首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Protein classification and characterization often rely on the information contained in the protein secondary structure. Protein class assignment is usually based on X-ray diffraction measurements, which need the protein in a crystallized form, or on NMR spectra, to obtain the structure of a protein in solution. Simple spectroscopic techniques, such as circular dichroism (CD) and infrared (IR) spectroscopies, are also known to be related to protein secondary structure, but they have seldom been used for protein classification. To see the potential of CD, IR, and combined CD/IR measurements for protein classification, unsupervised pattern recognition methods, Principal Component Analysis (PCA) and cluster analysis, are proposed first to check for natural grouping tendencies of proteins according to their measured spectra. Partial Least Squares Discriminant Analysis (PLS-DA), a supervised pattern recognition method, is used afterwards to test the possibility to model explicitly each protein class and to test these models in class assignment of unknown proteins. Determination of the protein secondary structure, understood as the prediction of the abundance of the different secondary structure motifs in the biomolecule, was carried out with the local regression method interval Partial Least Squares (iPLS). CD, IR, and CD/IR measurements were correlated to the fraction of the motif to be predicted, determined from X-ray measurements. iPLS builds models extracting the spectral information most correlated to a specific secondary motif and avoids the use of irrelevant spectral regions. Spectral intervals chosen by iPLS models provide structural information which can be used to confirm previous biochemical assignments or identify new motif-related spectral features. The predictive ability of the models built with the selected spectral regions has a quality similar to previous classical approaches.  相似文献   

2.
One of the most versatile methods for monitoring the structure of proteins, either in solution or in the solid state, is Fourier transform infrared spectroscopy. Also known as mid-range infrared, which covers the frequency range from 4000 to 400 cm(-1), this wavelength region includes bands that arise from three conformationally sensitive vibrations within the peptide backbone (amide I, II and III). Of these vibrations, amide I is the most widely used and can provide information on secondary structure composition and structural stability. One of the advantages of infrared spectroscopy is that it can be used with proteins that are either in solution or in the solid state. The use of infrared to monitor protein structure and stability is summarized herein. In addition, specialized infrared methods are presented, such as techniques for the study of membrane proteins and oriented samples. In addition, there is a growing body of literature on the use of infrared to follow reaction kinetics and ligand binding in proteins, as well as a number of infrared studies on protein dynamics. Finally, the potential for using near-infrared spectroscopy to study protein structure is introduced.  相似文献   

3.
Centrin is a low molecular mass (20 kDa) protein that belongs to the EF-hand superfamily of calcium-binding proteins. Local and overall changes were investigated for interactions between cations and Chlamydomonas centrin using Fourier transform infrared (FT-IR) and circular dichroic (CD) spectroscopies. FT-IR spectral features studied included the amide I' band and the side-chain absorbances for aspartate residues located almost exclusively at the calcium-binding sites in the spectral region of 1700-1500 cm(-1). The amide I' band is exquisitely sensitive to changes in protein secondary structure and is observed to shift from 1626.5 to 1642.7 cm(-1) in the presence and absence of calcium. These spectral bands are complex and were further studied using two-dimensional Fourier transform infrared (2D-FT-IR) correlation along with curve-fitting routines. Using these methods the secondary structure contributions were determined for holocentrin and apocentrin. The alpha-helical content in centrin was determined to be 60%-53% in the presence and absence of cations, respectively. Furthermore, the beta-strand content was determined to be 12%-36%, while the random coil component remained almost constant at 7%-13.5% in the presence and absence of cations, respectively. Changes in the side-chain band are mostly due to the monodentate coordination of aspartate to the cation. A shift of approximately 4 cm(-1) (for the COO- antisymmetric stretch in Asp) from 1565 to 1569 cm(-1) is observed for apocentrin and holocentrin, respectively. Thermal dependence revealed reversible conformational transition temperatures for apocentrin at 37 degrees C and holocentrin at 45 degrees C, suggesting greater stability for holocentrin.  相似文献   

