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.
Lu J  Nie K  Wang F  Tan T 《Bioresource technology》2008,99(14):6070-6074
The immobilized lipase Candida sp. 99-125 catalyzed methanolysis of glycerol trioleate was studied in twelve different solvents in order to deduce the solvent effect through an attempt to correlate the highest yield with such solvent properties as hydrophobicity (log P), dielectric constant (epsilon), and Hildebrand solubility parameter (delta). The results showed that the conversion of glycerol trioleate and yield of oleic acid methyl ester were quite dependent on the solvent. The catalyst lipase in various solvents also needed different optimum amount of water to keep its maximum activity, and generally this lipase in more hydrophobic solvents required more water. The correlation between the highest yield and log P value was found to be reasonable except deviation of data points of certain solvents, while no obvious correlation existed between the other two parameters, dielectric constant (epsilon) and Hildebrand solubility parameter (delta), and the enzyme activity. The study revealed that more hydrophobic solvents such as n-hexane or cyclohexane were more suitable solvents for Candida sp. 99-125 catalyzed transesterification of glycerol trioleate to oleic acid methyl ester.  相似文献   

2.
Forces between type I collagen triple helices are studied in solvents of varying hydrogen-bonding ability. The swelling of collagen fibers in reconstituted films is controlled by the concentration of soluble polymers that are excluded from the fibers and that compete osmotically with collagen for available solvent. The interaxial spacing between the triple helices as a function of the polymer concentration is measured by x-ray diffraction. Exponential-like changes in the spacing with increasing osmotic stress, qualitatively similar to the forces previously found in aqueous solution, are also seen in formamide and ethylene glycol. These are solvents that, like water, are capable of forming three-dimensional hydrogen-bond networks. In solvents that either cannot form a network or have a greatly impaired ability to form a hydrogen-bonded network, strikingly different behavior is observed. A hard-wall repulsion is seen with collagen solvated by ethanol, 2-propanol, and N,N-dimethylformamide. The spacing between helices hardly changes with increasing polymer concentration until the stress exceeds some threshold where removal of the solvent becomes energetically favorable. No solvation of collagen is observed in dimethoxyethane. In solvents with an intermediate ability to form hydrogen-bonded networks, methanol, 2-methoxyethanol, or N-methylformamide, the change in spacing with polymer concentration is intermediate between exponential-like and hard-wall. These results provide direct evidence that the exponential repulsion observed between collagen helices at 0-8-A surface separations in water is due to the energetic cost associated with perturbing the hydrogen-bonded network of solvent molecules between the collagen surfaces.  相似文献   

3.
Although the hydrophobic effect is generally considered to be one of the most important forces in stabilizing the folded structure of a globular protein molecule, there is a lack of consensus on the precise magnitude of this effect. The magnitude of the hydrophobic effect is most directly measured by observing the change in stability of a protein molecule when an internal hydrophobic residue is mutated to another of smaller size. Results of such measurements have, however, been confusing because they vary greatly and are generally considerably larger than expected from the transfer free energies of corresponding small molecules. In this article, a thermodynamic argument is presented to show (1) that the variation is mainly due to that in the flexibility of the protein molecule at the site of mutation, (2) that the maximum destabilization occurs when the protein at the site of mutation is rigid, in which case the value of the destabilization is approximately given by the work of cavity formation in water, and (3) that the transfer free energy approximately gives the minimum of the range of variations. The best numerical agreements between the small molecule and the protein systems are obtained when the data from the small molecule system are expressed as the molarity-based standard free energies without other corrections.  相似文献   

4.
Synthesis of N-CBZ-(N-Carbobenzoxy)-1-amino-acid methyl esters from N-CBZ-amino acids and methanol has been used as an assay to examine the properties of papain in organic solvents containing small amounts of water. Papain is active in solvents ranging in polarity from acetonitrile to tetrachloromethane. The optimal activity in each solvent varied only about three to four fold, but the amount of added water required to achieve it varied from 4% (v/v) in acetonitrile to 0.05% (v/v) in tetrachloromethane. The enzyme was generally more stable in hydrophobic solvents and at lower water contents. The apparent K(m) value of CBZ-glycine was 26 times higher in acetonitrile than in toluene due to differential partitioning of the substrate between aqueous and organic phases. The substrate specificity of the enzyme was qualitatively little different from that in aqueous solution, with amino acid derivatives still the best substrates. Nitrile analogs of substrates inhibited the enzyme, as they do in aqueous solution, and inhibition by a variety of substituted aromatic hydrocarbons showed that the main specificity of papain for hydrophobic side chains at its S(2) subsite, was little affected. The results show that papain can catalyze reactions under a variety of conditions in organic solvents but its substrate specificity is little changed from that in aqueous media.  相似文献   

