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.
The immobilization of the globular protein α‐1‐acid glycoprotein (AGP) onto silica gel led to the commercial availability of an AGP column, which has a high enantioselectivity. The enantioselectivity of AGP columns has been demonstrated in numerous applications. Due to potential AGP structural changes occurring upon its immobilization, the interaction between particular pairs of enantiomers and the stationary phase is very difficult to assess. Therefore, in this paper we report a mechanistic study that probes the nature of these types of interactions. As model ligands, we employed two LTD4 antagonists (L‐708, 738, MK0476, and their enantiomers) which have a rigid backbone consisting of a conjugated aromatic region and a side chain which is terminated with a carboxylic functional group. The difference between the two compounds is a two‐fluorine versus one‐chlorine substituent in the aromatic region of the molecule. To study the interaction between the two homologues and the AGP stationary phase, several parameters were varied, including pH, ionic strength, organic modifier, and temperature. van't Hoff plots were constructed and found to be nonlinear. They could, however, be divided into two linear regions, one from 0°C to ∼30°C, and another from 39°C to 50°C. The region at lower temperature implied that the separation was entropy‐dominated while the separation at higher temperature was enthalpically driven. The transition from the entropic to the enthalpically driven separation region suggested that bound AGP undergoes a conformational change. Fluorescence spectroscopy performed on the AGP stationary phase found evidence for a limited conformational transition at a similar temperature, consistent with this hypothesis. Chirality 11:224–232, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

2.
This study utilizes sensitive, modern isothermal titration calorimetric methods to characterize the microscopic thermodynamic parameters that drive the binding of basic drugs to α‐1‐acid glycoprotein (AGP) and thereby rationalize the thermodynamic data in relation to docking models and crystallographic structures of the drug–AGP complexes. The binding of basic compounds from the tricyclic antidepressant series, together with miaserine, chlorpromazine, disopyramide and cimetidine, all displayed an exothermically driven binding interaction with AGP. The impact of protonation/deprotonation events, ionic strength, temperature and the individual selectivity of the A and F1*S AGP variants on drug‐binding thermodynamics was characterized. A correlation plot of the thermodynamic parameters for all of the test compounds revealed that an enthalpy–entropy compensation is in effect. The exothermic binding energetics of the test compounds were driven by a combination of favorable (negative) enthalpic (?Hº) and favorable (positive) entropic (?Sº) contributions to the Gibbs free energy (?Gº). Collectively, the data imply that the free energies that drive drug binding to AGP and its relationship to drug serum residency evolve from the complex interplay of enthalpic and entropic forces from interactions with explicit combinations of hydrophobic and polar side‐chain sub‐domains within the multi‐lobed AGP ligand binding cavity.Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

