous pathways mapping to the surfaces of TMHs 2-5 of subunit a. TMH4 of subunit a is thought to pack close to TMH2 of subunit c based upon disulfide cross-link formation between Cys substitutions in both TMHs. Here we substituted Cys into the fifth TMH of subunit a and the second TMH of subunit c and tested for cross-linking using bis-methanethiosulfonate (bis-MTS) reagents. A total of 62 Cys pairs were tested and 12 positive cross-links were identified with variable alkyl length linkers. Cross-linking was achieved near the middle of the bilayer for the Cys pairs a248C/c62C, a248C/ c63C, a248C/c65C, a251C/c57C, a251C/c59C, a251C/c62C, a252C/c62C, and a252C/c65C. Cross-linking was achieved near the cytoplasmic side of the bilayer for Cys pairs a262C/c53C, a262C/c54C, a262C/c55C, and a263C/c54C. We conclude that both aTMH4 and aTMH5 pack proximately to cTMH2 of the c-ring. In other experiments we demonstrate that aTMH4 and aTMH5 can be simultaneously cross-linked to different subunit c monomers in the c-ring. Five mutants showed pH-dependent cross-linking consistent with aTMH5 changing conformation at lower pH values to facilitate cross-linking. We suggest that the pH-dependent conformational change may be related to the proposed role of aTMH5 in gating H+ access from the periplasm to the cAsp-61 residue in cTMH2.  相似文献   

4.
The structural changes involved in ligand-dependent activation of G protein-coupled receptors are not well understood at present. To address this issue, we developed an in situ disulfide cross-linking strategy using the rat M(3) muscarinic receptor, a prototypical G(q)-coupled receptor, as a model system. It is known that a tyrosine residue (Tyr(254)) located at the C terminus of transmembrane domain (TM) V and several primarily hydrophobic amino acids present within the cytoplasmic portion of TM VI play key roles in determining the G protein coupling selectivity of the M(3) receptor subtype. To examine whether M3 receptor activation involves changes in the relative orientations of these functionally critical residues, pairs of cysteine residues were substituted into a modified version of the M(3) receptor that contained a factor Xa cleavage site within the third intracellular loop and lacked most endogenous cysteine residues. All analyzed mutant receptors contained a Y254C point mutation and a second cysteine substitution within the segment Lys(484)-Ser(493) at the intracellular end of TM VI. Following their transient expression in COS-7 cells, mutant receptors present in their native membrane environment (in situ) were subjected to mild oxidizing conditions, either in the absence or in the presence of the muscarinic agonist, carbachol. The successful formation of disulfide cross-links was monitored by studying changes in the electrophoretic mobility of oxidized, factor Xa-treated receptors on SDS gels. The observed cross-linking patterns indicated that M(3) receptor activation leads to structural changes that allow the cytoplasmic ends of TM V and TM VI to move closer to each other and that also appear to involve a major change in secondary structure at the cytoplasmic end of TM VI. This is the first study employing an in situ disulfide cross-linking strategy to examine agonist-dependent dynamic structural changes in a G protein-coupled receptor.  相似文献   

5.
To probe proximities between amino acids in the cytoplasmic domain by using mutants containing engineered cysteine pairs, three sets of rhodopsin mutants have been prepared. In the first two sets, a cysteine was placed, one at a time, at positions 311-314 in helix VIII, while the second cysteine was fixed at position 246 (set I) and at position 250 (set II) at the cytoplasmic end of helix VI. In the third set, one cysteine was fixed at position 65 while the second cysteine was varied between amino acid positions 306 and 321 located at the cytoplasmic end of helix VII and throughout in helix VIII. Rapid disulfide bond formation in the dark was found between the cysteine pairs in mutants A246C/Q312C,A246C/K311C and in mutants H65C/C316, H65C/315C and H65C/312C. Disulfide bond formation at much lower rates was found in mutants A246C/F313C, V250C/Q312C, H65C/N310C, H65C/K311C, H65C/F313C, and H65C/R314C; the remaining mutants showed no significant disulfide bond formation. Comparisons of the results from disulfide bond formation in solution with the distances observed in the rhodopsin crystal structure showed that the rates of disulfide bond formation in most cases were consistent with the amino acid proximities as revealed in crystal structure. However, deviations were also found, in particular, in the set containing fixed cysteine at position Cys246 and cysteines at positions 311-314. The results implicate significant effects of structural dynamics on disulfide bond formation in solution.  相似文献   

