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.
We selected 49 icosahedral virus capsids whose crystal structures are reported in the Protein Data Bank. They belong to the T=1, T=3, pseudo T=3 and other lattice types. We identified in them 779 unique interfaces between pairs of subunits, all repeated by icosahedral symmetry. We analyzed the geometric and physical chemical properties of these interfaces and compared with interfaces in protein-protein complexes and homodimeric proteins, and with crystal packing contacts. The capsids contain one to 16 subunits implicated in three to 66 unique interfaces. Each subunit loses 40-60% of its accessible surface in contacts with an average of 8.5 neighbors. Many of the interfaces are very large with a buried surface area (BSA) that can exceed 10,000 A(2), yet 39% are small with a BSA<800 A(2) comparable to crystal packing contacts. Pairwise capsid interfaces overlap, so that one-third of the residues are part of more than one interface. Those with a BSA>800 A(2) resemble homodimer interfaces in their chemical composition. Relative to the protein surface, they are non-polar, enriched in aliphatic residues and depleted of charged residues, but not of neutral polar residues. They contain one H-bond per about 200 A(2) BSA. Small capsid interfaces (BSA<800 A(2)) are only slightly more polar. They have a similar amino acid composition, but they bury fewer atoms and contain fewer H-bonds for their size. Geometric parameters that estimate the quality of the atomic packing suggest that the small capsid interfaces are loosely packed like crystal packing contacts, whereas the larger interfaces are close-packed as in protein-protein complexes and homodimers. We discuss implications of these findings on the mechanism of capsid assembly, assuming that the larger interfaces form first to yield stable oligomeric species (capsomeres), and that medium-size interfaces allow the stepwise addition of capsomeres to build larger intermediates.  相似文献   

2.
Protein crystals contain two different types of interfaces: biologically relevant ones, observed in protein–protein complexes and oligomeric proteins, and nonspecific ones, corresponding to crystal lattice contacts. Because of the increasing complexity of the objects being tackled in structural biology, distinguishing biological contacts from crystal contacts is not always a trivial task and can lead to wrong interpretation of macromolecular structures. We devised an approach (CRK, core‐rim Ka/Ks ratio) for distinguishing biologically relevant interfaces from nonspecific ones. Given a protein–protein interface, CRK finds a set of homologs to the sequences of the proteins involved in the interface, retrieves and aligns the corresponding coding sequences, on which it carries out a residue‐by‐residue Ka/Ks ratio (ω) calculation. It divides interface residues into a “rim” and a “core” set and analyzes the selection pressure on the residues belonging to the two sets. We developed and tested CRK on different datasets and test cases, consisting of biologically relevant contacts, nonspecific ones or of both types. The method proves very effective in distinguishing the two categories of interfaces, with an overall accuracy rate of 84%. As it relies on different principles when compared with existing tools, CRK is optimally suited to be used in combination with them. In addition, CRK has potential applications in the validation of structures of oligomeric proteins and protein complexes. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

3.
We compare the geometric and physical-chemical properties of interfaces involved in specific and non-specific protein-protein interactions in crystal structures reported in the Protein Data Bank. Specific interactions are illustrated by 70 protein-protein complexes and by subunit contacts in 122 homodimeric proteins; non-specific interactions are illustrated by 188 pairs of monomeric proteins making crystal-packing contacts selected to bury more than 800 A2 of protein surface. A majority of these pairs have 2-fold symmetry and form "crystal dimers" that cannot be distinguished from real dimers on the basis of the interface size or symmetry. The chemical and amino acid compositions of the large crystal-packing interfaces resemble the protein solvent-accessible surface. These interfaces are less hydrophobic than in homodimers and contain much fewer fully buried atoms. We develop a residue propensity score and a hydrophobic interaction score to assess preferences seen in the chemical and amino acid compositions of the different types of interfaces, and we derive indexes to evaluate the atomic packing, which we find to be less compact at non-specific than at specific interfaces. We test the capacity of these parameters to identify homodimeric proteins in crystal structures, and show that a simple combination of the non-polar interface area and the fraction of buried interface atoms assigns the quaternary structure of 88% of the homodimers and 77% of the monomers in our data set correctly. These success rates increase to 93-95% when the residue propensity score of the interfaces is taken into consideration.  相似文献   

