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 125 毫秒
1.
Urea-induced unfolding of lipoxygenase-1 (LOX1) at pH 7.0 was followed by enzyme activity, spectroscopic measurements, and limited proteolysis experiments. Complete unfolding of LOX1 in 9 M urea in the presence of thiol reducing or thiol modifying reagents was observed. The aggregation and oxidative reactions prevented the reversible unfolding of the molecule. The loss of enzyme activity was much earlier than the structural loss of the molecule during the course of unfolding, with the midpoint concentrations being 4.5 and 7.0 M for activity and spectroscopic measurements, respectively. The equilibrium unfolding transition could be adequately fitted to a three-state, two-step model (N left arrow over right arrow I left arrow over right arrow U) and the intermediate fraction was maximally populated at 6.3 M urea. The free energy change (DeltaG(H(2)O)) for the unfolding of native (N) to intermediate (I) was 14.2 +/- 0.28 kcal/mol and for the intermediate to the unfolded state (U) was 11.9 +/- 0.12 kcal/mol. The ANS binding measurements as a function of urea concentration indicated that the maximum binding of ANS was in 6.3 M urea due to the exposure of hydrophobic groups; this intermediate showed significant amount of tertiary structure and retained nearly 60% of secondary structure. The limited proteolysis measurements showed that the initiation of unfolding was from the C-terminal domain. Thus, the stable intermediate observed could be the C-terminal domain unfolded with exposed hydrophobic domain-domain interface. Limited proteolysis experiments during refolding process suggested that the intermediate refolded prior to completely unfolded LOX1. These results confirmed the role of cysteine residues and domain-domain interactions in the reversible unfolding of LOX1. This is the first report of the reversible unfolding of a very large monomeric, multi-domain protein, which also has a prosthetic group.  相似文献   

2.
Single potassium M-channels in rat sympathetic neurons have multiple voltage-dependent kinetic components in their activity: short, medium, and long closed times (tau(CS), tau(CM), and tau(CL)) and short and long open times (tau(OS) and tau(OL)). All five components can be detected in cell-attached patches, but only four of them (tau(CS), tau(CM), tau(OS), and tau(OL)) in excised patches (, J. Physiol. (Lond.). 472:711-724; 1996, Neuron. 16:151-162; 1996, Neuropharmacology. 35:933-947). Analysis of the burst structure of activity recorded from cell-attached and excised inside-out patches showed it to be consistent with the sequential kinetic scheme C(L) left arrow over right arrow O(S) left arrow over right arrow C(M) left arrow over right arrow O(L) left arrow over right arrow C(S). Using this scheme and experimentally determined kinetic parameters, we successfully simulated the activity of M-channels both under steady-state conditions and during depolarizing voltage steps. Consistent with the characteristic behavior of macroscopic M-current, ensemble currents constructed from simulated M-channels had exponential activation and deactivation, with no delays, when tested in the range between -50 and -20 mV.  相似文献   

3.
The secondary structure of human recombinant interleukin-4 (IL-4) has been investigated by three-dimensional (3D) 15N- and 13C-edited nuclear Overhauser (NOE) spectroscopy on the basis of the 1H, 15N, and 13C assignments presented in the preceding paper [Powers, R., Garrett, D. S., March, C. J., Frieden, E. A., Gronenborn, A. M., & Clore, G. M. (1992) Biochemistry (preceding paper in this issue)]. Based on the NOE data involving the NH, C alpha H, and C beta H protons, as well as 3JHN alpha coupling constant, amide exchange, and 13C alpha and 13C beta secondary chemical shift data, it is shown that IL-4 consists of four long helices (residues 9-21, 45-64, 74-96, and 113-129), two small helical turns (residues 27-29 and 67-70), and a mini antiparallel beta-sheet (residues 32-34 and 110-112). In addition, the topological arrangement of the helices and the global fold could be readily deduced from a number of long-range interhelical NOEs identified in the 3D 13C-edited NOE spectrum in combination with the spatial restrictions imposed by three disulfide bridges. These data indicate that the helices of interleukin-4 are arranged in a left-handed four-helix bundle with two overhand connections.  相似文献   