3.
The binding of the (R)‐ and (S)‐enantiomers of amlodipine to bovine serum albumin (BSA), human serum albumin (HSA), α1‐acid glycoprotein (AGP), and human plasma (HP) was studied by equilibrium dialysis over the concentration range of 75–200 μM at a protein concentration of 150 μM. Unbound drug concentrations were determined by enantioselective capillary electrophoresis using 50 mM phosphate buffer, pH 2.5, containing 18 mM α‐cyclodextrin as background electrolyte. Saturation of the protein binding sites was not observed over the concentration range tested. Upon application of racemic amlodipine besylate, (S)‐amlodipine was bound to a higher extend by HSA and HP compared with (R)‐amlodipine, whereas the opposite binding of the enantiomers was observed for BSA and AGP. Scatchard analysis was used to illustrate the different binding affinities of amlodipine besylate enantiomers to BSA, HSA and AGP. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
Ligand binding of neutral progesterone, basic propranolol, and acidic warfarin to human α1‐acid glycoprotein (AGP) was investigated by Raman spectroscopy. The binding itself is characterized by a uniform conformational shift in which a tryptophan residue is involved. Slight differences corresponding to different contacts of the individual ligands inside the β‐barrel are described. Results are compared with in silico ligand docking into the available crystal structure of deglycosylated AGP using quantum/molecular mechanics. Calculated binding energies are ?18.2, ?14.5, and ?11.5 kcal/mol for warfarin, propranolol, and progesterone, respectively. These calculations are consistent with Raman difference spectroscopy; nevertheless, minor discrepancies in the precise positions of the ligands point to structural differences between deglycosylated and native AGP. Thermal dynamics of AGP with/without bounded warfarin was followed by Raman spectroscopy in a temperature range of 10–95 °C and analyzed by principal component analysis. With increasing temperature, a slight decrease of α‐helical content is observed that coincides with an increase in β‐sheet content. Above 45 °C, also β‐strands tend to unfold, and the observed decrease in β‐sheet coincides with an increase of β‐turns accompanied by a conformational shift of the nearby disulfide bridge from high‐energy trans‐gauche‐trans to more relaxed gauche‐gauche‐trans. This major rearrangement in the vicinity of the bridge is not only characterized by unfolding of the β‐sheet but also by subsequent ligand release. Hereby, ligand binding alters the protein dynamics, and the more rigid protein–ligand complex shows an improved thermal stability, a finding that contributes to the reported chaperone‐like function of AGP. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
The HPLC separation of the R,S and S,R enantiomers of pyrrolidinyl norephedrine on immobilized alpha-1 glycoprotein (AGP) was investigated. Conditions for the separation were varied using a premixed mobile phase containing an ammonium phosphate buffer and an organic modifier. The influence of mobile phase pH, ionic strength, organic modifier composition, modifier type, and temperature on the chiral selectivity and retention were investigated. The presented data demonstrate that independent phenomena govern the enantioselectivity and retention. Retention is a function of both ion exchange equilibria and hydrophobic adsorption. Thermodynamic data derived from van't Hoff plots illustrates that while enantioselectivity is also enthalpically driven, the magnitude of the enthalpy term is governed by pH. Enantioselectivity has little dependence on ionic strength. Hydrophobic interactions appear to foster hydrogen bonding interactions; the two appear to be mutually responsible for chiral selectivity. The chiral selectivity decreases as the pH is decreased and increases with mobile phase buffer strength.  相似文献   

6.
Phosphofructokinase‐1 from Saccharomyces cerevisiae is composed of two types of subunits, α and β. Subunit‐specific monoclonal antibodies were raised to elucidate structural and functional properties of both subunits. One monoclonal antibody, α‐F3, binds to an epitope either at the C‐terminal or at the N‐terminal part of the α‐polypeptide chain. By screening a heptapeptide library with this monoclonal antibody, a set of heptapeptides was selected, which contained the consensus sequences D–A–F and D–S–F. Two heptapeptides with these motifs were synthesized in order assess their capacity to inhibit the binding of antibody α‐F3 to native phosphofructokinase‐1. The peptide G–I–K–D–A–F–L inhibited the binding more strongly (IC50 = 1.5 µM) than the peptide A–P–W–H–D–S–F (IC50 = 33.3 µM). Sequence matching revealed the presence of the D–A–F motif in the polypeptide chain of phosphofructokinase‐1 at amino acid position 172–174. As a control, the nonapeptide A–P–T–S–K–D–A–F–L which corresponds to the sequence of the putative epitope was tested in the inhibition assay. In view of the high inhibitory capacity (IC50 = 0.3 µM) it was concluded that this nonapeptide represents the continuous epitope of phosphofructokinase‐1 that is recognized by antibody α‐F3. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

7.
Lipase from Pseudomonas cepacia was made soluble in 1,4‐dioxane by lyophilization of the enzyme from aqueous solutions containing methoxypoly(ethylene glycol) (PEG). The solubility of the enzyme–PEG complex depended both on protein concentration and PEG protein ratio. Intrinsic protein fluorescence and far‐ and near‐UV circular dichroism revealed that not only did the enzyme not unfold in the organic solvent, but rather became more compact. This was seen by the slight quenching of fluorescence intensity and by the enhancement of the near‐UV circular dichroism negative signals, which are indicative of stronger interactions of tryptophanyl and/or tyrosyl residues among themselves or with other parts of the enzyme molecule. The specific activity of the lipase–PEG complex in the organic solvent was at least 2 orders of magnitude higher than that of the enzyme powder. This can be attributed both to the maintenance of native conformation and to enzyme dissolution in the reaction medium which should minimize possible limitations to enzyme–substrate interactions. © 1999 John Wiley & Sons, Inc., Biotechnol Bioeng 64: 624–629, 1999.  相似文献   