4.
During the characterization of mutants and covalently inhibited complexes of Fusarium solani cutinase, nine different crystal forms have been obtained so far. Protein mutants with a different surface charge distribution form new intermolecular salt bridges or long-range electrostatic interactions that are accompanied by a change in the crystal packing. The whole protein surface is involved in the packing contacts and the hydrophobicities of the protein surfaces in mutual contact turned out to be noncorrelated, which indicates that the packing interactions are nonspecific. In the case of the hydrophobic variants, the packing contacts showed some specificity, as the protein in the crystal tends to form either crystallographic or noncrystallographic dimers, which shield the hydrophobic surface from the solvent. The likelihood of surface atoms to be involved in a crystal contact is the same for both polar and nonpolar atoms. However, when taking areas in the 200–600 Å2 range, instead of individual atoms, the either highly hydrophobic or highly polar surface regions were found to have an increased probability of establishing crystal lattice contacts. The protein surface surrounding the active-site crevice of cutinase constitutes a large hydrophobic area that is involved in packing contacts in all the various crystalline contexts. Proteins 31:320–333, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

5.
沉淀剂类型对蛋白质晶体分子堆积的影响   总被引:3,自引:0,他引:3  
以不对称单位只有一个分子的牛胰核糖核酸酶和T4溶菌酶晶体为材料,着重研究了无机盐、有机溶剂和PEG三类不同的沉淀剂对晶体分子堆积的影响,经研究发现两种蛋白质中用无机盐做沉淀剂的晶型几乎都含有面积较大的二次轴对称接触面和较少的相邻分子数,同时其含有的参与接触的非极性残基集中分布于二次轴对称接触面,而盐键则在二次轴对称接触面上分布稀少。用有机溶剂作沉淀剂的晶型却含有面积较小的非二次轴对称接触面和较多的相含分子数,而参与接触的非极性残基和直键在各个非二次轴对称接触面上随机分布,用PEG作沉淀剂的晶型其分子堆积特征总体上类似于用有机溶剂作沉淀剂的晶型,但个别晶型具有与用无机盐做沉淀剂的晶型相似的分子堆积特征,以上结果提示,用三类沉淀剂得到的不同的分子堆积特征可能与三类沉淀剂不同的诱导结晶机理密切相关。  相似文献   

6.
Mintseris J  Weng Z 《Proteins》2003,53(3):629-639
The ability to analyze and compare protein-protein interactions on the structural level is critical to our understanding of various aspects of molecular recognition and the functional interplay of components of biochemical networks. In this study, we introduce atomic contact vectors (ACVs) as an intuitive way to represent the physico-chemical characteristics of a protein-protein interface as well as a way to compare interfaces to each other. We test the utility of ACVs in classification by using them to distinguish between homodimers and crystal contacts. Our results compare favorably with those reported by other authors. We then apply ACVs to mine the PDB for all known protein-protein complexes and separate transient recognition complexes from permanent oligomeric ones. Getting at the basis of this difference is important for our understanding of recognition and we achieved a success rate of 91% for distinguishing these two classes of complexes. Although accessible surface area of the interface is a major discriminating feature, we also show that there are distinct differences in the contact preferences between the two kinds of complexes. Illustrating the superiority of ACVs as a basic comparison measure over a sequence-based approach, we derive a general rule of thumb to determine whether two protein-protein interfaces are redundant. With this method, we arrive at a nonredundant set of 209 recognition complexes--the largest set reported so far.  相似文献   

7.
《Inorganica chimica acta》1988,147(2):143-150
The crystal structures of trans- and cis-bis(diphenylphosphino) ethene (1, 2) have been determined by single crystal X-ray diffraction. The conformation of these free ligands is compared with structural data available in the literature for the corresponding 1:2 complexes with gold(I) chloride (4, 5). In the cis-ligand 2 the conformation of the Ph2P-groups is such, that the molecule approaches non-crystallographic Cs symmetry with the lone pairs at phosphorus pointing towards each other. Upon addition of AuCl, rotation of one Ph2P group around the PC bond by approximately 60° leads to a structure for 5 which allows an intramolecular Au···Au contact of 3.05(1)Å. The trans-ligand 1 undergoes little structural change upon adduct formation, but intermolecular Au···Au contacts of 3.043(1) Å are secured through aggregation. The synthesis, properties and 197Au Mössbauer spectra of 1:1 and 1:2 complexes of 1 and 2 with AuCl are summarized with reference to a recent controversy in the literature.The crystal structure of bis(diphenylphosphino)- methane (3) has also been determined and the results compared with those published previously for the 1:2 complex with AuCl (7, crystallographic C2 symmetry, Au···Au distance 3.351(2) Å). There is very little change of the ligand conformation upon coordination.  相似文献   

