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 USE OF TETRACHLORONITROBENZENE ISOMERS ON POTATOES   总被引:1,自引:0,他引:1  
In laboratory and field experiments from 1952 to 1955, 2:3:5:6-tetrachloro-I-nitrobenzene (tecnazene) delayed the development of sprouts on treated potato tubers, and 2:3:4:6-TCNB retarded sprout development in two laboratory experiments. At the usual commercial rates of application the 2:3:4:5 isomer had no effect on sprout development and very little effect at higher rates. The crop from 2:3:5:6-TCNB-treated seed tubers usually contained a higher proportion of small tubers than that from untreated or 2:3:4:5-TCNB-treated seed. Marketable yield from 2:3:5:6-TCNB-treated seed was sometimes below that from untreated seed, but when this happened the yield from 2:3:4:5-TCNB-treated seed was higher.
Both isomers produced good control of dry rot, Fusarium caeruleum (Lib.) Sacc., on five varieties of potatoes, but did not produce a commercial control on Doon Star. To achieve control, the isomers had to be present at the time of infection. There was some evidence that the development of disease symptoms was delayed in treated, infected tubers. Since 2:3:4:5-tetrachloro-1-nitrobenzene has no sprout depressant action it merits further trial as a fungicidal treatment for seed potatoes.  相似文献   

2.
The linear growth of Botrytis cinerea, Fusarium caeruleum and Phoma foveata in culture was reduced in the presence of vapour from any of the three isomers of tetrachloronitrobenzene. The isomers were fungistatic but not fungicidal.
Differences in activity were observed amongst the isomers. The 2:3:4:6 isomer was the most active against all three test fungi. 2:3:5:6-TCNB (tecnazene) was more active than 2:3:4:5-TCNB against Botrytis cinerea , but less active against Fusarium caeruleum and Phoma foveata . Two strains of Fusarium caeruleum resistant to the 2:3:5:6 isomer were not resistant to the other two isomers, although they were more resistant than their 2:3:5:6-TCNB sensitive parent strains.
Sporulation and sclerotial production by Botrytis cinerea were completely suppressed by 2:3:5:6-TCNB and 2:3:4:6-TCNB but not by the 2:3:4:5: isomer.  相似文献   

3.
The hydrolysis and acyl migration of biosynthetic S-naproxen-beta-l-O-acyl glucuronide (I) and R-naproxen-beta-l-O-acyl glucuronide (II) was followed by HPLC. Nine first-order kinetic rate constants for the hydrolysis and acyl migration between the beta-l-O-acyl glucuronide, its alpha/beta-2, alpha/beta-3-, alpha/beta-4-, and alpha-1-O-acyl isomers and naproxen aglycone were determined for I and II at pH 7.00, 7.40 and 8.00 at 37 degrees C by kinetic simulation. For I the 3-O-acyl isomer was the most stable isomer as the pseudo-equilibrium ratio for the major acyl-migrated isomers was 1:1.5:0.9 (2-O-acyl isomer:3-O-acyl isomer:4-O-acyl isomer). The 3- and 4-O-acyl isomers of II were equally stable as the pseudo-equilibrium ratio for the major acyl-migrated isomers was 1:1.4:1.4 (2-O-acyl isomer:3-O-acyl isomer:4-O-acyl isomer). For both I and II, the pseudo-equilibrium ratio between the major 2-O-acyl isomer and the minor alpha-l-O-acyl isomer was 10:1 (2-O-acyl isomer:alpha-l-O-acyl isomer). The pseudo-equilibrium found for the major acyl-migrated isomers of I and II in the present study corresponds with the pattern previously published for R- and S-ketoprofen-beta-l-O-acyl glucuronide acyl-migrated isomers, suggesting that these findings may be general for acyl-migrated beta-l-O-acyl glucuronides of enantiomeric 2-arylpropionic acids.  相似文献   