6.
Multiple computational methods have been employed in a comparative study of thyrotropin-releasing hormone receptors 1 and 2 (TRH-R1 and TRH-R2) to explore the structural bases for the different functional properties of these G protein-coupled receptors. Three-dimensional models of both murine TRH receptors have been built and optimized by means of homology modeling based on the crystal structure of bovine rhodopsin, molecular dynamics simulations, and energy minimizations in a membrane-aqueous environment. The comparison between the two models showed a correlation between the higher flexibility and higher basal activity of TRH-R2 versus the lesser flexibility and lower basal activity of TRH-R1 and supported the involvement of the highly conserved W6.48 in the signaling process. A correlation between the level of basal activity and conformational changes of TM5 was detected also. Comparison between models of the wild type receptors and their W6.48A mutants, which have reversed basal activities compared with their respective wild types, further supported these correlations. A flexible molecular docking procedure revealed that TRH establishes a direct interaction with W6.48 in TRH-R2 but not in TRH-R1. We designed and performed new mutagenesis experiments that strongly supported these observations.  相似文献   

7.
Putative transmembrane helices (TM) 1 and 11 in the metal-tetracycline/H(+) antiporter are predicted to be close to each other on the basis of disulfide cross-linking experiments of the double-cysteine mutants in the periplasmic loop regions (Kubo, Y., Konishi, S., Kawabe, T., Nada, S., and Yamaguchi, A. (2000) J. Biol. Chem. 275, 5270-5274). In this study, each amino acid from Asn-2 to Gly-44 in the putative TM1 and loop1-2 regions or that from Ser-328 to Gly-366 in TM11 and its flanking regions was individually replaced with cysteine. With respect to the TM1 region, 10 mutants, from T5C to L14C, were all not reactive with N-ethylmaleimide (NEM), and from D15C to I22C, NEM-reactive and non-reactive mutations periodically appeared every two residues. Three mutants, M23C to V25C, were all NEM-reactive, but the degree of the latter two mutants was very low. Seven mutants, from L26C to E32C, were all highly reactive with NEM. Therefore, the region of TM1 is composed of the 21 amino acid residues from Thr-5 to Val-25. It is a partially amphiphilic helix, that is, the N-terminal (cytoplasmic) half is embedded in the hydrophobic interior, and the C-terminal (periplasmic) half faces a water-filled channel. With respect to TM11, nine mutants, from S328C to G336C, and six mutants, from L361C to G366C, were all reactive with NEM. On the other hand, out of the 24 mutants, from L337C to S360C, 17 were not reactive with NEM, and the 7 NEM-reactive mutants were scattered, indicating that this region is a transmembrane segment. The 7 residues from Val-347 to Phe-353 including Pro-350 formed a central hydrophobic core, and the 7 NEM-reactive mutations were periodically distributed in its flanking regions, indicating that both ends of TM11 face a water-filled channel. Ala-354 is located at about 1/3 of the length from the periplasmic end of TM11. Disulfide cross-linking experiments on double-cysteine mutants having the combination of A354C and a cysteine-scanning mutation in the loop1-2 region indicated that loop1-2 is very flexible and close to the periplasmic end of TM11. Tetracycline prevented the cross-linking formation between the periplasmic ends of TM1 and TM11; however, it did not affect the cross-linking between loop1-2 and TM11, indicating that the substrate-induced conformational change involves a shift in the relative locations of TM1 and TM11.  相似文献   

8.
Determinants of glycosylation site usage were explored by using the hemagglutinin-neuraminidase (HN) glycoprotein of the paramyxovirus Newcastle disease virus. The amino acid sequence of the HN protein, a type II glycoprotein, has six N-linked glycosylation addition sites, G1 to G6, two of which, G5 and G6, are not used for the addition of carbohydrate (L. McGinnes and T. Morrison, Virology 212:398-410, 1995). The sequence of this protein also has 13 cysteine residues in the ectodomain (C2 to C14). Mutation of either cysteine 13 or cysteine 14 resulted in the addition of another oligosaccharide chain to the protein. These cysteine residues flank the normally unused G6 glycosylation addition site, and mutation of the G6 site eliminated the extra glycosylation found in the cysteine mutants. These results suggested that failure to form an intramolecular disulfide bond resulted in the usage of a normally unused glycosylation site. This conclusion was confirmed by preventing cotranslational disulfide bond formation in cells by using dithiothreitol. Under these conditions, the wild-type protein acquired extra glycosylation, which was eliminated by mutation of the G6 site. These results suggest that localized folding events on the nascent chain, such as disulfide bond formation, which block access to the oligosaccharyl transferase are a determinant of glycosylation site usage.  相似文献   