8.
We present an analysis of the water molecules immobilized at the protein-protein interfaces of 115 homodimeric proteins and 46 protein-protein complexes, and compare them with 173 large crystal packing interfaces representing nonspecific interactions. With an average of 15 waters per 1000 A2 of interface area, the crystal packing interfaces are more hydrated than the specific interfaces of homodimers and complexes, which have 10-11 waters per 1000 A2, reflecting the more hydrophilic composition of crystal packing interfaces. Very different patterns of hydration are observed: Water molecules may form a ring around interfaces that remain "dry," or they may permeate "wet" interfaces. A majority of the specific interfaces are dry and most of the crystal packing interfaces are wet, but counterexamples exist in both categories. Water molecules at interfaces form hydrogen bonds with protein groups, with a preference for the main-chain carbonyl and the charged side-chains of Glu, Asp, and Arg. These interactions are essentially the same in specific and nonspecific interfaces, and very similar to those observed elsewhere on the protein surface. Water-mediated polar interactions are as abundant at the interfaces as direct protein-protein hydrogen bonds, and they may contribute to the stability of the assembly.  相似文献   

9.
10.
Addition of millimolar concentrations of Co2+ to Escherichia coli glutamine synthetase induces aggregation along the 6-fold symmetry axes of the protein molecules, forming long strands. The strands subsequently aggregate laterally to form two types of helical cables, a large cable with six outer strands wrapped around a central strand (6+1-stranded cables) and a smaller cable in which three strands wrap around one another. Similar but less extensive aggregation is induced by other divalent metal cations: Cu2+, Ni2+ and Zn2+. The aggregates exhibit little enzymatic activity, and aggregation is completely reversible upon removal of Co2+ in the presence of millimolar Mn2+, including regeneration of nearly full enzyme activity.Each type of helical cable exists in a variety of related forms, which vary in their helical pitch: 6+1-stranded cables have 6-fold axial symmetry, and different specimens are observed with helical pitches from 320 to 540 nm; 3-stranded cables apparently do not have 3-fold axial symmetry and have pitches from 140 to 270 nm. The large variation in pitch for glutamine synthetase helical cables implies either a variation of the regions of intermolecular contacts of approximately 4–10 Å, or a movement of the bonding domains relative to the rest of the molecule by a similar amount.  相似文献   

11.
Bothropstoxin I (BthTX-I) from the venom of Bothrops jararacussuis a myotoxic phospholipase A2 (PLA2) homologue which, although catalytically inactive due to an Asp49→Lys substitution, disrupts the integrity of lipid membranes by a Ca2+-independent mechanism. The crystal structures of two dimeric forms of BthTX-I which diffract X-rays to resolutions of 3.1 and 2.1 Å have been determined. The monomers in both structures are related by an almost perfect twofold axis of rotation and the dimer interfaces are defined by contacts between the N-terminal α-helical regions and the tips of the β-wings of partner monomers. Significant differences in the relative orientation of the monomers in the two crystal forms results in “open” and “closed” dimer conformations. Spectroscopic investigations of BthTX-I in solution have correlated these conformational differences with changes in the intrinsic fluorescence emission of the single tryptophan residues located at the dimer interface. The possible relevance of this structural transition in the Ca2+-independent membrane damaging activity is discussed. Proteins 30:442–454, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