8.
A PMMA‐binding peptide (PMMA‐tag) was genetically fused with the C‐terminal region of an anti‐human chorionic gonadotropin (hCG) single‐domain antibody (VHH). It was over‐expressed in an insoluble fraction of E. coli cells, and recovered in the presence of 8 M urea via one‐step IMAC purification. Monomeric and denatured PMMA‐tag‐fused VHH (VHH‐PM) was successfully prepared via the reduction and oxidation of VHH‐PM at a concentration less than 1 mg/mL in the presence of 8 M of urea. Furthermore, the VHH‐PM was refolded with a recovery of more than 95% by dialysis against 50 mM TAPS at pH 8.5, because the genetic fusion of PMMA‐tag resulted in a decrease in the apparent isoelectric point (pI) of the fusion protein, and its solubility at weak alkaline pH was considerably increased. The antigen‐binding activities of VHH‐PM in the adsorptive state were 10‐fold higher than that of VHH without a PMMA‐tag. The density of VHH‐PM on a PMMA plate was twice that of VHH, indicating that the site‐directed attachment of a PMMA‐tag resulted in positive effects to the adsorption amount as well as to the orientation of VHH‐PM in its adsorptive state. The preparation and immobilization methods for VHH‐PM against hCG developed in the present study were further applied to VHH‐PMs against four different antigens, and consequently, those antigens with the concentrations lower than 1 ng/mL were detected by the sandwich ELISA. Thus, the VHH‐PMs developed in the present study are useful for preparation of high‐performance and economical immunosorbent for detection of biomarkers. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:1563–1570, 2015  相似文献   

9.
Rapid and simple isocratic high‐performance liquid chromatographic methods with UV detection were developed and validated for the direct resolution of racemic mixtures of hyoscyamine sulfate and zopiclone. The method involved the use of αl‐acid glycoprotein (AGP) as chiral stationary phase. The stereochemical separation factor (?) and the stereochemical resolution factor (Rs) obtained were 1.29 and 1.60 for hyoscyamine sulfate and 1.47 and 2.45 for zopiclone, respectively. The method was used for determination of chiral switching (eutomer) isomers: S‐hyoscyamine sulfate and eszopiclone. Several mobile phase parameters were investigated for controlling enantioselective retention and resolution on the chiral AGP column. The influence of mobile phase, concentration and type of uncharged organic modifier, ionic strength, and column temperature on enantioselectivity were studied. Calibration curves were linear in the ranges of 1–10 µg mL‐1 and 0.5–5 µg mL‐1 for S‐hyoscyamine sulfate and eszopiclone, respectively. The method is specific and sensitive, with lower limits of detection and quantifications of 0.156, 0.515 and 0.106, 0.349 for S‐hyoscyamine sulfate and eszopiclone, respectively. The method was used to identify quantitatively the enantiomers profile of the racemic mixtures of the studied drugs in their pharmaceutical preparations. Thermodynamic studies were performed to calculate the enthalpic ΔH and entropic ΔS terms. The results showed that enantiomer separation of the studied drugs were an enthalpic process. Chirality 28:49–57, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

10.
Human metallothionein 1a, a protein with two cysteine‐rich metal‐binding domains (α with 11 Cys and β with 9), was analyzed in its metal‐free form by selective, covalent Cys modification coupled with ESI‐MS. The modification profiles of the isolated β‐ and α‐fragments reacted with p‐benzoquinone (Bq), N‐ethylmalemide (NEM) and iodoacetamide (IAM) were compared with the full length protein using ESI‐mass spectral data to follow the reaction pathway. Under denaturing conditions at low pH, the reaction profile with each modifier followed pathways that resulted in stochastic, Normal distributions of species whose maxima was equal to the mol. eq. of modifier added. Our interpretation of modification at this pH is that reaction with the cysteines is unimpeded when the full protein or those of its isolated domains are denatured. At neutral pH, where the protein is expected to be folded in a more compact structure, there is a difference in the larger Bq and NEM modification, whose reaction profiles indicate a cooperative pattern. The reaction profile with IAM under native conditions follows a similar stochastic distribution as at low pH, suggesting that this modifier is small enough to access the cysteines unimpeded by the compact structure. The data emphasize the utility of residue modification coupled with electrospray ionization mass spectrometry for the study of protein structure.  相似文献   