4.
We present experimental evidence for a cooperative unfolding transition of an alpha-helix in the lac repressor headpiece bound to a symmetric variant of the lac operator, as inferred from hydrogen-deuterium (H-D) exchange experiments monitored by NMR spectroscopy. In the EX1 limit, observed exchange rates become pH-independent and exclusively sensitive to local structure fluctuations that expose the amide proton HN to exchange. Close to this regime, we measured decay rates of individual backbone HN signals in D2O, and of their mutual HN-HN NOE by time-resolved two-dimensional (2D) NMR experiments. The data revealed correlated exchange at the center of the lac headpiece recognition helix, Val20-Val23, and suggested that the correlation breaks down at Val24, at the C terminus of the helix. A lower degree of correlation was observed for the exchange of Val9 and Ala10 at the center of helix 1, while no correlation was observed for Val38 and Glu39 at the center of helix 3. We conclude that HN exchange in the recognition helix and, to some extent, in helix 1 is a cooperative event involving the unfolding of these helices, whereas the HN exchange in helix 3 is dominated by random local structure fluctuations.  相似文献   

5.
Summary Sequence-specific backbone 1H and 15N resonance assignments have been made for 95% of the amino acids in sperm whale myoglobin, complexed with carbon monoxide (MbCO). Many assignments for side-chain resonances have also been obtained. Assignments were made by analysis of an extensive series of homonuclear 2D spectra, measured with unlabeled protein, and both 2D and 3D 1H-15N-correlated spectra obtained from uniformly 15N-labeled myoglobin. Patterns of medium-range NOE connectivities indicate the presence of eight helices in positions that are very similar to those found in the crystal structures of sperm whale myoglobin. The resonance assignments of MbCO form the basis for determination of the solution structure and for hydrogen-exchange measurements to probe the stability and folding pathways of myoglobin. They will also form a basis for assignment of the spectra of single-site mutants with altered ligand-binding properties.  相似文献   

6.
SUMO-1 (1-97) is a crucial protein in the machinery of post-translational modifications. We observed by circular dichroism and fluorescence spectroscopy that urea-induced unfolding of this protein is a complex process with the possibility of occurrence of detectable intermediates along the way. The tertiary structure is completely lost around approximately 4.5 M urea with a transition mid-point at 2.53 M urea, while the secondary structure unfolding seems to show two transitions, with mid-points at 2.42 M and 5.69 M urea. We have elucidated by systematic urea titration, the equilibrium residue level structural and dynamics changes along the entire folding/unfolding transition by multidimensional NMR. With urea dilution, the protein is seen to progressively lose most of the broad beta-domain structural preferences present at 8 M urea, acquire some helical propensities at 5 M urea, and lose some of them again on further dilution of urea. Between 3 M and 2 M urea, the protein starts afresh to acquire native structural features. These observations are contrary to the conventional notion that proteins fold with monotonously increasing native-type preferences. For folding below approximately 3 M urea, the region around the alpha1 helix appears to be a potential folding initiation site. The folding seems to start with a collapse into native-like topologies, at least in parts, and is followed by formation of secondary and tertiary structure, perhaps by cooperative rearrangements. The motional characteristics of the protein show sequence-dependent variation as the concentration of urea is progressively reduced. At the sub-nanosecond level, the features are extremely unusual for denatured states, and only certain segments corresponding to the flexible regions in the native protein display these motions at the different concentrations of urea.  相似文献   

7.
Bolen DW  Yang M 《Biochemistry》2000,39(49):15208-15216
The DeltaG degrees (N)(-)(D) value obtained from extrapolation to zero denaturant concentration by the linear extrapolation method (LEM) is commonly interpreted to represent the Gibbs energy difference between native (N) and denatured (D) ensembles at the limit of zero denaturant concentration. For DeltaG degrees (N)(-)(D) to be interpreted solely in terms of N and D, as is common practice, it must be shown to be independent of denaturant concentration. Because DeltaG degrees (N)(-)(D) is often observed to be dependent on the nature of the denaturant, it is necessary to determine the circumstances under which DeltaG degrees (N)(-)(D) can be interpreted as a property solely of the protein. Here, we use proton inventory, a thermodynamic property of both the native and denatured ensembles, to monitor the thermodynamic character of denaturant-dependent aspects of N and D ensembles and the N right arrow over left arrow D transition. Use of a thermodynamic rather than a spectral parameter to monitor denaturation provides insight into the manner in which denaturant affects the meaning of DeltaG degrees (N)(-)(D) and the nature of the N right arrow over left arrow D transition. Three classes of proteins are defined in terms of the thermodynamic behaviors of their N right arrow over left arrow D transition and N and D ensembles. With guanidine hydrochloride as a denaturant, the classification of protein denaturations by these procedures determines when the LEM gives readily interpretable DeltaG degrees (N)(-)(D) values with this denaturant and when it does not.  相似文献   