12.
Hugo Schweke  Qifang Xu  Gerardo Tauriello  Lorenzo Pantolini  Torsten Schwede  Frédéric Cazals  Alix Lhéritier  Juan Fernandez-Recio  Luis Angel Rodríguez-Lumbreras  Ora Schueler-Furman  Julia K. Varga  Brian Jiménez-García  Manon F. Réau  Alexandre M. J. J. Bonvin  Castrense Savojardo  Pier-Luigi Martelli  Rita Casadio  Jérôme Tubiana  Haim J. Wolfson  Romina Oliva  Didier Barradas-Bautista  Tiziana Ricciardelli  Luigi Cavallo  Česlovas Venclovas  Kliment Olechnovič  Raphael Guerois  Jessica Andreani  Juliette Martin  Xiao Wang  Genki Terashi  Daipayan Sarkar  Charles Christoffer  Tunde Aderinwale  Jacob Verburgt  Daisuke Kihara  Anthony Marchand  Bruno E. Correia  Rui Duan  Liming Qiu  Xianjin Xu  Shuang Zhang  Xiaoqin Zou  Sucharita Dey  Roland L. Dunbrack  Emmanuel D. Levy  Shoshana J. Wodak 《Proteomics》2023,23(17):2200323
Reliably scoring and ranking candidate models of protein complexes and assigning their oligomeric state from the structure of the crystal lattice represent outstanding challenges. A community-wide effort was launched to tackle these challenges. The latest resources on protein complexes and interfaces were exploited to derive a benchmark dataset consisting of 1677 homodimer protein crystal structures, including a balanced mix of physiological and non-physiological complexes. The non-physiological complexes in the benchmark were selected to bury a similar or larger interface area than their physiological counterparts, making it more difficult for scoring functions to differentiate between them. Next, 252 functions for scoring protein-protein interfaces previously developed by 13 groups were collected and evaluated for their ability to discriminate between physiological and non-physiological complexes. A simple consensus score generated using the best performing score of each of the 13 groups, and a cross-validated Random Forest (RF) classifier were created. Both approaches showed excellent performance, with an area under the Receiver Operating Characteristic (ROC) curve of 0.93 and 0.94, respectively, outperforming individual scores developed by different groups. Additionally, AlphaFold2 engines recalled the physiological dimers with significantly higher accuracy than the non-physiological set, lending support to the reliability of our benchmark dataset annotations. Optimizing the combined power of interface scoring functions and evaluating it on challenging benchmark datasets appears to be a promising strategy.  相似文献   

13.
Quinolinic acid (QA), a biologically potent but neurodestructive metabolite is catabolized by quinolinic acid phosphoribosyltransferase (QPRT) in the first step of the de novo NAD+ biosynthesis pathway. This puts QPRT at the junction of two different pathways, that is, de novo NAD+ biosynthesis and the kynurenine pathway of tryptophan degradation. Thus, QPRT is an important enzyme in terms of its biological impact and its potential as a therapeutic target. Here, we report the crystal structure of human QPRT bound to its inhibitor phthalic acid (PHT) and kinetic analysis of PHT inhibition of human QPRT. This structure, determined at 2.55 Å resolution, shows an elaborate hydrogen bonding network that helps in recognition of PHT and consequently its substrate QA. In addition to this hydrogen bonding network, we observe extensive van der Waals contacts with the PHT ring that might be important for correctly orientating the substrate QA during catalysis. Moreover, our crystal form allows us to observe an intact hexamer in both the apo‐ and PHT‐bound forms in the same crystal system, which provides a direct comparison of unique subunit interfaces formed in hexameric human QPRT. We call these interfaces “nondimeric interfaces” to distinguish them from the typical dimeric interfaces observed in all QPRTs. We observe significant changes in the nondimeric interfaces in the QPRT hexamer upon binding PHT. Thus, the new structural and functional features of this enzyme we describe here will aid in understanding the function of hexameric QPRTs, which includes all eukaryotic and select prokaryotic QPRTs. Proteins 2014; 82:405–414. © 2013 Wiley Periodicals, Inc.  相似文献   