11.
A high-performance liquid chromatographic method using a chiral AGP column was developed to screen and determine the enantiomers of bunolol in human urine. The recovery of (+)- and (−)-bunolol from urine was 91.79–95.23% at different concentrations. The coefficients of variation (C.V.) were less than 2.1 and 2.3% for intra- and inter-assays, respectively. Urinary metabolites were detected using GC—MS after derivatization with N-methyl(trimethylsilyl)trifluroacetamide. The influences of pH and modifier on a chiral AGP column were studied.  相似文献   

12.
We investigated water/organic solvent sorption and residual enzyme activity to simultaneously monitor preferential solvation/hydration of protein macromolecules in the entire range of water content at 25°C. We applied this approach to estimate protein destabilization/stabilization due to the preferential interactions of bovine pancreatic α‐chymotrypsin with water‐acetone (moderate‐strength H‐bond acceptor) and water‐DMSO (strong H‐bond acceptor) mixtures. There are three concentration regimes for the dried α‐chymotrypsin. α‐Chymotrypsin is preferentially hydrated at high water content. The residual enzyme activity values are close to 100%. At intermediate water content, the dehydrated α‐chymotrypsin has a higher affinity for acetone/DMSO than for water. Residual enzyme activity is minimal in this concentration range. The acetone/DMSO molecules are preferentially excluded from the protein surface at the lowest water content, resulting in preferential hydration. The residual catalytic activity in the water‐poor acetone is ~80%, compared with that observed after incubation in pure water. This effect is very small for the water‐poor DMSO. Two different schemes are operative for the hydrated enzyme. At high and intermediate water content, α‐chymotrypsin exhibits preferential hydration. However, at intermediate water content, in contrast to the dried enzyme, the initially hydrated α‐chymotrypsin possesses increased preferential hydration parameters. At low water content, no residual enzyme activity was observed. Preferential binding of DMSO/acetone to α‐chymotrypsin was detected. Our data clearly demonstrate that the hydrogen bond accepting ability of organic solvents and the protein hydration level constitute key factors in determining the stability of protein–water–organic solvent systems.  相似文献   

13.
The function of sialic acid groups at the terminal of sugar chains of human α1-acid glycoprotein (AGP) was investigated with respect to chiral discrimination between optical isomers of basic drugs, using high-performance capillary electrophoresis/frontal analysis (HPCE/FA), a novel analytical method developed for the determination of unbound drug concentration with ultramicrosample volume (100–200 nl). Native human AGP and desialylated AGP were used as test proteins, and propranolol (PRO) and verapamil (VER) were used as model drugs. The unbound concentration of (S)-VER was 1.31 times higher than that of (R)-VER in native AGP solution. This selectivity was not affected by desialylation. Further, enzymatic elimination of galactose residues, which neighbored sialic acid groups, did not change the binding of either isomer of VER. On the other hand, the unbound concentration of (R)-PRO was 1.27 times higher than that of (S)-PRO in native AGP solution. Desialylation caused the unbound concentration of (S)-PRO to rise to the same level of (R)-PRO, resulting in loss of enantioselectivity. Thus, it follows that sialic acid groups of AGP, as a whole, are not responsible for chiral recognition between enantiomers of VER but are involved in enantioselectivity toward the isomers of PRO. Chirality 9:291–296, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