5.
The effect of organic solvent on the structure and dynamics of proteins was investigated by multiple molecular dynamics simulations (1 ns each) of Candida rugosa lipase in water and in carbon tetrachloride. The choice of solvent had only a minor structural effect. For both solvents the open and the closed conformation of the lipase were near to their experimental X-ray structures (C rms deviation 1–1.3 Å). However, the solvents had a highly specific effect on the flexibility of solvent-exposed side chains: polar side chains were more flexible in water, but less flexible in organic solvent. In contrast, hydrophobic residues were more flexible in organic solvent, but less flexible in water. As a major effect solvent changed the dynamics of the lid, a mobile element involved in activation of the lipase, which fluctuated as a rigid body about its average position. While in water the deviations were about 1.6 Å, organic solvent reduced flexibility to 0.9 Å. This increase rigidity was caused by two salt bridges (Lys85–Asp284, Lys75–Asp79) and a stable hydrogen bond (Lys75–Asn 292) in organic solvent. Thus, organic solvents stabilize the lid but render the side chains in the hydrophobic substrate-binding site more mobile. Figure Superimposition of open (black, PDB entry 1CRL) and closed (gray, PDB entry 1TRH) conformers of C. rugosa lipase. The mobile lid is indicatedThis revised version was published online in October 2004 with corrections to the Graphical Abstract.  相似文献   

6.
The role of hemoglobin solutions as oxygen carriers in biotechnology are numerous, such as in the oxygen supply to biocatalysts or in the preparation of blood substitutes. However, the major barrier to the successful use of hemoglobin in biological and medical engineering is the autoxidation of heme iron during preparation, storage, and utilization. Fifty-six solvents, chosen among the group of Parker's classification, were studied with regard to the autoxidation kinetics of oxyhemoglobin under nondenaturant conditions. Among these solvents 27 present a concentration range in which the autoxidation rates were reduced compared to autoxidation in water. Three groups of solvent have been observed: one exhibiting only a destabilizing effect regardless of the solvent proportion, a second showing a strong stabilizing effect (k(H2O)/k(solvent) greater than 20) and a third showing a low stabilization (k(H2O)/k(solvent) less than 20). The most effective stabilizing solvents were glycerol, glycols, and alcohols. The effect of hydroorganic solvents could be explained by taking into account the globin solvation by water molecules. The solvents that enhance the structure of the water and form few hydrophobic interactions with globin prevent oxyhemoglobin autoxidation.  相似文献   

7.
I L Karle 《Biopolymers》1989,28(1):1-14
Preferred conformation and types of molecular folding are some of the topics that can be addressed by structure analysis using x-ray diffraction of single crystals. The conformations of small linear peptide molecules with 2-6 residues are affected by polarity of solvent, presence of water molecules, hydrogen bonding with neighboring molecules, and other packing forces. Larger peptides, both cyclic and linear, have many intramolecular hydrogen bonds, the effect of which outweighs any intermolecular attractions. Numerous polymorphs of decapeptides grown from a variety of solvents, with different cocrystallized solvents, show a constant conformation for each peptide. Large conformational changes occur, however, upon complexation with metal ions. A new form of free valinomycin grown from DMSO exhibits near three-fold symmetry with only three intramolecular hydrogen bonds. The peptide is in the form of a shallow bowl with a hydrophobic exterior. Near the bottom of the interior of the bowl are three carbonyl oxygens, spaced and directed so that they are in position to form three ligands to a K+, e.g., complexation can be completed by the three lobes containing the beta-bends closing over and encapsulating the K+ ion. In another example, free antamanide and the biologically inactive perhydro analogue, in which four phenyl groups become cyclic hexyl groups, have essentially the same folding of backbone and side chains. The conformation changes drastically upon complexation with Li+ or Na+. However, the metal ion complex of natural antamanide has a hydrophobic globlar form whereas the metal ion complex of the inactive perhydro analogue has a polar band around the middle. The structure results indicate that the antamanide molecule is in a complexed form during its biological activity. Single crystal x-ray diffraction structure analyses have identified the manner in which water molecules are essential to creating minipolar areas on apolar helices. Completely apolar peptides, such as membrane-active peptides, can acquire amphiphilic character by insertion of a water molecule into the helical backbone of Boc-Aib-Ala-Leu-Aib-Ala-Leu-Aib-Ala-Leu-Aib-OMe, for example. The C-terminal half assumes an alpha-helix conformation, whereas the N-terminal half is distorted by an insertion of a water molecule W(1) between N(Ala5) and O(Ala2), forming hydrogen bonds N(5)H...W(1) and W(1)...O(2). The distortion of the helix exposes C = O(Aib1) and C = O(Aib4) to the outside environment with the consequence of attracting additional water molecules. The leucyl side chains are on the other side of the molecule. Thus a helix with an apolar sequence can mimic an amphiphilic helix.  相似文献   