4.
One carbonyl oxygen of the cyclic hexapeptide cyclo(-Gly1-Pro2-Phe3-Val4-Phe5-Phe6-) (A) can be selectively exchanged with sulphur using Yokoyama's reagent. Surprisingly it was not the C=] of Gly1 but that of Phe5 which was substituted and cyclo(-Gly1-Pro2-Phe3-Val4-Phe5 psi [CS-NH]Phe6-) (B) was obtained. Thionation results in a conformational change of the peptide backbone although the C=O of Phe5 and the corresponding C=S are not involved in internal hydrogen bonds. Two isomers in slow exchange, containing a cis Gly1-Pro2 bond in a beta VIa-turn (minor) and a trans Gly-Pro bond in a beta II'-turn (major), were analyzed by restrained molecular dynamics in vacuo and in DMSO as well as using time dependent distance constraints. It is impossible to fit all experimental data to a static structure of each isomer. Interpreting the conflicting NOEs, local segment flexibility is found. MD simulations lead to a dynamic model for each structure with evidence of an equilibrium between a beta I- and beta II-turn about the Val4-Phe5 amide bond in both the cis and trans isomers. Additionally proton relaxation rates in the rotating frame (R1 rho) were measured to verify the assumption of this fast beta I/beta II equilibrium within each isomer. Significant contributions to R1 rho-rates from intramolecular motions were found for both isomers. Therefore it is possible to distinguish between at least four conformers interconverting on different time scales based on NMR data and MD refinement. This work shows that thionation is a useful modification of peptides for conformation-activity investigations.  相似文献   

5.
Y S Yang  P A Frey 《Biochemistry》1986,25(25):8173-8178
The dihydrolipoyl transacetylase component (E2) of the pyruvate dehydrogenase complex catalyzes the reaction of acetyl coenzyme A (acetyl-CoA) with dihydrolipoamide, producing coenzyme A and S-acetyldihydrolipoamide. The acetyl group is shown by experiments reported herein to be bonded to S8 in the enzymatic product. 1H NMR analysis of synthetic samples of both structural isomers of S-acetyl-S-(phenylmercurio)dihydrolipoamide enabled structural assignments to be made. Reaction of 8-S-acetyl-6-S-(phenylmercurio)dihydrolipoamide with 3-mercaptopropionic acid in chloroform produced 8-S-acetyldihydrolipoamide which contained a small amount (5%) of the 6-S isomer. Reaction of 6,8-di-S-acetyldihydrolipoamide with NH2OH produced a 4:1 mixture of 6-S-acetyldihydrolipoamide and the 8-S isomer. These compounds did not isomerize at significant rates in chloroform but rapidly isomerized to the equilibrium mixture in aqueous solution (Keq = 3.4). The second-order rate constants for the hydroxide-catalyzed isomerization were found to be kf = (1.15 +/- 0.07) X 10(6) M-1 X s-1 and kr = (3.36 +/- 0.20) X 10(5) M-1 X s-1 in the direction of the formation of the 8-S isomer. The enzymatic product was trapped by addition of phenylmercuric hydroxide within 15 s-30 min after starting the reaction. 1H NMR analysis of the products obtained at various times showed that the enzymatic product was 8-S-acetyldihydrolipoamide, which underwent progressive isomerization to the mixture of isomers within a few minutes. In the reaction of acetyl-CoA with dihydrolipoamide, the latter substrate reacts in place of enzyme-bound dihydrolipoyl moieties. Therefore, acetylation occurs at the 8-S position of bound lipoyl groups.  相似文献   

6.
M Numazawa  Y Osawa 《Steroids》1979,34(3):347-360
The synthesis of epimeric 6-bromo-4-androstene-3,17-dione (1a and 1b), 6-bromotestosterone (2a and 2b) and its acetate (3a and 3b), and 6-bromo-16 alpha-acetoxy-4-androstene-3,17-dione (5a and 5b), and 6 beta-bromo-16 alpha-hydroxy-4-androstene-3,17-dione (4) is described. The interconversions among compounds 1, 2, and 3 are also studied. The 6 beta-isomer (1b, 2b, and 3b) was epimerized to the 6 alpha-isomer (1a, 2a and 3a) in carbon tetrachloride or chloroform-methanol (9:1) and the 6 alpha-isomer was isolated by fractional crystallization from the epimeric mixture. 6 alpha-Bromo isomer 1a was also epimerized back to 6 beta-bromo isomer 1b in chloroform-methanol (9:1). Two polymorphic forms of 6 beta-bromotestosterone acetate (3b) were isolated (mp. 114--117 degrees and 138--141 degrees). The 6 beta-bromo isomers were found to be unstable in methanol and decomposed to give 5 alpha-androstane-3,6-dione derivative (6). The results of irreversible inactivation of human placental androgen aromatase with some of these 6-bromoandrogens are discussed.  相似文献   