14.
Structure and refinement of penicillopepsin at 1.8 A resolution   总被引:15,自引:0,他引:15  
Penicillopepsin, the aspartyl protease from the mould Penicillium janthinellum, has had its molecular structure refined by a restrained-parameter least-squares procedure at 1.8 Å resolution to a conventional R-factor of 0.136. The estimated co-ordinate accuracy for the majority of the 2363 atoms of the enzyme is better than 0.12 Å. The average atomic thermal vibration parameter, B, for the atoms of the enzyme is 14.5 Å2. One determining factor of this low average B value is the large central hydrophobic core, in which there are two prominent clusters of aromatic residues, one of nine, the other of seven residues. The N and C-terminal domains of penicillopepsin display an approximate 2-fold symmetry: 70 residue pairs are topologically equivalent, related by a rotation of 177 ° and a translation of 1.2 Å. The analysis of the secondary structural features of the molecule reveals non-linear hydrogen bonding. In penicillopepsin, there is no difference in the mean hydrogen-bond parameters for the elements of α-helix, parallel or antiparallel β-pleated sheet. The mean values for these structural elements are: NO, 2.90 Å; NHO, 1.95 Å; N?O, 160 °. The average hydrogen-bond parameters of the reverse β-turns and the 310 helices are distinctly different from the above values. The analysis of sidechain conformational angles χ1 and χ2 penicillopepsin and other enzyme structures refined in this laboratory shows much narrower distributions as compared with those compiled from unrefined protein structures. The close proximity of the carboxyl groups of Asp33 and Asp213 suggests that they share a proton in a tight hydrogen-bonded environment (Asp33OD2 to Asp213OD1 is 2.87 Å). There are several solvent molecules in the active site region and, in particular, O39 forms hydrogen-bonded interactions with both aspartate residues. The disposition of the two carboxyl groups suggests that neither is likely to be involved in a direct nucleophilic attack on the scissile bond of a substrate. The average atomic B-factors of the residues in this region of the molecule are between 5 and 8 Å2, confirming the proposal that conformational mobility of the active site residues has no role in the enzymatic mechanism. However, conformational mobility of neighbouring regions of the molecule e.g. the “flap” containing Tyr75, is verified by the high B-factors for those residues. The positions of 319 solvent sites per asymmetric unit have been selected from difference electron density maps and refined. Thirteen have been classified as internal, and several of these may have key roles during catalysis. The positively charged Nζ atom of Lys304 forms hydrogen bonds to the carboxylate of Asp14 (internal ion pair) and to two internal water molecules O5 and O25. The protonated side-chain of Asp300 forms a hydrogen bond to Thr214O, 2.78 Å, and is the recipient of a hydrogen bond from a surface pocket water molecule O46. There is no possibility for direct interaction between Asp300 and Lys304 without large conformational changes of their environment. The intermolecular packing involves many protein-protein contacts (66 residues) with a large number of solvent molecules involved in bridging between polar residues at the contact surface. The penicillopepsin molecules resemble an approximate hexagonal close-packing of spheres with each molecule having 12 “nearest” neighbours.  相似文献   

15.
Lactoferrin (LF) is a multifunctional acute-phase protein involved in nonspecific defense against bacteria, viruses, and cancer diseases and is present in human barrier fluids, blood, and milk. Small-angle X-ray scattering (SAXS) and light scattering (LS) demonstrated for the first time that LF occurs in the form of oligomers, with a high monomer unit number in the solution. The degree of LF oligomerization depends on the LF concentration and the storage period of non-frozen neutral LF solutions. The average inertial radius of scattering particles (R g) reaches 100–450 Å at LF concentrations comparable with those in human milk, while R g of LF monomers is 26.7 Å. LF forms complexes with various nucleotides and hydrolyzes them. The addition of ATP or AMP to LF solutions accelerates LF oligomerization and increases R g to 600–700 Å, regardless of the initial degree of LF oligomerization. According to the different models (sphere, plate, and cylinder) of LF aggregates, its complexes with such R g presumably contain several tens to thousands of LF monomers. The possible role of oligomeric complexes in multiple biological functions of LF is discussed.  相似文献   

16.
The ClpP component Clp protease fromEscherichia colihas been crystallized and examined by X-ray crystallography and self-rotation function calculations. The crystal belongs to the monoclinic space groupP21with unit cell dimensions ofa=196.9 Å,b=104.3 Å,c=162.4 Å and β=98.3°. The X-ray diffraction pattern extends at least to 2.5 Å Bragg spacing when exposed to CuKα X-rays. Self-rotation function analyses indicate that the ClpP oligomer has 72-point group symmetry. This symmetry suggests that the ClpP oligomer is a tetradecamer, (ClpP)14, consisting of two heptamers, (ClpP)7stacked on top of each other in a head-to-head fashion. The measurement of crystal density indicates that two independent copies of the ClpP oligomers are present in the asymmetric unit, giving a crystal volume per protein mass (VM) of 2.73 Å3/Da and a solvent content of 54.9% (v/v). Self-rotation function calculations are consistent with the presence of two ClpP tetradecamers in the asymmetric unit. The Patterson function suggests that a translation ofx=0.5 andy=0.5 relates a pair of ClpP oligomers in one asymmetric unit to another pair in the other asymmetric unit. And the two independent tetradecamers in one asymmetric unit are related by a relative rotation of about 18° around the 7-fold axis.  相似文献   