8.
Most enzymes provide their optimum performance at a given water activity (aw), which is generally solvent independent. For a given organic liquid solvent at a specific temperature or for a supercritical solvent at a specific temperature and pressure this corresponds to a water concentration in which water has the desired activity. We present here a methodology for predicting this water concentration thus reducing substantially the amount of experimental work needed to find the optimum solvent with respect to equilibrium conversion.

If the enzyme optimum water activity is known, the methodology predicts the required water content in the solvent to achieve this aw value. If, in addition, the enzyme water activity curve is available, this methodology provides the total water that must be added to the system (enzyme plus solvent) so that a specific water activity can be obtained.

The same methodology can also be applied to predict the effect of the total water content of the system (initial or initial plus produced) on the water activity values. It is shown that: (a) for esterification reactions taking place in hydrophobic organic solvents, the produced water can lead to a substantial change in water activity, but not for less hydrophobic solvents; (b) introduction of dry CO2 into a system, pre-equilibrated to a certain water activity at atmospheric pressure, can lead to a substantial decrease in the water activity especially at temperatures just above the critical one of the solvent and pressures larger than that.  相似文献   

9.
Hydrophobicity of the peptide C=O...H-N hydrogen-bonded group   总被引:1,自引:0,他引:1  
The hydrophobicity of the peptide C=O ... H-N hydrogen-bonded group is an important parameter that determines the structure of proteins in water and in biological membranes, and therefore the free energy of transferring this group from water to non-polar solvents should be determined accurately. The essential work on this problem was carried out by Klotz and co-workers, and has been summarized elsewhere. Using N-methylacetamide as a model peptide, the free energies of the following processes were determined; (1) formation of the C=O ... H-N bond in water, (2) formation of the C=O ... N-N bond in CCl4, and (3) transfer of N-methylacetamide from water to CCl4. (4) From (3), the free energy of transferring the non-hydrogen bonded (C=O, H-N) group from water to CCl4 was calculated. When the free energies of (1), (2) and (4) are combined, one finds that the free energy of transferring the C=O ... H-N group from water to CCl4 is a surprising -1.4 kcal/mol (1 cal = 4.184 J). This number does not seem reasonable, since it implies that the C=O ... H-N group is about as hydrophobic as an isopropyl group, i.e. the side-chain of valine. In the present report, it is shown that this apparent hydrophobicity results from an underestimation of the free energy contribution that the methyl groups make to the transfer of N-methylacetamide from water to CCl4. When appropriate methyl group transfer free energies are used, one finds that the free energy of transferring the C=O ... H-N group from water to CCl4 is +0.62 kcal/mol. Therefore, this group is relatively insensitive to solvent polarity. A similar calculation shows that the free energy of transferring the C=O ... H-O hydrogen-bonded group from water to benzene is +0.55 kcal/mol.  相似文献   

10.
The solvent effect on the equilibrium position of the transesterification reaction of hexanol with ethyl acetate catalyzed by a lipase has been investigated in a variety of non-polar and polar solvents - and binary mixtures. The results obtained indicate that the solvent effect on the equilibrium conversion is very small as compared to that for the direct esterification reactions.

Equilibrium conversions were then predicted using the equilibrium constant for the reaction obtained from Gibbs free energy of formation information for reactants and products in combination with the UNIFAC activity coefficient model. A solvent independent equilibrium conversion was obtained, which was in good agreement with the observed average value for all solvents. This indicates that UNIFAC provides satisfactory estimates of the activity coefficients but its group contribution structure does not allow the prediction of the small differences in conversion among the solvents examined.