8.
The conformational stability of RNase Rs was determined with chemical and thermal denaturants over the pH range of 1-10. Equilibrium unfolding with urea showed that values of D(1/2) (5.7 M) and DeltaG(H(2)O) (12.8 kcal/mol) were highest at pH 5.0, its pI and the maximum conformational stability of RNase Rs was observed near pH 5.0. Denaturation with guanidine hydrochloride (GdnHCl), at pH 5.0, gave similar values of DeltaG(H(2)O) although GdnHCl was 2-fold more potent denaturant with D(1/2) value of 3.1 M. The curves of fraction unfolded (f(U)) obtained with fluorescence and CD measurements overlapped at pH 5.0. Denaturation of RNase Rs with urea in the pH range studied was reversible but the enzyme denatured irreversibly >pH 11.0. Thermal denaturation of RNase Rs was reversible in the pH range of 2.0-3.0 and 6.0-9.0. Thermal denaturation in the pH range 4.0-5.5 resulted in aggregation and precipitation of the protein above 55 degrees C. The aggregate was amorphous or disordered precipitate as observed in TE micrographs. Blue shift in emission lambda(max) and enhancement of fluorescence intensity of ANS at 70 degrees C indicated the presence of solvent exposed hydrophobic surfaces as a result of heat treatment. Aggregation could be prevented partially with alpha-cyclodextrin (0.15 M) and completely with urea at concentrations >3 M. Aggregation was probably due to intermolecular hydrophobic interaction favored by minimum charge-charge repulsion at the pI of the enzyme. Both urea and temperature-induced denaturation studies showed that RNase Rs unfolds through a two-state F right arrow over left arrow U mechanism. The pH dependence of stability described by DeltaG(H(2)O) (urea) and DeltaG (25 degrees C) suggested that electrostatic interactions among the charged groups make a significant contribution to the conformational stability of RNase Rs. Since RNase Rs is a disulfide-containing protein, the major element for structural stability are the covalent disulfide bonds.  相似文献   

9.
The urea-induced equilibrium unfolding of the alpha-subunit of tryptophan synthase (alphaTS) from Escherichia coli can be described by a four-state model, N right harpoon over left harpoon I1 right harpoon over left harpoon I2 right harpoon over left harpoon U, involving two highly populated intermediates, I1 and I2 [Gualfetti, P. J., Bilsel, O., and Matthews, C. R. (1999) Protein Sci. 8, 1623-1635]. To extend the physical characterization of these stable forms, the apparent radius was measured by several techniques. Size-exclusion chromatography (SEC), analytical ultracentrifugation (UC), and dynamic light scattering (DLS) experiments yield an apparent Stokes radius, R(s), of approximately 24 A for the native state of alphaTS. The small-angle X-ray scattering (SAXS) experiment yields a radius of gyration, R(g), of 19.1 A, consistent with the value predicted from the X-ray structure and the Stokes radius. As the equilibrium is shifted to favor I1 at approximately 3.2 M and I2 at 5.0 M urea, SEC and UC show that R(s) increases from approximately 38 to approximately 52 A. Measurements of the radius by DLS and SAXS between 2 and 4.5 M urea were complicated by the self-association of the I1 species at the relatively high concentrations required by those techniques. Above 6 M urea, SEC and UC reveal that R(s) increases linearly with increasing urea concentration to approximately 54 A at 8 M urea. The measurements of R(s) by DLS and R(g) by SAXS are sufficiently imprecise that both values appear to be identical for the I2 and U states and, considering the errors, are in good agreement with the results from SEC and UC. Thermodynamic parameters extracted from the SEC data for the N right harpoon over left harpoon I1 and I1 right harpoon over left harpoon I2 transitions agree with those from the optical data, showing that this technique accurately monitors a part of the equilibrium model. The lack of sensitivity to the I2 right harpoon over left harpoon U transition, beyond a simple swelling of both species with increasing urea concentration, implies that the Stokes radii for the I2 and U states are not distinguishable. Surprisingly, the hydrophobic core known to stabilize I2 at 5.0 M urea [Saab-Rincón, G., Gualfetti, P. J., and Matthews, C. R. (1996) Biochemistry 35, 1988-1994] develops without a significant contraction of the polypeptide, i.e., beyond that experienced by the unfolded form at decreasing urea concentrations. Kratky plots of the SAXS data, however, reveal that I2, similar to N and I1, has a globular structure while U has a more random coil-like form. By contrast, the formation of substantial secondary structure and the burial of aromatic side chains in I1 and, eventually, N are accompanied by substantial decreases in their Stokes radii and, presumably, the size of their respective conformational ensembles.  相似文献   

