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 91 毫秒
1.
A method for the preparation of insulin derivatives having protected sulfhydryl group(s) on definite site(s) on the molecule which uses anion-exchange high performance liquid chromatography on a TSKgel DEAE-2SW column for separation is described. Porcine insulin reacts with N-succinimidyl S-acetylthioacetate to afford four species of insulin derivatives that have S-acetylthioglycoloyl group(s) at: i) Gly(A1), ii) Gly(A1) and Phe(B1), iii) Gly(A1) and Lys(B29), and iv) Gly(A1), Phe(B1) and Lys(B29) positions. An insulin derivative which has a group at the Lys(B29)-position is prepared by the S-acetylthioglycoloylation of Gly(A1), Phe(B1)-dicitraconyl insulin followed by decitraconylation. The five derivatives are readily deacetylated with hydroxylamine to yield the corresponding sulfhydryl insulin derivatives.  相似文献   

2.
Bovine insulin was glycated under hyperglycemic reducing conditions and in nonreducing conditions. Purification through HPLC allowed isolating glycated forms of insulin and a novel triglycated form (6224.5 Da) was purified. Endoproteinase Glu-C digestion combined with mass spectrometry (MALDI-TOF/TOF) allowed determining the exact location of the glycation sites in each of the isolated glycated insulins. For the first time, a triglycated form of insulin was isolated and characterized accordingly to its glycation sites. These glucose binding sites were identified as the N-terminals of both chains (Gly1 and Phe1) and residue Lys29 of B-chain. Moreover, in diglycated insulin we found the coexistence of one specie glycated at the N-terminals of both chains (Gly1 and Phe1) and another specie containing the two glucitol adducts in B-chain (Phe1 and Lys29). Also, in monoglycated insulin generated in reducing and nonreducing conditions, one specie glycated at Phe1 and another specie glycated at Lys29, both B-chain residues coexist.  相似文献   

3.
Ricin and its corresponding polypeptides (A & B chain) were purified from castor seed. The molecular weight of ricin subunits were 29,000 and 28,000 daltons. The amino acids in ricin determined were Asp45 The22 Ser40 Glu53 Cys4 Gly96 His5 Ile21 Leu33 Lys20 Met4 Phe13 Pro37 Tyr11 Ala45 Val23 Arg20 indicating that ricin contains approximately 516 amino acid residues. The amino acids of the two subunits of ricin A and B chains were Asp23 The12 Ser21 Glu29 Cys2 Gly48 His3 Ile12, Leu17 Lys10 Met2 Phe6 Pro17 Tyr7 Ala35 Val13 Arg13 while in B chain the amino acids were Asp22 The10 Ser19 Glu25 Cys2 Gly47 His1 Ile10, Leu15 Lys11 Met1 Phe7 Pro6 Tyr5 Ala32Val11 Arg10. The total helical content of ricin came around 53.6% which is a new observation.  相似文献   

4.
KIA7, a peptide with a highly restricted set of amino acids (Lys, Ile, Ala, Gly and Tyr), adopts a specifically folded structure. Some amino acids, including Lys, Ile, Ala, Gly and His, form under the same putative prebiotic conditions, whereas different conditions are needed for producing Tyr, Phe and Trp. Herein, we report the 3D structure and conformational stability of the peptide KIA7H, which is composed of only Lys, Ile, Ala, Gly and His. When the imidazole group is neutral, this 20‐mer peptide adopts a four‐helix bundle with a specifically packed hydrophobic core. Therefore, one‐pot prebiotic proteins with well‐defined structures might have arisen early in chemical evolution. The Trp variant, KIA7W, was also studied. It adopts a 3D structure similar to that of KIA7H and its previously studied Tyr and Phe variants, but is remarkably more stable. When tested for ribonucleolytic activity, KIA7H, KIA7W and even short, unstructured peptides rich in His and Lys, in combination with Mg++, Mn++ or Ni++ (but not Cu++, Zn++ or EDTA) specifically cleave the single‐stranded region in an RNA stem–loop. This suggests that prebiotic peptide–divalent cation complexes with ribonucleolytic activity might have co‐inhabited the RNA world.  相似文献   

5.
The results from the study on the separation, purification, amino acid composition and amino acid sequence of CBa, one of the four CNBr degradation fragments of crystalline trichosanthin, are presented. Its amino acid composition is: Asp3, Thr2, Ser2, Hse1, Glu2, Gly2, Ala6, Val1, Tyr3, Phe3, Lys2, Arg1. The sequence of the CBa is Gly-Tyr-Arg-Ala-Gly-Asp-Thr-Ser- Tyr-Phe-Phe-Asn-Glu-Ala-Ser-Ala-Thr-Glu-Ala-Ala-Lys-Tyr-Val- Phe-Lys-Asp-Ala-Hso.  相似文献   