Finally plots of these conversions versus the solvent octanol/water partition coefficient or the solubility of water in the solvent, that provide the correct trend in direct esterification reactions, did not achieve the same for transesterification.  相似文献   

11.
For insight into the solvent structure around protein molecules and its role in phase transformations, we investigate the thermodynamics of crystallization of the rhombohedral form of porcine insulin crystals. We determine the temperature dependence of the solubility at varying concentration of the co-solvent acetone, Cac=0%, 5%, 10%, 15%, and 20%, and find that, as a rule, the solubility of insulin increases as temperature increases. The enthalpy of crystallization, undergoes a stepwise shift from approximately -20 kJ mol(-1) at Cac=0%, 5%, and 10% to approximately -55 kJ mol(-1) at Cac=15% and 20%. The entropy change upon crystallization is approximately 35 J mol(-1) K(-1) for the first three acetone concentrations, and drops to approximately -110 J mol(-1) K(-1) at Cac=15% and 20%. DeltaS degrees cryst>0 indicates release of solvent, mostly water, molecules structured around the hydrophobic patches on the insulin molecules' surface in the solution. As Cac increases to 15% and above, unstructured acetone molecules apparently displace the waters and their contribution to DeltaS degrees cryst is minimal. This shifts DeltaS degrees cryst to a negative value close to the value expected for tying up of one insulin molecule from the solution. The accompanying increase in DeltaH degrees cryst suggests that the water structured around the hydrophobic surface moieties has a minimal enthalpy effect, likely due to the small size of these moieties. These findings provide values of the parameters needed to better control insulin crystallization, elucidate the role of organic additives in the crystallization of proteins, and help us to understand the thermodynamics of the hydrophobicity of protein molecules and other large molecules.  相似文献   

12.
Large changes in heat capacity (deltaCp) have long been regarded as the characteristic thermodynamic signature of hydrophobic interactions. However, similar effects arise quite generally in order-disorder transitions in homogeneous systems, particularly those comprising hydrogen-bonded networks, and this may have significance for our understanding of protein folding and other biomolecular processes. The positive deltaCp associated with unfolding of globular proteins in water, thought to be due to hydrophobic interactions, is also typical of the values found for the melting of crystalline solids, where the effect is greatest for the melting of polar compounds, including pure water. This suggests an alternative model of protein folding based on the thermodynamics of phase transitions in hydrogen-bonded networks. Folded proteins may be viewed as islands of cooperatively-ordered hydrogen-bonded structure, floating in an aqueous network of less-well-ordered H-bonds in which the degree of hydrogen bonding decreases with increasing temperature. The enthalpy of melting of the protein consequently increases with temperature. A simple algebraic model, based on the overall number of protein and solvent hydrogen bonds in folded and unfolded states, shows how deltaCp from this source could match the hydrophobic contribution. This confirms the growing view that the thermodynamics of protein folding, and other interactions in aqueous systems, are best described in terms of a mixture of polar and non-polar effects in which no one contribution is necessarily dominant.  相似文献   

13.
The formation of planar bilayer membranes from lipid monolayers as described by Montal and Mueller (Proc. Natl. Acad. Sci. 1972. 69:3561) is analyzed. Bilayers absolutely free of alkane solvents or other nonpolar hydrocarbons can be formed on polytetrafluoroethylene (PTFE) (e.g. Teflon) septa only if certain boundary conditions are satisfied. Measurements have been made of the contact angles between monolayer-coated water and PTFE in the presence and absence of alkane solvents. The measurement suggest that the boundary conditions for formation of stable bilayers can be satisfied only when a nonpolar solvent is present. We conclude that the bilayer must be surrounded by a torus of alkane solvent, petroleum jelly, or silicone grease depending upon the details of technique used to form the bilayer. The non-polar solvent used in the formation of the bilayer may or may not be present in the bilayer depending upon the water solubility and size of the solvent molecule relative to the size of the alkyl chain of the lipid. Detailed sketches describing the formation of bilayers from monolayers are presented.  相似文献   

14.
The stability of biocatalysis in systems containing organic solvents is reviewed. Among the examples presented are homogeneous mixtures of water and water-miscible organic solvents, aqueous/organic two-phase systems, solid biocatalysts suspended in organic solvents, enzymes in reverse micelles and modified enzymes soluble in water immiscible solvents. The stability of biocatalysts in organic solvents depends very much on the conditions. The hydrophobicity or the polarity of the solvent is clearly of great importance. More hydrophobic solvents (higher log P values) are less harmful to enzymes than less hydrophobic solvents. The water content of the system is a very important parameter. Some water is essential for enzymatic activity; however, the stability of enzymes decreases with increasing water content. Mechanisms of enzyme inactivation are discussed.  相似文献   