10.
Using 1488 NOE constraints, 19 stereo-specific assignments, 13 pairs of H-bond constraints, and 140 pseudo-contact shift constraints, a family of 35 structures of bovine microsomal cytochrome b(5) mutant V61H has been obtained through the program PSEUDYANA. The family has been further refined by restrained energy minimization to give a family of final structures. The RMSD values of final structures with respect to the average structure are 0.45+/-0.11 and 0.96+/-0.10A for backbone and heavy atoms, respectively. The final Deltachi(ax) and Deltachi(rh) values are 2.34 x 10(-32) and -0.67 x 10(-32)m(3), respectively. The comparisons between the solution structures of mutant V61H and WT cytochrome b(5), and X-ray structure of the mutant V61H show that the global folding of the molecule in solution is unchanged and the side-chain of His61 deviates from the heme pocket and extends into the solvent like in its crystal structure. However, the helices around the heme pocket undergo outward global displacement while their local conformations are well maintained. Meanwhile, the heme ring shows a little off the heme pocket, which accounts for the lower stability of the mutant. Additionally, the axial ligand rings counterclockwise rotate around His39 N-Fe axis due to the mutation, which is confirmed by variation of the hyperfine shifts of the heme protons of V61H compared to those of WT cytochrome b(5).  相似文献   

11.
Streptomyces subtilisin inhibitor (SSI) is known to exist in at least two distinct denatured states, cold-denatured (D') and heat-denatured (D) under acidic conditions. In the present work, we investigated the manner how increasing urea concentration from 0 to 8 M changes the polypeptide chain conformation of SSI that exists initially in the D' and D states as well as in the native state (N), in terms of the secondary structure, the tertiary structure, and the chain form, based on the results of the experiments using circular dichroism (CD), small-angle X-ray scattering (SAXS) and 1H-NMR spectroscopy. Our results indicate that the urea-induced conformational transitions of SSI under typical conditions of D' (pH 1.8, 3 degrees C) occur at least in two steps. In the urea concentration range of 0-2 M (step 1), a cooperative destruction of the tertiary structure occurs, resulting in a mildly denatured state (DU), which may still contain a little amount of secondary structures. In the concentration range of 2-4 M urea (step 2), the DU state gradually loses its residual secondary structure, and increases the radius of gyration nearly to a maximum value. At 4 M urea, the polypeptide chain is highly disordered with highly mobile side chains. Increasing the urea concentration up to 8 M probably results in the more highly denatured or alternatively the stiffer chain conformations. The conformational transition starting from the N state proceeds essentially the same way as in the above scheme in which D' is replaced with N. The conformational transition starting from the D state lacks step 1 because the D state contains no tertiary structures and is similar to the DU state. The fact that similar conformations are reached at urea concentrations above 2 M from different conformations of D', D, and N indicates that the effect of urea dominates in determining the polypeptide conformation of SSI in the denatured states rather than the pH and temperature.  相似文献   