9.
In a previous report we showed that TRH-induced down-regulation of the density of its receptors (TRH-Rs) on rat pituitary tumor (GH3) cells was preceded by a decrease in the activity of the mRNA for the TRH-R, as assayed in Xenopus oocytes. Here we report the effects of TRH, elevation of cytoplasmic free Ca2+ concentration, phorbol myristate acetate (PMA), and H-7 [1-(5-isoquinolinesulfonyl)2-methylpiperazine dihydrochloride], an inhibitor of protein kinases, on the levels of TRH-R mRNA, which were measured by Northern analysis and in nuclease protection assays using probes made from mouse pituitary TRH-R cDNA, in GH3 cells. These agents were studied to gain insight into the mechanism of the TRH effect, because signal transduction by TRH involves generation of inositol 1,4,5-trisphosphate and elevation of cytoplasmic free Ca2+ concentration, which leads to activation of Ca2+/calmodulin-dependent protein kinase, and of 1,2-diacylglycerol, which leads to activation of protein kinase-C. TRH (1 microM TRH, a maximally effective dose) caused a marked transient decrease in TRH-R mRNA that attained a nadir of 20-45% of control by 3-6 h, increased after 9 h, but was still below control levels after 24 h. Elevation of the cytoplasmic free Ca2+ concentration had no effect on TRH-R mRNA. A maximally effective dose of PMA (1 microM) caused decreases in TRH-R mRNA that were similar in magnitude and time course to those induced by 1 microM TRH. H-7 (20 microM) blocked the effects of TRH and PMA to lower TRH-R mRNA to similar extents.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Multiple transmembrane (TM) segments line the pore of the cystic fibrosis transmembrane conductance regulator Cl(-) channel; however, the relative alignment of these TMs and their relative movements during channel gating are unknown. To gain three-dimensional structural information on the outer pore, we have used patch clamp recording to study the proximity of pairs of cysteine side chains introduced into TMs 6 and 11, using both disulfide cross-linking and Cd(2+) coordination. Following channel activation, disulfide bonds could apparently be formed between three cysteine pairs (of 15 studied): R334C/T1122C, R334C/G1127C, and T338C/S1118C. To examine the state dependence of cross-linking, we combined these cysteine mutations with a nucleotide-binding domain mutation (E1371Q) that stabilizes the channel open state. Investigation of the effects of the E1371Q mutation on disulfide bond formation and Cd(2+) coordination suggests that although R334C/T1122C and T338C/S1118C are closer together in the channel open state, R334C/G1127C are close together and can form disulfide bonds only when the channel is closed. These results provide important new information on the three-dimensional structure of the outer mouth of the cystic fibrosis transmembrane conductance regulator channel pore: TMs 6 and 11 are close enough together to form disulfide bonds in both open and closed channels. Moreover, the altered relative locations of residues in open and in closed channels that we infer allow us to propose that channel opening and closing may be associated with a relative translational movement of TMs 6 and 11, with TM6 moving "down" (toward the cytoplasm) during channel opening.  相似文献   

11.
In rat pituitary GH3 cells, thyrotropin-releasing hormone (TRH) down-regulates TRH receptor (TRH-R) mRNA (Fujimoto, J., Straub, R.E., and Gershengorn, M.C. (1991) Mol. Endocrinol. 5, 1527-1532), at least in part, by stimulating its degradation (Fujimoto, J., Narayanan, C.S., Benjamin, J.E., Heinflink, M., and Gershengorn, M.C. (1992) Endocrinology 130, 1879-1884). Here we show that TRH regulates RNase activity in GH3 cells and that specific mRNA sequences are needed for in vivo regulation of TRH-R mRNA by TRH. TRH affected RNase activity in a biphasic manner with rapid stimulation (by 10 min) followed by a decrease to a rate slower than in control lysates within 6 h. This time course paralleled the effects of TRH on degradation of TRH-R mRNA in vivo. The regulated RNase activity was in a polysome-free fraction of the lysates and was not specific for TRH-R RNA. A truncated form of TRH-R RNA that was missing the entire 3'-untranslated region (TRHR-R5) was more stable than full-length TRH-R RNA (TRHR-WT). In contrast to TRHR-WT mRNA, TRHR-R5 mRNA and TRHR-D9 mRNA, which was missing the 143 nucleotides 5' of the poly(A) tail, were not down-regulated by TRH in stably transfected GH3 cells as their rates of degradation were not increased. These data show that TRH regulates RNase activity in GH3 cells, that the 3'-untranslated region bestows decreased stability on TRH-R mRNA and that the 3' end of the mRNA is necessary for regulation by TRH of TRH-R mRNA degradation. We present an hypothesis that explains specific regulation of TRH-R mRNA degradation by TRH in GH3 pituitary cells.  相似文献   