4.
Changes in the amide bands in Fourier transform infrared spectra of proteins are generally attributed to alterations in protein secondary structure. In this study spectra of five different globular proteins were compared in the solid and solution states recorded with several sampling techniques. Spectral differences for each protein were observed between the various sampling techniques and physical states, which could not all be explained by a change in protein secondary structure. For example, lyophilization in the absence of lyoprotectants caused spectral changes that could (partially) have been caused by the removal of hydrating water molecules rather than secondary structural changes. Moreover, attenuated total reflectance spectra of proteins in H2O were not directly comparable to transmission spectra due to the anomalous dispersion effect. Our study also revealed that the amide I, II, and III bands differ in their sensitivities to changes in protein conformation: For example, strong bands in the region 1620-1630 and 1685-1695 cm(-1) were seen in the amide I region of aggregated protein spectra. Surprisingly, absorbance of such magnitudes was not observed in the amide II and III region. It appears, therefore, that only the amide I can be used to distinguish between intra- and intermolecular beta-sheet formation. Considering the differing sensitivity of the different amide modes to structural changes, it is advisable to utilize not only the amide I band, but also the amide II and III bands, to determine changes in protein secondary structure. Finally, it is important to realize that changes in these bands may not always correspond to secondary structural changes of the proteins.  相似文献   

5.
A fast search algorithm to reveal similar polypeptide backbone structural motifs in proteins is proposed. It is based on the vector representation of a polypeptide chain fold in which the elements of regular secondary structures are approximated by linear segments (Abagyan and Maiorov, J. Biomol. Struct. Dyn. 5, 1267-1279 (1988)). The algorithm permits insertions and deletions in the polypeptide chain fragments to be compared. The fast search algorithm implemented in FASEAR program is used for collecting beta alpha beta supersecondary structure units in a number of alpha/beta proteins of Brookhaven Data Bank. Variation of geometrical parameters specifying backbone chain fold is estimated. It appears that the conformation of the majority of the fragments, although almost all of them are right-handed, is quite different from that of standard beta alpha beta units. Apart from searching for specific type of secondary structure motif, the algorithm allows automatically to identify new recurrent folding patterns in proteins. It may be of particular interest for the development of tertiary template approach for prediction of protein three-dimensional structure as well for constructing artificial polypeptides with goal-oriented conformation.  相似文献   

6.
The conformational changes associated with the redox transition of plastocyanin (PC) were investigated by absorption and reaction-induced infrared spectroscopy. In addition to spectral features readily ascribed to beta and turn protein secondary structures, the amide I band shows a major component band at 1647 cm(-1) in both redox states of the protein. The sensitivity of this component to deuteration and increasing temperature suggests that PC adopts an unusual secondary structure in solution, which differs from those described for other type I copper proteins, such as azurin and halocyanin. The conformations of oxidized and reduced PC are different, as evidenced (1) by analysis of their amide I band contour and the electrochemically induced oxidized-minus-reduced difference spectrum and (2) by their different thermal stability. The redox-induced difference spectrum exhibits a number of difference bands within the conformationally sensitive amide I band that could be assigned to peptide C=O modes, in light of their small shift upon deuteration, and to signals attributable to side chain vibrational modes of Tyr residues. Lowering the pH to 4.8 induces destabilization of both redox states of the protein, more pronounced for reduced PC, without significantly affecting their secondary structure. Besides the conformational differences obtained at neutral pH, the oxidized-minus-reduced difference spectrum shows two broad and strong negative bands at 1405 and 1571 cm(-1), assigned to COO(-) vibrations, and a broad positive band at 1710 cm(-1), attributed to the C=O vibration of a COOH group(s). These bands are indicative of a protonation of (an) Asp or Glu side chain(s) upon plastocyanin oxidation at acidic pH.  相似文献   

7.
Fourier transform infrared (FTIR) spectroscopy is the most common spectroscopic technique used for study of protein structure. Initially, band deconvolution techniques were applied to determine the secondary structure of proteins. Recently, several multivariate regression methods have been used to predict the secondary structure of proteins as an alternative to the previous methods. Multivariate curve resolution-alternating least squares (MCR-ALS) was applied on the FTIR spectra of proteins to resolve the fraction and spectral profiles of different structural motifs. Initial estimates of spectral profiles of different protein motifs were built using orthogonal projection approach (OPA). Predicted fractions of α-helix and β-sheet obtained by MCR-ALS technique were compared with those from partial least squares (PLS) modeling which revealed superiority of the former. If we consider the possibility of pure spectra prediction in addition to the prediction of secondary structure from the data set, MCR-ALS can be proposed as a very valuable alternative for qualitative and quantitative study of protein structures.  相似文献   