7.
A series of inhibitors containing all possible isomers of 4-amino-3-hydroxy-5-phenylpentanoic acid was synthesized and tested for inhibition of HIV-1 protease. Incorporation of the (3S,4S) isomer of the t-butyloxycarbonyl protected amino acid into the sequence Glu-Phe resulted in a potent inhibitor of HIV-1 protease (Ki = 63 nM). This inhibitor is at least 47-times more potent than the inhibitors containing other isomers of 4-amino-3-hydroxy-5-phenylpentanoic acid, indicating that the (3S,4S) isomer is the preferred isomer for binding to HIV-1 protease.  相似文献   

8.
Oligothymidylate analogues having stereoregular, alternating methylphosphonate/phosphodiester backbones, d-Tp(TpTp)4T isomers I and II and d-Tp(TpTp)3T(pT)1-5 isomers I and II, were prepared by methods analogous to the phosphotriester synthetic technique. The designations isomer I nd isomer II refer to the configuration of the methylphosphonate linkage, which is the same through each isomer. Analogues with the type I methylphosphonate configuration form very stable duplexes with poly(dA) while those with the type II configuration form either 2T:1A triplexes or 1T:1A duplexes with poly(dA) of considerably lower stabilities. The oligothymidylate analogues were tested for their ability to initiate polymerizations catalyzed by Escherichia coli DNA polymerase I or calf thymus DNA polymerase alpha on a poly(dA) template. Neither d-Tp(TpTp)4T nor d-Tp(T]Tp)3TpT served as initiators of polymerization while d-Tp(TpTp)3T(pT)2-5 showed increasing priming ability as the length of the 3'-oligothymidylate tail increased. Analogues with type I methylphosphonate configuration were more effective initiators than the type II analogues at 37 degrees C. The apparent activation energies of polymerizations initiated by d-Tp(TpTp)3T-(pT)4 and 5 isomer I were greater than those for reactions initiated by isomer II or d-(Tp)11T. The results suggest that DNA polymerase interacts with the charged phosphodiester groups of the primer molecule and may help stabilize primer/template interaction. At least two contiguous phosphodiester groups are required at the 3' end of the analogue primers in order for polymerization to occur. Interactions between the polymerase and primer also appear to occur with phosphodiester groups located at sites remote from the 3'-OH polymerization site and may be influenced by the configuration of the methylphosphonate group.  相似文献   

9.
Inhibitory properties of 6E (compound 1) and 6Z (compound 2) isomers of 2,3-epoxy-10-aza-10,11-dihydrosqualene against oxidosqualene-lanosterol cyclase were assayed on microsomes and whole cells of Saccharomyces cerevisiae and Candida albicans. Only the 6E isomer (compound 1), bearing a correct substrate-like configuration, strongly inhibited the enzyme both in microsomes and cell cultures. The difference between compounds 1 and 2 (which had an unfavorable geometry) was especially evident when measuring [14C]acetate incorporation into non-saponifiable lipids extracted from treated cells. While isomer Z was totally ineffective at up to 30 μM, in cells treated with 5 μM isomer E, labelled oxidosqualene, the level of which was negligible in the control, rose to over 60% of the non-saponifiable lipids.  相似文献   