12.
Chow CY  Wu MC  Fang HJ  Hu CK  Chen HM  Tsong TY 《Proteins》2008,72(3):901-909
Fluorescence and circular dichroism stopped-flow have been widely used to determine the kinetics of protein folding including folding rates and possible folding pathways. Yet, these measurements are not able to provide spatial information of protein folding/unfolding. Especially, conformations of denatured states cannot be elaborated in detail. In this study, we apply the method of fluorescence energy transfer with a stopped-flow technique to study global structural changes of the staphylococcal nuclease (SNase) mutant K45C, where lysine 45 is replaced by cysteine, during folding and unfolding. By labeling the thiol group of cysteine with TNB (5,5'-dithiobis-2-nitrobenzoic acid) as an energy acceptor and the tryptophan at position 140 as a donor, distance changes between the acceptor and the donor during folding and unfolding are measured from the efficiency of energy transfer. Results indicate that the denatured states of SNase are highly compact regardless of how the denatured states (pH-induced or GdmCl-induced) are induced. The range of distance changes between two probes is between 25.6 and 25.4 A while it is 20.4 A for the native state. Furthermore, the folding process consists of three kinetic phases while the unfolding process is a single phase. These observations agree with our previous sequential model: N(0) left arrow over right arrow D(1) left arrow over right arrow D(2) left arrow over right arrow D(3) (Chen et al., J Mol Biol 1991;220:771-778). The efficiency of protein folding may be attributed to initiating the folding process from these compact denatured structures.  相似文献   

13.
The colicin immunity protein Im7 folds from its unfolded state in 6 M urea to its native four-helix structure through an on-pathway intermediate that lacks one of the helices of the native structure (helix III). In order to further characterize the folding mechanism of Im7, we have studied the conformational properties of the protein unfolded in 6 M urea in detail using heteronuclear NMR. Triple-resonance experiments with 13C/15N-labelled Im7 in 6 M urea provided almost complete resonance assignments for the backbone nuclei, and measurement of backbone 15N relaxation parameters allowed dynamic ordering of the unfolded polypeptide chain to be investigated. Reduced spectral density mapping and fitting backbone R2 relaxation rates to a polymer dynamics model identified four clusters of interacting residues, each predicted by the average area buried upon folding for each residue. Chemical shift analyses and measurement of NOEs detected with a long mixing-time 1H-1H-15N NOESY-HSQC spectrum confirmed the formation of four clusters. Each cluster of interacting side-chains in urea-unfolded Im7 occurs in a region of the protein that forms a helix in the protein, with the largest clusters being associated with the three long helices that are formed in the on-pathway folding intermediate, whilst the smallest cluster forms a helix only in the native state. NMR studies of a Phe15Ala Im7 variant and a protein in which residues 51-56 are replaced by three glycine residues (H3G3 Im7*), indicated that the clusters do not interact with each other, possibly because they are solvated by urea, as indicated by analysis of NOEs between the protein and the solvent. Based on these data, we suggest that dilution of the chaotrope to initiate refolding will result in collapse of the clusters, leading to the formation of persistent helical structure and the generation of the three-helix folding intermediate.  相似文献   

14.
The structure and dynamics of a 37-residue antimicrobial peptide gaegurin 4 (GGN4) isolated from the skin of the native Korean frog, Rana rugosa, was determined in SDS micelles by NMR spectroscopy. The solution structure of the peptide in SDS micelles was determined from 352 NOE-derived distance constraints and 22 backbone torsion angle constraints. Dynamic properties for the amide backbone were characterized by (1)H-(15)N heteronuclear NOE experiments. The structural study revealed two amphipathic helices spanning residues 2-10 and 16-32 and that the helices were connected by a flexible loop. An intraresidue disulfide bridge was formed between residues Cys31 and Cys37 near the C-terminus. The loop region (11-15) connecting the two helices are were slightly more flexible than these helices themselves. From the fact that since there is no contact NOEs between two helices, it is implied that the GGN4 peptide shows an independent motion of both helices which has an angle of about 60 degrees -120 degrees from each other.  相似文献   

15.
16.
The aerial parts of Teucrium oliverianum yielded two neo-clerodane diterpenoids, teucrolin F and G, together with the known teucrolin E. The previously proposed structure for teucrolin E was revised so that it contains a tetrahydrofuran ring instead of an oxetane ring. This was based on analysis of the NMR spectroscopic data of its diacetate, including its NOE spectra. In addition, the structural assignments of the new diterpenoids were based on 1H and 13C NMR spectroscopic studies, mainly 2D NMR experiments, including homonuclear and heteronuclear correlations.  相似文献   