15.
Hydrophobic interactions are essential for stabilizing protein-protein complexes, whose interfaces generally consist of a central cluster of hot spot residues surrounded by less important peripheral residues. According to the O-ring hypothesis, a condition for high affinity binding is solvent exclusion from interacting residues. This hypothesis predicts that the hydrophobicity at the center is significantly greater than at the periphery, which we estimated at 21 cal mol(-1) A(-2). To measure the hydrophobicity at the center, structures of an antigen-antibody complex where a buried phenylalanine was replaced by smaller hydrophobic residues were determined. By correlating structural changes with binding free energies, we estimate the hydrophobicity at this central site to be 46 cal mol(-1) A(-2), twice that at the periphery. This context dependence of the hydrophobic effect explains the clustering of hot spots at interface centers and has implications for hot spot prediction and the design of small molecule inhibitors.  相似文献   

16.
DNA undergoes a helix-to-coil transition (also called denaturation transition) upon heating. This transition can also be facilitated by using solvent mixtures (for example water–alcohol). An increase in the hydrophobic tail of the second solvent molecule first decreases then increases the melting temperature appreciably. Measurement on 4% DNA in a series of water–alcohol mixtures shows that the helix-to-coil melting transition is driven by the solvent ability to cross the hydrophobic sugar-rich region. DNA is behaving like a cylindrical micelle.  相似文献   

17.
The conformation of microtubule-bound paclitaxel has been examined by fluorescence and solid-state NMR spectroscopy. A fluorescent derivative of paclitaxel, 3'-N-debenzoyl-3'-N-(m-aminobenzoyl)paclitaxel (N-AB-PT), was prepared by semisynthesis. No differences in the microtubule-promoting activity between N-AB-PT and paclitaxel were observed, demonstrating that addition of the amino group did not adversely affect the ligand-receptor association. The distance between the fluorophore N-AB-PT and the colchicine binding site on tubulin polymers was determined through time-resolved measurements of fluorescence resonance energy transfer to be 29 +/- 2 A. The absorption and emission spectra of N-AB-PT bound to microtubules and in various solvents were measured. A plot of the Stokes shift as a function of solvent polarity was highly unusual. The Stokes shift increased linearly with solvent polarity in protic solvents, which is expected due to the nature of the fluorophore. In aprotic solvents, however, the Stokes shift was invariant with solvent polarity, indicating that the fluorophore was somehow shielded from the effects of the solvent. These data are best explained by considering the solution-state conformational properties of paclitaxel. It is known that paclitaxel adopts different conformations depending on the nature of the solvent, and these fluorescence data are consistent with the molecule adopting a "hydrophobic collapsed" conformation in protic solvents and an "extended" conformation in aprotic solvents. The Stokes shift of microtubule-bound N-AB-PT was within the protic solvent region, demonstrating that microtubule-bound paclitaxel is in a hydrophobic collapsed conformation. Microtubule-bound paclitaxel was also investigated by solid-state NMR. Paclitaxel was labeled with (19)F at the para position of the C-2 benzoyl substituent and with (13)C and (15)N in the side chain. Distances between the fluorine and carbon nuclei were determined by REDOR. The distance between the fluorine and the 3'-amide carbonyl carbon was 9.8 +/- 0.5 A, and the distance between the fluorine atom and the 3'-methine carbon was 10. 3 +/- 0.5 A. These spectroscopic data were used in conjunction with molecular modeling to refine the microtubule-bound conformation of paclitaxel and to suggest an alternative orientation of the ligand within the paclitaxel binding site.  相似文献   

18.
Osmolytes are small organic solutes accumulated at high concentrations by cells/tissues in response to osmotic stress. Osmolytes increase thermodynamic stability of folded proteins and provide protection against denaturing stresses. The mechanism of osmolyte compatibility and osmolyte-induced stability has, therefore, attracted considerable attention in recent years. However, to our knowledge, no quantitative study of osmolyte effects on the strength of hydrophobic interactions has been reported. Here, we present a detailed molecular dynamics simulation study of the effect of the osmolyte trimethylamine-N-oxide (TMAO) on hydrophobic phenomena at molecular and nanoscopic length scales. Specifically, we investigate the effects of TMAO on the thermodynamics of hydrophobic hydration and interactions of small solutes as well as on the folding-unfolding conformational equilibrium of a hydrophobic polymer in water. The major conclusion of our study is that TMAO has almost no effect either on the thermodynamics of hydration of small nonpolar solutes or on the hydrophobic interactions at the pair and many-body level. We propose that this neutrality of TMAO toward hydrophobic interactions-one of the primary driving forces in protein folding-is at least partially responsible for making TMAO a "compatible" osmolyte. That is, TMAO can be tolerated at high concentrations in organisms without affecting nonspecific hydrophobic effects. Our study implies that protein stabilization by TMAO occurs through other mechanisms, such as unfavorable water-mediated interaction of TMAO with the protein backbone, as suggested by recent experimental studies. We complement the above calculations with analysis of TMAO hydration and changes in water structure in the presence of TMAO molecules. TMAO is an amphiphilic molecule containing both hydrophobic and hydrophilic parts. The precise balance of the effects of hydrophobic and hydrophilic segments of the molecule appears to explain the virtual noneffect of TMAO on the strength of hydrophobic interactions.  相似文献   