10.
The substrate specificity of the different molecular forms of biliverdin reductase (bilirubin:NAD(P)+ oxidoreductase, EC 1.3.1.24) using biliverdin XIII alpha, XIII beta and XIII gamma was examined. It was found that molecular form 1 (the major form in normal rat liver) reduced biliverdin XIII alpha at a much higher rate than the other two isomers. Molecular form 2 (the minor form) reduced isomers XIII alpha and XIII beta at similar rates, while molecular form 3 (the major form induced by CoCl2 treatment) reduced the XIII beta isomer at a slightly higher rate than the XIII alpha isomer. Molecular forms 2 and 3, both reduced isomer XIII gamma more slowly than they reduced the XIII alpha and XIII beta isomers. These results are similar to those obtained previously using biliverdins IX alpha, IX beta and IX gamma, suggesting that biliverdin reductase specificity is related to the type of the isomer rather than to the series (IX or XIII) of the isomer.  相似文献   

11.
Abstract

The preparation of 6 substituted pyrimidine nucleosides has received limited attention and undoubtedly reflects the difficulty in synthesizing nucleosides of this type. Condensation of & substituted pyrimidines with suitable sugar derivatives leads to the formation of mixtures of N3 and N1 nucleosides where the N3 isomer usually predominates1. This is exemplified by the direct ribosylation of the silyl derivative of 6-methyl-thiouracil, which furnished only the N3 ribonucleoside2. Ueda and coworkers' adcfessed this problem with moderate success. When 5′- O-acetyl-2′,3′-O-isopropylidine5bromouridine c1) was reacted with cyan- ide ion, a Michael-type addition occurred at C6 with concomitant dehycfo- brominatim to give the corresponding Gcyanowidine in quantitative yield. Treatment of 1(Scheme 1) with benzyl mercaptan, however furnished a 1:1 mixture of the C6 and C5 isomers 2 and 3 grespectively4. Attempts to alter the course of this reaction so that 2 predominated met with little success. It is worth mentioning that in ouFhands when this reaction was scaled-up, 3 predominated (2:3=1:4). Also the use of other sulfur nucleophiles, such as SEt, afforded only the C5-substituted derivative3. Thus, a new synthetic approach was sought which would furnish only the desired C6-substituted isomer and in reasonable yield.  相似文献   

12.
Numazawa M  Nagaoka M  Sohtome N 《Biochemistry》2005,44(32):10839-10845
Aromatase is a cytochrome P-450 enzyme complex that catalyzes the conversion of androst-4-ene-3,17-dione (AD) to estrone and formic acid through three sequential oxygenations of the 19-methyl group. To gain insight into the catalytic function of aromatase as well as the mechanism of the hitherto uncertain third oxygenation step, we focused on the aromatase-catalyzed 19-oxygenation of 3-deoxyandrogens: 3-deoxy-AD (1), which is a very powerful competitive inhibitor but poor substrate of aromatase, and its 5-ene isomer 4, which is a good competitive inhibitor and effective substrate of the enzyme. In incubations of their 19S-(3)H-labeled 19-hydroxy derivatives 2 and 5 and the corresponding 19R-(3)H isomers with human placental microsomes in the presence of NADPH under air, the radioactivity was liberated in both water and formic acid. The productions of (3)H(2)O and (3)HCOOH were blocked by the substrate AD or the inhibitor 4-hydroxy-AD, indicating that these productions are due to a catalytic function of aromatase. A comparison of the (3)H(2)O production from S-(3)H substrates 2 and 5 with that from the corresponding R-(3)H isomers revealed that the 19-pro-R hydrogen atom was stereospecifically (pro-R:pro-S = 100:0) removed in the conversion of 5-ene substrate 5 into the 19-oxo product 6, whereas 75:25 stereoselectivity for the loss of the pro-R and pro-S hydrogen atoms was observed in the oxygenation of the other substrate, 2. The present results reveal that human placental aromatase catalyzes three sequential oxygenations at C-19 of 3-deoxyandrogens 1 and 4 to cause the cleavage of the C(10)-C(19) bond through their 19-hydroxy (2 and 5) and 19-oxo (3 and 6) intermediates, respectively, where there is a difference in the stereochemistry between the two androgens in the second 19-hydroxylation. It is implied that the aromatase-catalyzed 19-oxygenation of 5-ene steroid 4 but not the 4-ene isomer 1 would proceed in the same steric mechanism as that involved in the AD aromatization.  相似文献   