12.
It has been postulated that the hydrophobic loop of actin (residues 262-274) swings out and inserts into the opposite strand in the filament, stabilizing the filament structure. Here, we analyzed the hydrophobic loop dynamics utilizing four mutants that have cysteine residues introduced at a single location along the yeast actin loop. Lateral, copper-catalyzed disulfide cross-linking of the mutant cysteine residues to the native C374 in the neighboring strand within the filament was fastest for S265C, followed by V266C, L267C, and then L269C. Site-directed spin labeling (SDSL) studies revealed that C265 lies closest to C374 within the filament, followed by C266, C267, and then C269. These results are not predicted by the Holmes extended loop model of F-actin. Furthermore, we find that disulfide cross-linking destroys L267C and L269C filaments; only small filaments are observed via electron microscopy. Conversely, phalloidin protects the L267C and L269C filaments and inhibits their disulfide cross-linking. Combined, our data indicate that, in solution, the loop resides predominantly in a "parked" position within the filament but is able to dynamically populate other conformational states which stabilize or destabilize the filament. Such states may be exploited within a cell by filament-stabilizing and -destabilizing factors.  相似文献   

13.
The third intracellular loop (IL3) of G protein-coupled receptors (GPCRs) is an important contact domain between GPCRs and their G proteins. Previously, the IL3 of Ste2p, a Saccharomyces cerevisiae GPCR, was suggested to undergo a conformational change upon activation as detected by differential protease susceptibility in the presence and absence of ligand. In this study using disulfide cross-linking experiments we show that the Ste2p cytoplasmic ends of helix 5 (TM5) and helix 6 (TM6) that flank the amino and carboxyl sides of IL3 undergo conformational changes upon ligand binding, whereas the center of the IL3 loop does not. Single Cys substitution of residues in the middle of IL3 led to receptors that formed high levels of cross-linked Ste2p, whereas Cys substitution at the interface of IL3 and the contiguous cytoplasmic ends of TM5 and TM6 resulted in minimal disulfide-mediated cross-linked receptor. The alternating pattern of residues involved in cross-linking suggested the presence of a 3(10) helix in the middle of IL3. Agonist (WHWLQLKPGQPNleY) induced Ste2p activation reduced cross-linking mediated by Cys substitutions at the cytoplasmic ends of TM5 and TM6 but not by residues in the middle of IL3. Thus, the cytoplasmic ends of TM5 and TM6 undergo conformational change upon ligand binding. An α-factor antagonist (des-Trp, des-His-α-factor) did not influence disulfide-mediated Ste2p cross-linking, suggesting that the interaction of the N-terminus of α-factor with Ste2p is critical for inducing conformational changes at TM5 and TM6. We propose that the changes in conformation revealed for residues at the ends of TM5 and TM6 are affected by the presence of G protein but not G protein activation. This study provides new information about role of specific residues of a GPCR in signal transduction and how peptide ligand binding activates the receptor.  相似文献   

14.
Drebrin is a mammalian neuronal protein that binds to and organizes filamentous actin (F-actin) in dendritic spines, the receptive regions of most excitatory synapses that play a crucial role in higher brain functions. Here, the structural effects of drebrin on F-actin were examined in solution. Depolymerization and differential scanning calorimetry assays show that F-actin is stabilized by the binding of drebrin. Drebrin inhibits depolymerization mainly at the barbed end of F-actin. Full-length drebrin and its C-terminal truncated constructs were used to clarify the domain requirements for these effects. The actin binding domain of drebrin decreases the intrastrand disulfide cross-linking of Cys-41 (in the DNase I binding loop) to Cys-374 (C-terminal) but increases the interstrand disulfide cross-linking of Cys-265 (hydrophobic loop) to Cys-374 in the yeast mutants Q41C and S265C, respectively. We also demonstrate, using solution biochemistry methods and EM, the rescue of filament formation by drebrin in different cases of longitudinal interprotomer contact perturbation: the T203C/C374S yeast actin mutant and grimelysin-cleaved skeletal actin (between Gly-42 and Val-43). Additionally, we show that drebrin rescues the polymerization of V266G/L267G, a hydrophobic loop yeast actin mutant with an impaired lateral interface formation between the two filament strands. Overall, our data suggest that drebrin stabilizes actin filaments through its effect on their interstrand and intrastrand contacts.  相似文献   