8.
Raman spectroscopy was employed to examine the secondary structure of the cAMP receptor protein (CRP). Spectra were obtained over the range 400-1900 cm-1 from solutions of CRP and from CRP-cAMP cocrystals. The spectra of CRP dissolved in 30 mM sodium phosphate and 0.15 M NaCl buffered at either pH 6 or pH 8 or dissolved in 0.15-0.2 M NaCl at protein concentrations of 5, 15, and 30 mg/mL were examined. Estimates of the secondary structure distribution were made by analyzing the amide I region of the spectra (1630-1700 cm-1). CRP secondary structure distributions were essentially the same in either pH and at all protein concentrations examined. The amide I analyses indicated a structural distribution of 44% alpha-helix, 28% beta-strand, 18% turn, and 10% undefined for CRP in solution. Raman spectra of CRP-cAMP cocrystals differed from the spectra of CRP in solution. Some differences were assigned to interfering background bands, whereas other spectral differences were attributed to changes in CRP structure. Differences in the amide III region and in the intensity at 935 cm-1 were consistent with alterations in secondary structure. Analysis of the amide I region of the CRP-cAMP cocrystal spectrum indicated a secondary structure distribution of 37% alpha-helix, 33% beta-strand, 17% turn, and 12% undefined. This result is in agreement with a published secondary structure distribution derived from X-ray analysis of CRP-cAMP cocrystals (37% alpha-helix and 36% beta-strand).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
One of the most versatile methods for monitoring the structure of proteins, either in solution or in the solid state, is Fourier transform infrared spectroscopy. Also known as mid-range infrared, which covers the frequency range from 4000 to 400 cm-1, this wavelength region includes bands that arise from three conformationally sensitive vibrations within the peptide backbone (amide I, II and III). Of these vibrations, amide I is the most widely used and can provide information on secondary structure composition and structural stability. One of the advantages of infrared spectroscopy is that it can be used with proteins that are either in solution or in the solid state. The use of infrared to monitor protein structure and stability is summarized herein. In addition, specialized infrared methods are presented, such as techniques for the study of membrane proteins and oriented samples. In addition, there is a growing body of literature on the use of infrared to follow reaction kinetics and ligand binding in proteins, as well as a number of infrared studies on protein dynamics. Finally, the potential for using near-infrared spectroscopy to study protein structure is introduced.  相似文献   

10.
Functional RNA regions are often related to recurrent secondary structure patterns (or motifs), which can exert their role in several different ways, particularly in dictating the interaction with RNA-binding proteins, and acting in the regulation of a large number of cellular processes. Among the available motif-finding tools, the majority focuses on sequence patterns, sometimes including secondary structure as additional constraints to improve their performance. Nonetheless, secondary structures motifs may be concurrent to their sequence counterparts or even encode a stronger functional signal. Current methods for searching structural motifs generally require long pipelines and/or high computational efforts or previously aligned sequences. Here, we present BEAM (BEAr Motif finder), a novel method for structural motif discovery from a set of unaligned RNAs, taking advantage of a recently developed encoding for RNA secondary structure named BEAR (Brand nEw Alphabet for RNAs) and of evolutionary substitution rates of secondary structure elements. Tested in a varied set of scenarios, from small- to large-scale, BEAM is successful in retrieving structural motifs even in highly noisy data sets, such as those that can arise in CLIP-Seq or other high-throughput experiments.  相似文献   

11.
Abstract

A fast search algorithm to reveal similar polypeptide backbone structural motifs in proteins is proposed. It is based on the vector representation of a polypeptide chain fold in which the elements of regular secondary structures are approximated by linear segments (Abagyan and Maiorov, J. Biomol. Struct. Dyn. 5, 1267–1279 (1988)). The algorithm permits insertions and deletions in the polypeptide chain fragments to be compared. The fast search algorithm implemented in FASEAR program is used for collecting βαβ supersecondary structure units in a number of α/β proteins of Brookhaven Data Bank. Variation of geometrical parameters specifying backbone chain fold is estimated. It appears that the conformation of the majority of the fragments, although almost all of them are right-handed, is quite different from that of standard βαβ units. Apart from searching for specific type of secondary structure motif, the algorithm allows automatically to identify new recurrent folding patterns in proteins. It may be of particular interest for the development of tertiary template approach for prediction of protein three-dimensional structure as well for constructing artificial polypeptides with goal-oriented conformation.  相似文献   