13.
As a continuation of our interest in novel 99mTc chelating systems, several pyridine-containing HYNIC (6-hydrazinonicotinamide) derivatives (L1-L5) have been synthesized and characterized by NMR (1H and 13C) and LC-MS. 99mTc complexes of L1-L5 were prepared by the reaction of the HYNIC derivative with 99mTcO4- in the presence of excess tricine and stannous chloride. Results from this study show that the attachment site of the linker is critical for the formation of macrocyclic 99mTc complexes. For example, the pyridine-N in L3 is not able to bond to the Tc, because the lysine linker is attached to the 4-position. When the linker is at the 2-position, L1 forms the macrocyclic complex [99mTc(L1)(tricine)], but the radiochemical purity is relatively low. If the linker is attached to the 3-position of the pyridine ring, the HYNIC derivatives form macrocyclic complexes [99mTc(L)(tricine)] (L2, L4, and L5) in high yield (>95%). The HPLC data suggest that the macrocyclic complex [(99m)Tc(L2)(tricine)] exists in solution as four isomers: two diastereomers and two conformational isomers. Diastereomers are due to a combination of the chirality of the lysine linker and of the Tc chelate. Replacing lysine with a pentamethylenediamine linker results in the macrocyclic complex [99mTc(L4)(tricine)] with two conformational isomers, which interconvert rapidly at room temperature. Changing the linker from pentamethylenediamine to hexamethylenediamine did not eliminate the minor isomer; but the percentage of the minor isomer was reduced from approximately 10% for [99mTc(L4)(tricine)] to only 6% for [99mTc(L5)(tricine)]. The linker length is an important parameter to minimize the minor isomer. LC-MS data of complexes [99mTc(L)(tricine)] (L2, L4, and L5) are completely consistent with their proposed compositions. On the basis of these data, it is concluded that pyridine-containing HYNIC derivatives have the potential as bifunctional chelators for 99mTc-labeling of small biomolecules if the linker is attached to the 3-position of the pyridine ring.  相似文献   

14.
The solution conformations of two potent antagonists of bradykinin (Arg1-Pro2-Pro3-Gly4-Phe5-Ser6-Pro7-Phe8-Arg9), [Aca(-1),DArg0,Hyp3,Thi5,DPhe7,(N-Bzl)Gly8]BK (1) and [Aaa(-1),DArg0,Hyp3,Thi5,(2-DNal)7,Thi8]BK (2), were studied by using 2D NMR spectroscopy in DMSO-d6 and molecular dynamics simulations. The NMR spectra of peptide 1 reveals the existence of at least two isomers arising from isomerization across the DPhe7-(N-Bzl)Gly8 peptide bond. The more populated isomer possesses the cis peptide bond at this position. The ratio of cis/trans isomers amounted to 7:3. With both antagonists, the NMR data indicate a beta-turn structure for the Hyp3-Gly4 residues. In addition, for peptide 2, position 2,3 is likely to be occupied by turn-like structures. The cis peptide bond between DPhe7 and (N-Bzl)Gly8 in analogue 1 suggests type VI beta-turn at position 7,8. The molecular dynamics runs were performed on both peptides in DMSO solution. The results indicate that the structure of peptide 1 is characterized by type VIb beta-turn comprising residues Ser6-Arg9 and the betaI or betaII-turn involving the Pro2-Thi5 fragment, whereas peptide 2 shows the tendency towards the formation of type I beta-turn at position 2,3. The structures of both antagonists are stabilized by a salt bridge between the guanidine moiety of Arg1 and the carboxyl group of Arg9. Moreover, the side chain of DArg0 is apart of the rest of molecule and is not involved in structural elements except for a few calculated structures.  相似文献   