6.
From the roots of Pseudostellaria heterophylla, three cyclopeptides and three amides were isolated, besides heterophyllin A and B. Their structures were determined as cyclo (Ala‐Gly‐Pro‐Val‐Tyr‐) (heterophyllin J; 1 ), cyclo (Ala‐Gly‐Pro‐Tyr‐Leu‐) (pseudostellarin A; 2 ), cyclo (Gly‐Gly‐Gly‐Pro‐Pro‐Phe‐Gly‐Ile‐) (pseudostellarin B; 3 ), methyl γ‐hydroxypyroglutamate ( 4 ), methyl pyroglutamate ( 5 ), and pyroglutamic acid ( 6 ) on the basis of spectral data, especially 2D‐NMR data. Among them, compounds 1 and 4 are new compounds.  相似文献   

7.
Interaction of cis-[Pt(en)(H2O)2]2+ and [CuL(H2O)]2+, where L is 2-[bis(2-aminoethyl)amino]ethanol, with oxidized insulin B chain in molar ratio of 1 : 1, 1 : 2 and 1 : 3 at pH 2.5 and 40 degrees C has been investigated by electrospray ionization mass spectrometry (ESI-MS) and tandem mass spectrometry (MS/MS). The results show that the binding sites of the two complexes with oxidized insulin B chain are terminal NH2, imidazole groups of His5 and His10. The hydrolytic cleavage studies show that the [CuL(H2O)]2+, upon a pendant hydroxyl group of the ligand, selectively cleaves the peptide bonds at Gly8-Ser9, Asn3-Gln4 and Phe1-Val2, and the cis-[Pt(en)(H2O)2]2+ only cleaves the peptide bond at His10-Leu11. This is the first report of cis-[Pt(en)(H2O)2]2+-promoted cleavage of His-X peptide bond.  相似文献   

8.
Two novel mono-PEGylated derivatives of hGRF(1-29)-NH(2) [human growth hormone-releasing factor, fragment 1-29] have been synthesized by regio-specific conjugation of Lys(12) or Lys(21) to a monomethoxy-PEG(5000) chain (compounds Lys(12)PEG-GRF and Lys(21)PEG-GRF). The PEG moiety has been covalently linked at the amino group of a norleucine residue via a carbamate bond. The Lys(12)PEG-GRF regioisomer was found to be slightly less active in vitro than both the unmodified peptide and Lys(21)PEG-GRF. To assess whether the differences in the biological activity of the PEGylated analogues could be related to conformational rearrangements induced by the PEG moiety, the structure of these PEGylated derivatives has been worked out (TFE solution) by means of NMR spectroscopy and molecular dynamics. Secondary structure shifts, hydrogen/deuterium exchange kinetics, temperature coefficients of amide protons, and NOE-based molecular models point out that hGRF(1-29)-NH(2), Lys(21)PEG-GRF and Lys(12)PEG-GRF share a remarkably similar pattern of secondary structure. All three compounds adopt an alpha-helix conformation which spans the whole length of the molecule, and which becomes increasingly rigid on going from the N-terminus to the C-terminus. Residues Lys(12) and Lys(21) are enclosed in all the compounds considered into well-defined alpha-helical domains, indicating that PEGylation either at Lys(12) or Lys(21) does not alter the tendency of the peptide to adopt a stable alpha-helix conformation, nor does it induce appreciable conformational mobility in the proximity of the PEGylation sites. No significant variation of the amphiphilic organization of the alpha-helix is observed among the three peptides. Therefore, the different biological activities observed for the PEGylated analogues are not due to conformational effects, but are rather due to sterical hindrance effects. The relationship between the biological activitiy of the mono-PEGylated derivatives and sterical hindrance is discussed in terms of the topology of interaction between hGRF(1-29)-NH(2) and its receptor.  相似文献   