15.
The conformational dynamics of filamentous actin (F-actin) is essential for the regulation and functions of cellular actin networks. The main contribution to F-actin dynamics and its multiple conformational states arises from the mobility and flexibility of the DNase I binding loop (D-loop; residues 40-50) on subdomain 2. Therefore, we explored the structural constraints on D-loop plasticity at the F-actin interprotomer space by probing its dynamic interactions with the hydrophobic loop (H-loop), the C-terminus, and the W-loop via mutational disulfide cross-linking. To this end, residues of the D-loop were mutated to cysteines on yeast actin with a C374A background. These mutants showed no major changes in their polymerization and nucleotide exchange properties compared to wild-type actin. Copper-catalyzed disulfide cross-linking was investigated in equimolar copolymers of cysteine mutants from the D-loop with either wild-type (C374) actin or mutant S265C/C374A (on the H-loop) or mutant F169C/C374A (on the W-loop). Remarkably, all tested residues of the D-loop could be cross-linked to residues 374, 265, and 169 by disulfide bonds, demonstrating the plasticity of the interprotomer region. However, each cross-link resulted in different effects on the filament structure, as detected by electron microscopy and light-scattering measurements. Disulfide cross-linking in the longitudinal orientation produced mostly no visible changes in filament morphology, whereas the cross-linking of D-loop residues > 45 to the H-loop, in the lateral direction, resulted in filament disruption and the presence of amorphous aggregates on electron microscopy images. A similar aggregation was also observed upon cross-linking the residues of the D-loop (> 41) to residue 169. The effects of disulfide cross-links on F-actin stability were only partially accounted for by the simulations of current F-actin models. Thus, our results present evidence for the high level of conformational plasticity in the interprotomer space and document the link between D-loop interactions and F-actin stability.  相似文献   

16.
Subunit a of F(1)F(0) ATP synthase is required in the H(+) transport driven rotation of the c-ring of F(0), the rotation of which is coupled to ATP synthesis in F(1). The three-dimensional structure of subunit a is unknown. In this study, Cys substitutions were introduced into two different transmembrane helices (TMHs) of subunit a, and the proximity of the thiol side chains was tested via attempted oxidative cross-linking to form the disulfide bond. Pairs of Cys substitutions were made in TMHs 2/3, 2/4, 2/5, 3/4, 3/5, and 4/5. Cu(+2)-catalyzed oxidation led to cross-link formation between Cys pairs L120C(TMH2) and S144C(TMH3), L120C(TMH2) and G218C(TMH4), L120C(TMH2) and H245C(TMH5), L120C(TMH2) and I246C(TMH5), N148C(TMH3) and E219C(TMH4), N148C(TMH3) and H245C(TMH5), and G218C(TMH4) and I248C(TMH5). Iodine, but not Cu(+2), was found to catalyze cross-link formation between D119C(TMH2) and G218C(TMH4). The results suggest that TMHs 2, 3, 4, and 5 form a four-helix bundle with one set of key functional residues in TMH4 (Ser-206, Arg-210, and Asn-214) located at the periphery facing subunit c. Other key residues in TMHs 2, 4, and 5, which were concluded previously to compose a possible aqueous access pathway from the periplasm, were found to locate to the inside of the four-helix bundle.  相似文献   

17.
P-glycoprotein (Pgp), a member of the ABC transporter family, functions as an ATP hydrolysis-driven efflux pump to rid the cell of toxic organic compounds, including a variety of drugs used in anti-cancer chemotherapy. We have recently obtained EM projection images of lipid-bound Pgp without nucleotide and transport substrate that showed the two halves of the transporter separated by a central cavity (Lee, J. Y., Urbatsch, I. L., Senior, A. E., and Wilkens, S. (2002) J. Biol. Chem. 277, 40125-40131). Addition of nucleotide and/or substrate lead to a close association of the two halves of the transporter, thereby closing the central cavity (Lee, J. Y., Urbatsch, I. L., Senior, A. E., and Wilkens, S. (2008) J. Biol. Chem. 283, 5769-5779). Here, we used cysteine-mediated disulfide cross-linking to further delineate the structural rearrangements of the two nucleotide binding domains (NBD1 and NBD2) that take place during catalysis. Cysteines introduced at or near the C-terminal ends of NBD1 and NBD2 allowed for spontaneous disulfide cross-linking under nonreducing conditions. For mutant A627C/S1276C, disulfide formation was with high efficiency and cross-linked Pgp retained 30-68% drug-stimulated ATPase activity compared with reduced or cysteine-less Pgp. Two other cysteine pairs (K615C/S1276C and A627C/K1260C) also formed a disulfide but to a lesser extent, and the cross-linked form of these two mutants had lower drug-stimulated ATPase activity. The data suggest that the C-terminal ends of the two NBDs of Pgp are not required to undergo significant motion with respect to one another during the catalytic cycle.  相似文献   