15.
Equilibrium binding studies and viscosity experiments are described that characterize the interaction of delta- and lambda-[Ru(o-phen)3]2+ with calf thymus DNA. The mode of binding of these compounds to DNA is a matter of controversy. Both isomers of [Ru(o-phen)3]2+ were found to bind but weakly to DNA, with binding constants of 4.9 (+/- 0.3) x 10(4) M-1 and 2.8 (+/- 0.2) x 10(4) M-1 determined for the delta and lambda isomers, respectively, at 20 degrees C in a solution containing 5 mM Tris-HCl (pH 7.1) and 10 mM NaCl. We determined that the quantity delta log K/delta log [Na+] equals 1.37 and 1.24 for the delta and lambda isomers, respectively. Application of polyelectrolyte theory allows us to use these values to show quantitatively that both the delta and lambda isomers are essentially electrostatically bound to DNA. Viscosity experiments show that binding the lambda isomer does not alter the relative viscosity of DNA to any appreciable extent, while binding of the delta isomer decreases the relative viscosity of DNA. From these viscosity results, we conclude that neither isomer of [Ru(o-phen)3]2+ binds to DNA by classical intercalation.  相似文献   

16.
F Eckstein  R S Goody 《Biochemistry》1976,15(8):1685-1691
The chemical synthesis of adenosine 5'-(O-1-thiotriphosphate) (ATPalphaS) and adenosine 5'-(O-2-thiotriphosphate) (ATPbetaS) is described. Both exist as a pair of diastereomers, A and B. The isomers of ATPalphaS can be distinguished on the basis of their different reaction rates with myokinase as well as nucleoside diphosphate kinase. With both enzymes, isomer A reacts fast whereas isomer B reacts considerably more slowly. Phosphorylation of a mixture of isomers of ADPalphaS with pyruvate or acetate kinase yields ATPalphaS, isomer A, whereas the phosphoryl transfer with creatine or arginine kinase yields isomer B. The isomers of ATPbetaS differ in their reactivity with myosin. Isomer A is readily hydrolyzed, whereas isomer B is not. However, isomer B reacts faster with nucleoside diphosphate kinase and ADP than isomer A. Phosphoryl transfer with pyruvate kinase onto ADPbetaS yields ATPbetaS, isomer A, with acetate kinase, isomer B.  相似文献   

17.
Adducts of catechols and histidine, which are produced by reactions of 1,2-quinones and p-quinone methides with histidyl residues in proteins incorporated into the insect exoskeleton, were characterized using electrospray ionization mass spectrometry (ESMS), tandem electrospray mass spectrometry (ESMS-MS, collision-induced dissociation), and ion trap mass spectrometry (ITMS). Compounds examined included adducts obtained from acid hydrolysates of Manduca sexta (tobacco hornworm) pupal cuticle exuviae and products obtained from model reactions under defined conditions. The ESMS and ITMS spectra of 6-(N-3')-histidyldopamine [6-(N-3')-His-DA, pi isomer] isolated from M. sexta cuticle were dominated by a [M + H]+ ion at m/z 308, rather than the expected m/z 307. High-resolution fast atom bombardment MS yielded an empirical formula of C14H18N3O5, which was consistent with this compound being 6-(N-1')-histidyl-2-(3, 4-dihydroxyphenyl)ethanol [6-(N-1')-His-DOPET] instead of a DA adduct. Similar results were obtained when histidyl-catechol compounds linked at C-7 of the catechol were examined; the (N-1') isomer was confirmed as a DA adduct, and the (N-3') isomer identified as an (N-1')-DOPET derivative. Direct MS analysis of unfractionated cuticle hydrolysate revealed intense parent and product ions characteristic of 6- and 7-linked adducts of histidine and DOPET. Mass spectrometric analysis of model adducts synthesized by electrochemical oxidative coupling of N-acetyldopamine (NADA) quinone and N-acetylhistidine (NAcH) identified the point of attachment in the two isomers. A prominent product ion corresponding to loss of CO2 from [M + H]+ of 2-NAcH-NADA confirmed this as being the (N-3') isomer. Loss of (H2O + CO) from 6-NAcH-NADA suggested that this adduct was the (N-1') isomer. The results support the hypothesis that insect cuticle sclerotization involves the formation of C-N cross-links between histidine residues in cuticular proteins, and both ring and side-chain carbons of three catechols: NADA, N-beta-alanyldopamine, and DOPET.  相似文献   