14.
A new experimental approach based on FTIR spectroscopic measurements was proposed to study simultaneously the adsorption/desorption of water and organic solvent on solid enzyme and corresponding changes in the enzyme secondary structure in the water activity range from 0 to 1.0 at 25 degrees C. The effect of dioxane on the hydration/dehydration and structure of bovine pancreatic alpha-chymotrypsin (CT) was characterized by means of this approach. Dioxane sorption exhibits pronounced hysteresis. No sorbed dioxane was observed at low water activities (a(w)<0.5) during hydration. At a(w) about 0.5, a sharp increase in the amount of sorbed dioxane was observed. Dioxane sorption isotherm obtained during dehydration resembles a smooth curve. In this case, CT binds about 150 mol dioxane/mol enzyme at the lowest water activities. Three different effects of dioxane on the water binding by the initially dried CT were observed. At a(w)<0.5, water adsorption is similar in the presence and absence of dioxane. It was concluded that the presence of dioxane has little effect on the interaction between enzyme and tightly bound water at low a(w). At a(w)>0.5, dioxane increases the amount of water bound by CT during hydration. This behavior was interpreted as a dioxane-assisted effect on water binding. Upon dehydration at low water activities, dioxane decreases the water content at a given a(w). This behavior suggests that the suppression in the uptake of water during dehydration may be due to a competition for water-binding sites on chymotrypsin by dioxane. Changes in the secondary structure of CT were determined from infrared spectra by analyzing the structure of amide I band. Dioxane induced a strong band at 1628 cm(-1) that was assigned to the intermolecular beta-sheet aggregation. Changes in the intensity of the 1628 cm(-1) band agree well with changes in the dioxane sorption by CT. An explanation of the dioxane effect on the CT hydration and structure was provided on the basis of hypothesis on water-assisted disruption of polar contacts in the solid enzyme. The reported results demonstrate that the hydration and structure of alpha-chymotrypsin depend markedly on how enzyme has been hydrated - whether in the presence or in the absence of organic solvent. A qualitative model was proposed to classify the effect of hydration history on the enzyme activity-a(w) profiles.  相似文献   

15.
A new method for preparing immobilized alpha1-acid glycoprotein (AGP) for use in drug-protein binding studies was developed and optimized. In this approach, periodate was used under mild conditions to oxidize the carbohydrate chains in AGP for attachment to a hydrazide-activated support. The final conditions chosen for this oxidation involved the reaction of 5.0 mg/mL AGP at 4 degrees C and pH 7.0 with 5-20 mM periodic acid for 10 min. These conditions helped maximize the immobilization of AGP without significantly affecting its activity. This method was evaluated by using it to attach AGP to silica for use in high-performance affinity chromatography and self-competition zonal elution studies. In work with R- and S-propranolol, only one type of binding site was observed for both enantiomers on the immobilized AGP, in agreement with previous studies using soluble AGP. The association equilibrium constants measured for the immobilized AGP with R- and S-propranolol at pH 7.4 and 37 degrees C were 2.7 x 10(6) and 4.2 x 10(6) M(-1), respectively, with linear van't Hoff plots being obtained between 5 and 37 degrees C. Work performed with other drugs also gave good agreement between the behavior seen for immobilized AGP and that for soluble AGP. The same immobilization method described in this work could be used to attach AGP to other materials, such as those used for surface plasmon resonance or alternative biosensors.  相似文献   

16.
For most therapeutic proteins, a long serum half‐life is desired. Studies have shown that decreased antigen binding at acidic pH can increase serum half‐life. In this study, we aimed to investigate whether pH‐dependent binding sites can be introduced into antigen binding crystallizable fragments of immunoglobulin G1 (Fcab). The C‐terminal structural loops of an Fcab were engineered for reduced binding to the extracellular domain of human epidermal growth factor receptor 2 (Her2‐ECD) at pH 6 compared to pH 7.4. A yeast‐displayed Fcab‐library was alternately selected for binding at pH 7.4 and non‐binding at pH 6.0. Selected Fcab variants showed clear pH‐dependent binding to soluble Her2‐ECD (decrease in affinity at pH 6.0 compared to pH 7.4) when displayed on yeast. Additionally, some solubly expressed variants exhibited pH‐dependent interactions with Her2‐positive cells whereas their conformational and thermal stability was pH‐independent. Interestingly, two of the three Fcabs did not contain a single histidine mutation but all of them contained variations next to histidines that already occurred in loops of the lead Fcab. The study demonstrates that yeast surface display is a valuable tool for directed evolution of pH‐dependent binding sites in proteins.  相似文献   