12.
The transmembrane helix of glycophorin A contains a seven-residue motif, LIxxGVxxGVxxT, that mediates protein dimerization. Threonine is the only polar amino acid in this motif with the potential to stabilize the dimer through hydrogen-bonding interactions. Polarized Fourier transform infrared spectroscopy is used to establish a robust protocol for incorporating glycophorin A transmembrane peptides into membrane bilayers. Analysis of the dichroic ratio of the 1655-cm(-1) amide I vibration indicates that peptides reconstituted by detergent dialysis have a transmembrane orientation with a helix crossing angle of <35 degrees. Solid-state nuclear magnetic resonance spectroscopy is used to establish high resolution structural restraints on the conformation and packing of Thr-87 in the dimer interface. Rotational resonance measurement of a 2.9-A distance between the gamma-methyl and backbone carbonyl carbons of Thr-87 is consistent with a gauche- conformation for the chi1 torsion angle. Rotational-echo double-resonance measurements demonstrate close packing (4.0 +/- 0.2 A) of the Thr-87 gamma-methyl group with the backbone nitrogen of Ile-88 across the dimer interface. The short interhelical distance places the beta-hydroxyl of Thr-87 within hydrogen-bonding range of the backbone carbonyl of Val-84 on the opposing helix. These results refine the structure of the glycophorin A dimer in membrane bilayers and highlight the complementary role of small and polar residues in the tight association of transmembrane helices in membrane proteins.  相似文献   

13.
The Raman spectrum of the isometric bacteriophage phi X174 contains a number of well-resolved bands which have been assigned unambiguously to proteins of the capsid or to the single-stranded DNA (ssDNA) genome. Additional Raman bands of protein and DNA, which are partially overlapped in the spectrum of virus, have been resolution enhanced by Fourier deconvolution to permit improved semiquantitative measurement of spectral intensities and frequencies for structural conclusions. Raman conformation markers indicate that the ssDNA molecule within the capsid contains nucleosides of C2'-endo sugar pucker and anti-glycoside bond orientation, but the nucleic acid backbone lacks the geometry characteristic of B-form DNA. The Raman profile of encapsidated phi X DNA indicates a backbone more similar to heat-denatured DNA than to DNA containing hairpinlike secondary structure. This finding suggests limited interbase interactions in the packaged genome, which is presumably the result of constraints imposed by the viral capsid. Thus, the extensive pairing and stacking of bases indicated by Raman profiles from ssRNA viruses are not evident for the phi X174 chromosome. Overall, the proteins of the virion contain extensive beta-sheet and irregular secondary structures. Fourier deconvolution of the Raman amide I band provides an estimate of the percentage of total beta-sheet structure (approximately 60%) in all proteins of the virion. The amide III region of the spectrum confirms that beta-sheet and irregular domains are the predominant protein secondary structures. Samples of phi X174 concentrated for Raman spectroscopy by either ultracentrifugation or ultrafiltration exhibit nearly identical Raman spectra, indicating that either method can be employed to prepare intact virus without significant loss of DNA or protein components.  相似文献   

14.
Mnemiopsin 2 from a luminous ctenophore with two functional EF-hand motifs is a calcium-regulated photoprotein that is responsible for emitting a bright blue bioluminescence upon reacting with coelenterazine and calcium ions in Mnemiopsis leidyi. Synchrotron radiation-based Fourier-transform infrared (SR-FTIR) spectroscopy was applied to analyze the distribution of secondary structures, the conformational changes resulting from calcium binding and the structural stabilities in wild-type mnemiopsin 2, as well as its mutant type that possesses three EF-hand motifs. The distribution of secondary structures of these proteins indicates that mutant apo-mnemiopsin 2 has a more stable secondary structure than the wild-type. Analyses of the SR-FTIR spectra revealed that the conformational changes at the secondary structures of both mnemiopsin 2 depend on the calcium concentrations, such that the most noticeable changes in structures of wild-type and mutant mnemiopsin 2 occur at optimum concentrations 6 and 2 mM of calcium chloride, respectively. The addition of calcium to both proteins increases the proportions of their secondary structures in the amide I and II regions. The major amide I bands in the IR spectra of both mnemiopsin‑calcium complexes shift towards smaller wavenumbers, whereas their main amide II bands are identified at larger wavenumbers.  相似文献   