17.
The environment of amino acid residues in protein tertiary structures and three types of interfaces formed by protein-protein association--in complexes, homodimers, and crystal lattices of monomeric proteins--has been analyzed in terms of the propensity values of the 20 amino acid residues to be in contact with a given residue. On the basis of the similarity of the environment, twenty residues can be divided into nine classes, which may correspond to a set of reduced amino acid alphabet. There is no appreciable change in the environment in going from the tertiary structure to the interface, those participating in the crystal contacts showing the maximum deviation. Contacts between identical residues are very prominent in homodimers and crystal dimers and arise due to 2-fold related association of residues lining the axis of rotation. These two types of interfaces, representing specific and nonspecific associations, are characterized by the types of residues that partake in "self-contacts"--most notably Leu in the former and Glu in the latter. The relative preference of residues to be involved in "self-contacts" can be used to develop a scoring function to identify homodimeric proteins from crystal structures. Thirty-four percent of such residues are fully conserved among homologous proteins in the homodimer dataset, as opposed to only 20% in crystal dimers. Results point to Leu being the stickiest of all amino acid residues, hence its widespread use in motifs, such as leucine zippers.  相似文献   

18.
《Biophysical journal》2022,121(7):1289-1298
Get3/4/5 chaperone complex is responsible for targeting C-terminal tail-anchored membrane proteins to the endoplasmic reticulum. Despite the availability of several crystal structures of independent proteins and partial structures of subcomplexes, different models of oligomeric states and structural organization have been proposed for the protein complexes involved. Here, using native mass spectrometry (Native-MS), coupled with intact dissociation, we show that Get4/5 exclusively forms a tetramer using both Get5/5 and a novel Get4/4 dimerization interface. Addition of Get3 to this leads to a hexameric (Get3)2-(Get4)2-(Get5)2 complex with closed-ring cyclic architecture. We further validate our claims through molecular modeling and mutational abrogation of the proposed interfaces. Native-MS has become a principal tool to determine the state of oligomeric organization of proteins. The work demonstrates that for multiprotein complexes, native-MS, coupled with molecular modeling and mutational perturbation, can provide an alternative route to render a detailed view of both the oligomeric states as well as the molecular interfaces involved. This is especially useful for large multiprotein complexes with large unstructured domains that make it recalcitrant to conventional structure determination approaches.  相似文献   

19.
The geometry of interactions of planar residues is nonrandom in protein tertiary structures and gives rise to conventional, as well as nonconventional (X--H...pi, X--H...O, where X = C, N, or O) hydrogen bonds. Whether a similar geometry is maintained when the interaction is across the protein-protein interface is addressed here. The relative geometries of interactions involving planar residues, and the percentage of contacts giving rise to different types of hydrogen bonds are quite similar in protein structures and the biological interfaces formed by protein chains in homodimers and protein-protein heterocomplexes--thus pointing to the similarity of chemical interactions that occurs during protein folding and binding. However, the percentage is considerably smaller in the nonspecific and nonphysiological interfaces that are formed in crystal lattices of monomeric proteins. The C--H...O interaction linking the aromatic and the peptide groups is quite common in protein structures as well as the three types of interfaces. However, as the interfaces formed by crystal contacts are depleted in aromatic residues, the weaker hydrogen bond interactions would contribute less toward their stability.  相似文献   

20.
Crystal packing in six crystal forms of pancreatic ribonuclease.   总被引:7,自引:0,他引:7  
We compare the molecular packing of bovine pancreatic ribonuclease A (RNase A) in six crystal forms, two grown with alcohol, three with high salt and one with polyethylene glycol as a precipitant. The six packings differ in the number of molecules in contact and in the extent of the contacts, which bury 1570 A2 to 2790 A2 of the RNase surface. Regions of the protein surface involved in the six packings cover almost the whole RNase molecule. The abundance of polar interactions, about one per 200 A2, is the same in all types of precipitants. All molecule-to-molecule contacts are different in the six crystal forms, except for the one that forms a RNase dimer. The dimer has a large interface covering 1800 A2 and eight to ten polar interactions. Its presence in the three salt-grown crystal forms suggests that it is an intermediate in salt induced crystallization. In contrast, the two alcohol-grown forms contain only small interfaces, implying a different mechanism of nucleation.  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