18.
Residues Tyr-110 through Gly-115 of serotonin transporter were replaced, one at a time, with cysteine. Of these mutants, only G113C retained full activity for transport, Q111C and N112C retained partial activity, but Y110C, G114C and G115C were inactive. Poor surface expression was at least partly responsible for the lack of transport by G114C and G115C. In membrane preparations, Y110C through G113C all bound a high affinity cocaine analog similarly to the wild type. Treatment with methanethiosulfonate reagents increased the transport activity of Q111C and N112C to essentially wild-type levels but had no measurable effect on the other mutants. The decreased activity of Q111C and N112C resulted from an increase in the KM for serotonin that was not accompanied by a decrease in serotonin binding affinity. Superfusion experiments indicated a defect in 5-HT exchange. Modification of the inserted cysteine residues reversed the increase in KM and the poor exchange, also with no effect on serotonin affinity. The results suggest that Gln-111 and Asn-112 are not required for substrate binding but participate in subsequent steps in the transport cycle.  相似文献   

19.
Residues Tyr-110 through Gly-115 of serotonin transporter were replaced, one at a time, with cysteine. Of these mutants, only G113C retained full activity for transport, Q111C and N112C retained partial activity, but Y110C, G114C and G115C were inactive. Poor surface expression was at least partly responsible for the lack of transport by G114C and G115C. In membrane preparations, Y110C through G113C all bound a high affinity cocaine analog similarly to the wild type. Treatment with methanethiosulfonate reagents increased the transport activity of Q111C and N112C to essentially wild-type levels but had no measurable effect on the other mutants. The decreased activity of Q111C and N112C resulted from an increase in the K(M) for serotonin that was not accompanied by a decrease in serotonin binding affinity. Superfusion experiments indicated a defect in 5-HT exchange. Modification of the inserted cysteine residues reversed the increase in K(M) and the poor exchange, also with no effect on serotonin affinity. The results suggest that Gln-111 and Asn-112 are not required for substrate binding but participate in subsequent steps in the transport cycle.  相似文献   

20.
In this work, the site saturation mutagenesis of tyrosine 195, tyrosine 260 and glutamine 265 in the cyclodextrin glycosyltransferase (CGTase) from Paenibacillus macerans was conducted to improve the specificity of CGTase for maltodextrin, which can be used as a cheap and easily soluble glycosyl donor for the synthesis of 2-O-d-glucopyranosyl-l-ascorbic acid (AA-2G). Specifically, the site-saturation mutagenesis of three sites—tyrosine 195, tyrosine 260, and glutamine 265—was performed, and it was found that the resulting mutants (containing the mutations Y195S [tyrosine → serine], Y260R [tyrosine → arginine], and Q265K [glutamine → lysine]) produced higher AA-2G yields than the wild type and the other mutant CGTases when maltodextrin was used as the glycosyl donor. Furthermore, double and triple mutations were introduced, and four mutants (containing Y195S/Y260R, Y195S/Q265K, Y260R/Q265K, and Y260R/Q265K/Y195S) were obtained and evaluated for the capacity to produce AA-2G. The Y260R/Q265K/Y195S triple mutant produced the highest titer of AA-2G at 1.92 g/liter, which was 60% higher than that (1.20 g/liter) produced by the wild-type CGTase. The kinetics analysis of AA-2G synthesis by the mutant CGTases confirmed the enhanced maltodextrin specificity, and it was also found that compared with the wild-type CGTase, all seven mutants had lower cyclization activities and higher hydrolysis and disproportionation activities. Finally, the mechanism responsible for the enhanced substrate specificity was explored by structure modeling, which indicated that the enhancement of maltodextrin specificity may be related to the changes of hydrogen bonding interactions between the side chain of residue at the three positions (195, 260, and 265) and the substrate sugars. This work adds to our understanding of the synthesis of AA-2G and makes the Y260R/Q265K/Y195S mutant a good starting point for further development by protein engineering.  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