15.
A soft-modeling multivariate numerical approach that combines self-modeling curve resolution (SMCR) and mixed Lorentzian-Gaussian curve fitting was successfully implemented for the first time to elucidate spatially and spectroscopically resolved spectral information from infrared imaging data of oral mucosa cells. A novel variant form of the robust band-target entropy minimization (BTEM) SMCR technique, coined as hierarchical BTEM (hBTEM), was introduced to first cluster similar cellular infrared spectra using the unsupervised hierarchical leader-follower cluster analysis (LFCA) and subsequently apply BTEM to clustered subsets of data to reconstruct three protein secondary structure (PSS) pure component spectra—α-helix, β-sheet, and ambiguous structures—that associate with spatially differentiated regions of the cell infrared image. The Pearson VII curve-fitting procedure, which approximates a mixed Lorentzian-Gaussian model for spectral band shape, was used to optimally curve fit the resolved amide I and II bands of various hBTEM reconstructed PSS pure component spectra. The optimized Pearson VII band-shape parameters and peak center positions serve as means to characterize amide bands of PSS spectra found in various cell locations and for approximating their actual amide I/II intensity ratios. The new hBTEM methodology can also be potentially applied to vibrational spectroscopic datasets with dynamic or spatial variations arising from chemical reactions, physical perturbations, pathological states, and the like.  相似文献   

16.
We have studied the conformation of the peptide Ac-EPKRSVAFKKTKKEVKKVATPKK (CH-1), free in solution and bound to the DNA, by Fourier-transform infrared spectroscopy. The peptide belongs to the COOH-terminal domain of histone H1(0) (residues 99-121) and is adjacent to the central globular domain of the protein. In aqueous (D(2)O) solution the amide I' is dominated by component bands at 1643 cm(-1) and 1662 cm(-1), which have been assigned to random coil conformations and turns, respectively. In accordance with previous NMR results, the latter component has been interpreted as arising in turn-like conformations in rapid equilibrium with unfolded states. The peptide becomes fully structured either in 90% trifluoroethanol (TFE) solution or upon interaction with the DNA. In these conditions, the contributions of turn (1662 cm(-1)) and random coil components virtually disappear. In TFE, the spectrum is dominated by the alpha-helical component (1654 cm(-1)). The band at 1662 cm(-1) shifts to 1670 cm(-1), and has been assigned to the COOH-terminal TPKK motif in a more stable turn conformation. A band at 1637 cm(-1), also present in TFE, has been assigned to 3(10) helical structure. The amide I' band of the complexes with the DNA retains the components that were attributed to 3(10) helix and the TPKK turn. In the complexes with the DNA, the alpha-helical component observed in TFE splits into two components at 1657 cm(-1) and 1647 cm(-1). Both components are inside the spectral region of alpha-helical structures. Our results support the presence of inducible helical and turn elements, both sharing the character of DNA-binding motifs.  相似文献   

17.
Band splitting and/or bands shifting in opposite directions due to coupling of vibrations of neighboring groups observed in the infrared spectra of tRNAPhe and 23 S RNA give information on the secondary structure. The base pairing, dependent on temperature, is investigated, discussing coupling effects with the base residues' vibrations in the region 1700-1500 cm-1. The secondary structure of the backbone is studied, discussing coupling effects with vibrations in the region 1300-1000 cm-1. The 2'OH groups are cross-linked with the O atoms of the neighboring ribose residues via hydrogen bonds. Probably the greater than PO-2 groups are turned inward at the backbone, i.e. towards the base residues. The base pairs as well as the secondary structure of the backbone melt with increasing temperature and with dialysis against distilled water. The comparison of the Mg2+ and the K+ salts of the tRNAPhe shows that the changes of base pairing due to Mg2+ are small. At the backbone, however, Mg2+ favor somewhat more the discussed secondary structure than K+ does. All Mg2+ effects on secondary structure are, however, too small to explain the considerable increase in melting temperature due to Mg2+. Thus it is supposed that the rise in the melting temperature due to Mg2+ is not caused by a change in secondary but in the tertiary structure of tRNAPhe. Furthermore, the influence of Mg2+ on the secondary structure of 23 S RNA is studied. The following results are obtained: (1) The double helical regions become more compact and probably increase due to the influence of Mg2+. (2) At the backbone, Mg2+ induces strong hydrogen bonding between the 2'OH groups and the ether O atoms of neighboring ribose residues. Probably they turn the greater than PO-2 groups toward the base residues, i.e., inward at the backbone. Schulte, Morrison and Garrett found that a critical level of Mg2+ is required for binding certain proteins to rRNA (Biochemistry (1974) 13, 1032). Thus the observed conformation is probably necessary for binding these proteins.  相似文献   