17.
The complete mitochondrial genomes (mitogenomes) of Cnaphalocrocis medinalis and Chilo suppressalis (Lepidoptera: Pyralidae) were determined and analyzed. The circular genomes were 15,388 bp long for C. medinalis and 15,395 bp long for C. suppressalis. Both mitogenomes contained 37 genes, with gene order similar to that of other lepidopterans. Notably, 12 protein-coding genes (PCGs) utilized the standard ATN, but the cox1 gene used CGA as the initiation codon; the cox1, cox2, and nad4 genes in the two mitogenomes had the truncated termination codons T, T, and TA, respectively, but the nad5 gene was found to use T as the termination codon only in the C. medinalis mitogenome. Additionally, the codon distribution and Relative Synonymous Codon Usage of the 13 PCGs in the C. medinalis mitogenome were very different from those in other pyralid moth mitogenomes. Most of the tRNA genes had typical cloverleaf secondary structures. However, the dihydrouridine (DHU) arm of the trnS1(AGN) gene did not form a stable stem-loop structure. Forty-nine helices in six domains, and 33 helices in three domains were present in the secondary structures of the rrnL and rrnS genes of the two mitogenomes, respectively. There were four major intergenic spacers, except for the A+T-rich region, spanning at least 12 bp in the two mitogenomes. The A+T-rich region contained an 'ATAGT(A)'-like motif followed by a poly-T stretch in the two mitogenomes. In addition, there were a potential stem-loop structure, a duplicated 25-bp repeat element, and a microsatellite '(TA)(13)' observed in the A+T-rich region of the C. medinalis mitogenome. A poly-T motif, a duplicated 31-bp repeat element, and a 19-bp triplication were found in the C. suppressalis mitogenome. However, there are many differences in the A+T-rich regions between the C. suppressalis mitogenome sequence in the present study and previous reports. Finally, the phylogenetic relationships of these insects were reconstructed based on amino acid sequences of mitochondrial 13 PCGs using Bayesian inference and maximum likelihood methods. These molecular-based phylogenies support the traditional morphologically based view of relationships within the Pyralidae.  相似文献   

18.
Combined use of shielding constant computations, measurements of chemical shifts and NOE studies reveal that poly(dG-dC).(poly)dG-dC) in low salt solutions exist as a right-handed B-DNA double helix described by Gupta, Dhingra, Sarma, Sarma, Rajagopalan and Sasisekharan, J. Biomole. Str. Dyn. 1. 395, 1983. We present a simple and direct method to determine the handedness of DNA double helices from NOE difference spectra. This method takes advantage of the NOE between base protons and the H2'H2" sugar protons; and in the difference NOE spectra in the H2'H2" region the signatures of the right and left-handed helices become imprinted.  相似文献   

19.
The three-dimensional structure of the rhodanese homology domain At4g01050(175-195) from Arabidopsis thaliana has been determined by solution nuclear magnetic resonance methods based on 3043 upper distance limits derived from NOE intensities measured in three-dimensional NOESY spectra. The structure shows a backbone root mean square deviation to the mean coordinates of 0.43 A for the structured residues 7-125. The fold consists of a central parallel beta-sheet with five strands in the order 1-5-4-2-3 and arranged in the conventional counterclockwise twist, and helices packing against each side of the beta-sheet. Comparison with the sequences of other proteins with a rhodanese homology domain in Arabidopsis thaliana indicated residues that could play an important role in the scaffold of the rhodanese homology domain. Finally, a three-dimensional structure comparison of the present noncatalytic rhodanese homology domain with the noncatalytic rhodanese domains of sulfurtransferases from other organisms discloses differences in the length and conformation of loops that could throw light on the role of the noncatalytic rhodanese domain in sulfurtransferases.  相似文献   

20.
Poly(d2NH2A-dT): two-dimensional NMR shows a B to A conversion in high salt   总被引:6,自引:0,他引:6  
B Borah  J S Cohen  F B Howard  H T Miles 《Biochemistry》1985,24(25):7456-7462
Poly(d2NH2A-dT) forms a structure in high salt that is clearly distinct from the B form present in low salt. Two-dimensional nuclear Overhauser effect (2D NOE) NMR spectra establish that the conformation of the high-salt form is not Z. Correlations of observed cross peaks in the 2D NOE spectra and estimated interproton distances of the common DNA conformations are consistent only with an A form or closely related structure. This interpretation is also consistent with the negative circular dichroic band observed in the high-salt form of poly(d2NH2A-dT) and in A-form ribohomopolymer helices containing 2NH2A.  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