19.
The crystal structure analysis of horse liver alcohol dehydrogenase has been extended to 2.4 Å resolution. From the corresponding electron density map of the apoenzyme we have determined the positions of the 374 amino acids in the polypeptide chain of each subunit.The coenzyme binding domain of the subunit comprises residues 176 to 318. 45% of these residues are helical and 32% are in the central six-stranded pleated sheet structure. The positions and orientations of the helices with respect to the pleated sheet indicate a possible folding mechanism for this part of the subunit structure. The coenzyme analogue ADP-ribose binds to this domain in a position and orientation very similar to coenzyme binding to lactate dehydrogenase. The adenine part binds in a hydrophobic pocket, the adenosine ribose is hydrogen-bonded to the side chain of Asp223, the pyrophosphate is positioned by interaction with Arg47 and the nicotinamide ribose is 6Å away from the catalytic zinc atom.The catalytic domain is mainly built up from three distinct antiparallel pleated-sheet regions. Residues within this domain provide ligands to the catalytic zinc atom; Cys46, His67 and Cys174. An approximate tetrahedral coordination of this zinc is completed by a water molecule or hydroxyl ion depending on the pH. Residues 95 to 113 form a lobe that binds the second zinc atom of the subunit. This zinc is liganded in a distorted tetrahedral arrangement by four sulphur atoms from the cysteine residues 97, 100, 103 and 111. The lobe forms one side of a significant cleft in the enzyme surface suggesting that this region might constitute a second catalytic centre of unknown function.The two domains of the subunit are separated by a crevice that contains a wide and deep hydrophobic pocket. The catalytic zinc atom is at the bottom of this pocket, with the zinc-bound water molecule projecting out into the pocket. This water molecule is hydrogen-bonded to the side chain of Ser48 which in turn is hydrogen-bonded to His51. The pocket which in all probability is the binding site for the substrate and the nicotinamide moiety of the coenzyme, is lined almost exclusively with hydrophobic side chains. Both subunits contribute residues to each of the two substrate binding pockets of the molecule. The only accessible polar groups in the vicinity of the catalytic centre are Ser48 and Thr178 apart from zinc and the zinc-bound water molecule.  相似文献   

20.
Aims:  Solvent-tolerant bacteria have emerged as a new class of micro-organisms able to grow at high concentrations of toxic solvents. Such bacteria and their solvent-stable enzymes are perceived to be useful for biotransformations in nonaqueous media. In the present study, the solvent-responsive features of a lipase–producing, solvent-tolerant strain Pseudomonas aeruginosa PseA have been investigated to understand the cellular mechanisms followed under solvent-rich conditions.
Methods and Results:  The solvents, cyclohexane and tetradecane with differing log P -values (3·2 and 7·6 respectively), have been used as model systems. Effect of solvents on (i) the cell morphology and structure (ii) surface hydrophobicity and (iii) permeability of cell membrane have been examined using transmission electron microscopy, atomic force microscopy and other biochemical techniques. The results show that (i) less hydrophobic (low log P -value) solvent cyclohexane alters the cell membrane integrity and (ii) cells adapt to organic solvents by changing morphology, size, permeability and surface hydrophobicity. However, no such changes were observed in the cells grown in tetradecane.
Conclusions:  It may be concluded that P. aeruginosa PseA responds differently to solvents of different hydrophobicities. Bacterial cell membrane is more permeable to less hydrophobic solvents that eventually accumulate in the cytoplasm, while highly hydrophobic solvents have lesser tendency to access the membrane.
Significance and Impact of the Study:  To the best of our knowledge, these are first time observations that show that way of bacterial solvent adaptability depends on nature of solvent. Difference in cellular responses towards solvents of varying log P -values (hydrophobicity) might prove useful to search for a suitable solvent for carrying out whole-cell biocatalysis.  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