18.
We present an improved technique for estimating protein secondary structure content from amide I and amide III band infrared spectra. This technique combines the superposition of reference spectra of pure secondary structure elements with simultaneous aromatic side chain, water vapor, and solvent background subtraction. Previous attempts to generate structural reference spectra from a basis set of reference protein spectra have had limited success because of inaccuracies arising from sequential background subtractions and spectral normalization, arbitrary spectral band truncation, and attempted resolution of spectroscopically degenerate structure classes. We eliminated these inaccuracies by defining a single mathematical function for protein spectra, permitting all subtractions, normalizations, and amide band deconvolution steps to be performed simultaneously using a single optimization algorithm. This approach circumvents many of the problems associated with the sequential nature of previous methods, especially with regard to removing the subjectivity involved in each processing step. A key element of this technique was the calculation of reference spectra for ordered helix, unordered helix, sheet, turns, and unordered structures from a basis set of spectra of well-characterized proteins. Structural reference spectra were generated in the amide I and amide III bands, both of which have been shown to be sensitive to protein secondary structure content. We accurately account for overlaps between amide and nonamide regions and allow different structure types to have different extinction coefficients. The agreement between our structure estimates, for proteins both inside and outside the basis set, and the corresponding determinations from X-ray crystallography is good.  相似文献   

19.
Fourier-transform infrared spectroscopy is a valuable method for the study of protein conformation in solution primarily because of the sensitivity to conformation of the amide I band (1700-1620 cm-1) which arises from the backbone C = O stretching vibration. Combined with resolution-enhancement techniques such as derivative spectroscopy and self-deconvolution, plus the application of iterative curve-fitting techniques, this method provides a wealth of information concerning protein secondary structure. Further extraction of conformational information from the amide I band is dependent upon discerning the correlations between specific conformational types and component bands in the amide I region. In this paper, we report spectra-structure correlations derived from conformational perturbations in bovine trypsin which arise from autolytic processing, zymogen activation, and active-site inhibition. IR spectra were collected for the single-chain (beta-trypsin) and once-cleaved, double-chain (alpha-trypsin) forms as well as at various times during the course of autolysis and also for zymogen, trypsinogen, and beta-trypsin inhibited with diisopropyl fluorophosphate. Spectral differences among the various molecular forms were interpreted in light of previous biochemical studies of autolysis and the known three-dimensional structures of the zymogen, the active enzyme, and the DIP-inhibited form. Our spectroscopic results from these proteins in D2O imply that certain loop structures may absorb in the region of 1655 cm-1. Previously, amide I' infrared bands near 1655 cm-1 have been interpreted as arising solely from alpha-helices. These new data suggest caution in interpreting this band. We have also proposed that regions of protein molecules which are known from crystallographic experiments to be disordered absorb in the 1645 cm-1 region and that type II beta-turns absorb in the region of 1672-1685 cm-1. Our results also corroborate assignment of the low-frequency component of extended strands to bands below 1636 cm-1. Additionally, the results of multiple measurements have allowed us to estimate the variability present in component band areas calculated by curve fitting the resolution-enhanced IR spectra. We estimate that this approach to data analysis and interpretation is sensitive to changes of 0.01 unit or less in the relative integrated intensities of component bands in spectra whose peaks are well resolved.  相似文献   

20.
Computational grafting of functional motifs onto scaffold proteins is a promising way to engineer novel proteins with pre-specified functionalities. Typically, protein grafting involves the transplantation of protein side chains from a functional motif onto structurally homologous regions of scaffold proteins. Using this approach, we previously transplanted the human immunodeficiency virus 2F5 and 4E10 epitopes onto heterologous proteins to design novel "epitope-scaffold" antigens. However, side-chain grafting is limited by the availability of scaffolds with compatible backbone for a given epitope structure and offers no route to modify backbone structure to improve mimicry or binding affinity. To address this, we report here a new and more aggressive computational method-backbone grafting of linear motifs-that transplants the backbone and side chains of linear functional motifs onto scaffold proteins. To test this method, we first used side-chain grafting to design new 2F5 epitope scaffolds with improved biophysical characteristics. We then independently transplanted the 2F5 epitope onto three of the same parent scaffolds using the newly developed backbone grafting procedure. Crystal structures of side-chain and backbone grafting designs showed close agreement with both the computational models and the desired epitope structure. In two cases, backbone grafting scaffolds bound antibody 2F5 with 30- and 9-fold higher affinity than corresponding side-chain grafting designs. These results demonstrate that flexible backbone methods for epitope grafting can significantly improve binding affinities over those achieved by fixed backbone methods alone. Backbone grafting of linear motifs is a general method to transplant functional motifs when backbone remodeling of the target scaffold is necessary.  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