17.
Neutral salts activate and stabilize thermolysin. In this study, to explore the mechanism, we analyzed the interaction of 8-anilinonaphthalene 1-sulphonate (ANS) and thermolysin by ANS fluorescence. At pH 7.5, the fluorescence of ANS increased and blue-shifted with increasing concentrations (0–2.0?μM) of thermolysin, indicating that the anilinonaphthalene group of ANS binds with thermolysin through hydrophobic interaction. ANS did not alter thermolysin activity. The dissociation constants (Kd) of the complex between ANS and thermolysin was 33?±?2?μM at 0?M NaCl at pH 7.5, decreased with increasing NaCl concentrations, and reached 9?±?3?μM at 4?M NaCl. The Kd values were not varied (31?34?μM) in a pH range of 5.5?8.5. This suggests that at high NaCl concentrations, Na+ and/or Cl ions bind with thermolysin and affect the binding of ANS with thermolysin. Our results also suggest that the activation and stabilization of thermolysin by NaCl are partially brought about by the binding of Na+ and/or Cl ions with thermolysin.  相似文献   

18.
Some reactions of organic synthesis require to be performed in rather aggressive media, like organic solvents, that frequently impair enzyme operational stability to a considerable extent. We have studied the option of developing a reactivation strategy to increase biocatalyst lifespan under such conditions, under the hypothesis that organic solvent enzyme inactivation is a reversible process. Glyoxyl agarose immobilized penicillin G acylase and cross‐linked enzyme aggregates of the enzyme were considered as biocatalysts performing in dioxane medium. Reactivation strategy consisted in re‐incubation in aqueous medium of the partly inactivated biocatalysts in organic medium, best conditions of reactivation being studied with respect to dioxane concentration and level of enzyme inactivation attained prior to reactivation. Best results were obtained with glyoxyl agarose immobilized penicillin G acylase at all levels of residual activity studied, with reactivations up to 50%; for the case of a biocatalyst inactivated down to 75% of its initial activity, full recovery of enzyme activity was obtained after reactivation. The potential of this strategy was evaluated in the thermodynamically controlled synthesis of deacetoxycephalosporin G in a sequential batch reactor operation, where a 20% increase in the cumulative productivity was obtained by including an intermediate stage of reactivation after 50% inactivation. Biotechnol. Bioeng. 2009;103: 472–479. © 2009 Wiley Periodicals, Inc.  相似文献   

19.
The chiral separation of (±)‐catechin was investigated by capillary electrophoresis using characterized succinyl‐β‐cyclodextrins (Suc‐β‐CDs) with one to three degree of substitution values. The effects of nature and concentration of Suc‐β‐CDs and running buffer pH on the migration time and resolution of (±)‐catechin are discussed. All three kinds of Suc‐β‐CDs show a clear baseline separation of (±)‐catechin in capillary electrophoresis. Mono‐Suc‐β‐CD effectively separated (±)‐catechin, and additional substituted CDs (di‐ and tri‐Suc‐β‐CD) were capable of chiral separation at a broad pH range. The optimum running conditions were found to be 100 mM borate buffer (pH 9.8) containing 5 mM mono‐Suc‐β‐CD with no methanol organic modifier. Chirality, 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
The immunologically important (1 → 6) comb‐like branched (1 → 3)‐β‐D ‐glucans scleroglucan, schizophyllan, lentinan, and others, exist mainly as linear triple‐helical structures in aqueous solution. Partial interconversion from linear to circular topology has been reported to take place following conformational transition of the triple‐helical structure and subsequent regeneration of the triplex conformation. We here report on experimental data indicating that complete strand separation of the triple‐helical structure is required for this interconversion. NaOH or dimethylsulfoxide was used to induce dissociation of the triplex at combinations of concentrations and temperatures shown by calorimetry to yield a conformational transition of the triplex structures. For the alkaline treatment at 55°C, it is found that up to about 30% of the material readily can be converted to the cyclic topology. This fraction increased to about 60% when the subsequent annealing of the scleroglucan in aqueous solution at pH 7 was carried out at 100°C. Further increase of the annealing temperature yielded a smaller relative amount of cyclic species. The data indicate that the lower molecular weight fraction of the molecular weight distributions can be converted selectively to the macrocyclic topology by conditions that do not yield complete strand separation of the whole sample. These findings add to previous reports by providing more details about how the conditions required for the linear triplex to macrocycle interconversion relate to the conformational properties of the triple‐helical structure. © 1999 John Wiley & Sons, Inc. Biopoly 50: 496–512, 1999  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