18.
Immune response suppressors are used in the medical praxis to prevent graft rejection after organ transplantation and in the therapy of some autoimmune diseases. As a continuation of our previous work searching for new, effective suppressors devoid of toxicity, we present the synthesis, conformational analysis, and biological activity of nonapeptides 1-6, analogs of naturally existing immunomodulatory peptide CLA. New CLA analogs were modified with (S)-beta(2)-iso-proline 7 or (S)-beta(3)-homo-proline 8, respectively. The conformational influence of the beta-iso-proline and beta-homo-proline building blocks was analyzed by NMR spectroscopy. Peptides 1-6 exist as a mixture of four isomers due to cis/trans isomerization of the Xxx-Pro peptide bond. The major isomers of peptides 1, 3, and 4 contain all peptide bonds of the trans geometry. The geometry of the proline-proline bond of the second populated isomer of peptides 3 and 4 is cis. The proline-proline peptide bond is cis for the major isomers of peptides 2, 5, and 6. The peptides were tested for their ability to suppress the proliferative response of mouse splenocytes to T- and B-cell mitogens and the secondary humoral immune response to sheep erythrocytes in vitro in parallel with a reference drug-cyclosporine A. The immunoregulatory actions of the peptides depended on the position and content of proline isomers and were, with some exceptions, strongly inhibitory at the highest dose tested (100 microg/ml). In addition, the peptides were practically devoid of toxicity at that dose. In conclusion, the replacement of Pro by beta-Pro may be useful for fine-tuning CLA immunosuppressive potency and undesirable toxicity.  相似文献   

19.
To find a cost effective alternative substrate, Pleurotus cornucopiae 608 (yellow basidiomata) was grown on: (1) chopped, pasteurized switch grass (Panicum virgatum, 99%) with 1% ground limestone and (2) a mixture of pasteurized cottonseed hulls (75% dry wt.), 24% chopped wheat straw, and 1% ground limestone (all ingredients wt./wt.). The substrates were spawned at various levels (2.5%, 3.75% or 5% wet wt., crop I) and non-supplemented or supplemented with commercial delayed release nutrient (Campbell's S-41) at various levels (0%, 1.5%, 3%, 4.5%, 6%, 7.5% and 9% dry wt., crop II). Maximum yield (weight of fresh mushrooms harvested at maturity) was obtained on cottonseed hull/wheat straw substrate at a 3.75-5% spawn level and 6% S-41 supplement. On switch grass substrate, increasing spawn levels and supplement levels stimulated yields in a linear fashion. However, maximum yields were only 46% or less for those of similar treatments on cottonseed hull/wheat straw substrate. Yields were three times higher on switch grass that was harvested after the grass had senesced (winter; beige color) compared to material that was harvested when the grass was green (summer; time of flowering). Additional physical processing of the material, such as milling, may improve yield potential of this material.  相似文献   

20.
We investigated a novel method for the selective separation of beta-carotene isomers from a freeze-dried powder of the algae Dunaliella bardawil using supercritical fluid extraction. The separation method relies on the different dissolution rate of the 9Z and all-E isomers of beta-carotene in SC-CO(2). At first, the equilibrium solubility of the two isomers in SC-CO(2) was determined at the extraction conditions of 44.8 MPa and 40 degrees C. The solubility of the 9Z isomer was found to be nearly 4 times higher than that of the all-E isomer (1.92 x 10(-5) g all-E isomer/g CO(2) compared to 7.64 x 10(-5) g 9Z isomer/g CO(2)). When supercritical fluid extraction was applied to a carotenoid concentrate from the algae (29 wt% beta-carotene) or a freeze-dried powder of the algae (3.1% beta-carotene), a selective separation of the 9Z/all-E isomers of beta-carotene was obtained. Thirty-nine percent recovery of beta-carotene with 80% purity of 9Z isomer was achieved at the initial stages of extraction (40 mL CO(2)). The extraction rate of beta-carotene from the freeze-dried algae powder was slower than that from the carotenoid concentrate, resulting in a reduction in the recovery and purity of the 9Z isomer. This indicates that even at the initial stage of the extraction the internal mass resistance is significant. Isomer purity and recovery could be enhanced upon grinding of the algae powder.  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