9.
Cation-pi interactions between amino acid side chains are increasingly being recognized as important structural and functional features of proteins and other biomolecules. Although these interactions have been found in static protein structures, they have not yet been detected in dynamic biomolecular systems. We determined, by (1)H NMR spectroscopic titrations, the energies of cation-pi interactions of the amino acid derivative AcLysOMe (1) with AcPheOEt (2) and with AcTyrOEt (3) in aqueous and three organic solvents. The interaction energy is substantial; it ranges from -2.1 to -3.4 kcal/mol and depends only slightly on the dielectric constant of the solvent. To assess the effects of auxiliary interactions and structural preorganization on formation of cation-pi interactions, we studied these interactions in the association of pentapeptides. Upon binding of the positively-charged peptide AcLysLysLysLysLysNH(2) (5) to the negatively-charged partner AcAspAspXAspAspNH(2) (6), in which X is Leu (6a), Tyr (6b), and Phe (6c), multiple interactions occur. Association of the two pentapeptides is dynamic. Free peptides and their complex are in fast exchange on the NMR time-scale, and 2D (1)H ROESY spectra of the complex of the two pentapeptides do not show intermolecular ROESY peaks. Perturbations of the chemical shifts indicated that the aromatic groups in peptides 6b and 6c were affected by the association with 5. The association constants K(A) for 5 with 6a and with 6b are nearly equal, (4.0 +/- 0.7) x 10(3) and (5.0 +/- 1.0) x 10(3) M(-)(1), respectively, while K(A) for 5 with 6c is larger, (8.3 +/- 1.3) x 10(3) M(-)(1). Molecular-dynamics (MD) simulations of the pentapeptide pairs confirmed that their association is dynamic and showed that cation-pi contacts between the two peptides are stereochemically possible. A transient complex between 5 and 6 with a prominent cation-pi interaction, obtained from MD simulations, was used as a template to design cyclic peptides C(X) featuring persistent cation-pi interactions. The cyclic peptide C(X) had a sequence in which X is Tyr, Phe, and Leu. The first two peptides do, but the third does not, contain the aromatic residue capable of interacting with a cationic Lys residue. This covalent construct offered conformational stability over the noncovalent complexes and allowed thorough studies by 2D NMR spectroscopy. Multiple conformations of the cyclic peptides C(Tyr) and C(Phe) are in slow exchange on the NMR time-scale. In one of these conformations, cation-pi interaction between Lys3 and Tyr9/Phe9 is clearly evident. Multiple NOEs between the side chains of residues 3 and 9 are observed; chemical-shift changes are consistent with the placement of the side chain of Lys3 over the aromatic ring. In contrast, the cyclic peptide C(Leu) showed no evidence for close approach of the side chains of Lys3 and Leu9. The cation-pi interaction persists in both DMSO and aqueous solvents. When the disulfide bond in the cyclic peptide C(Phe) was removed, the cation-pi interaction in the acyclic peptide AC(Phe) remained. To test the reliability of the pK(a) criterion for the existence of cation-pi interactions, we determined residue-specific pK(a) values of all four Lys side chains in all three cyclic peptides C(X). While NOE cross-peaks and perturbations of the chemical shifts clearly show the existence of the cation-pi interaction, pK(a) values of Lys3 in C(Tyr) and in C(Phe) differ only marginally from those values of other lysines in these dynamic peptides. Our experimental results with dynamic peptide systems highlight the role of cation-pi interactions in both intermolecular recognition at the protein-protein interface and intramolecular processes such as protein folding.  相似文献   

10.
The gas‐phase geometries of neutral, protonated, and deprotonated forms of some biologically important molecules, alanine (Ala), glycine (Gly), phenylalanine (Phe), and tyrosine (Tyr), were optimized using density functional theory at B3LYP/6‐311++G(d) and the ab initio HF/6‐311++G(d) level of theories. The neutral and different stable ionic states of Ala, Gly, Phe, and Tyr have also been solvated in aqueous medium using polarizable continuum model for the determination of solvation free energies in the aqueous solution. The gas‐phase acidity constants of above four molecules have been also calculated at both levels of theories and found that the values calculated at HF/6‐311++G(d) method are in good agreement with experimental results. A thermodynamic cycle was used to determine the solvation free energies for the proton dissociation process in aqueous solution and the corresponding pKa values of these molecules. The pKa values calculated at B3LYP/6‐311++G(d) method are well supported by the experimental data with a mean absolute deviation 0.12 pKa units. Additionally, the chemical hardness and the ionization potential (IP) for these molecules have been also explored at both the level of theories. The Tyr has less value of chemical hardness and IP at both levels of theories compared with other three molecules, Ala, Gly, and Phe. The calculated values of chemical hardness and IP are decreasing gradually with the substitution of the various functional groups in the side chain of the amino acids. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

11.
Synthesis and conformational studies of peptides containing the E-vinylogous prolines 1 (VPro1) and 2 (VPro2), Boc-Ala-Val-VPro1-Xaa-Leu-OMe (3, Xaa = Gly; 4, Xaa = Phe), Boc-Ala-Val-VPro2-Xaa-Leu-OMe (5, Xaa = Gly; 6, Xaa = Phe), Boc-Leu-Ile-Val-VPro1-Xaa-Leu-OMe (7, Xaa = Gly; 8, Xaa = Phe), and Boc-Leu-Ile-Val-VPro2-Xaa-Leu-OMe (9, Xaa = Gly; 10, Xaa = Phe), were carried out. It has been shown that both VPro1 and VPro2 lead to the formation of 12-membered intramolecularly hydrogen bonded structures very similar to type VI beta-turns with a cis Xaa-VPro amide bond in the major conformers in all the peptides 3-10, resulting in the nucleation of beta-hairpin type structures in these molecules in CDCl(3).  相似文献   

12.
Upon hexanal-modification in the presence of NaCNBH(3), the oxidized B chain of insulin becomes mono- and further dialkylated on both the N-terminal and Lys(29) residues. A pseudo-MS(3) study was performed with a triple-quadrupole mass spectrometer on the different modified lysine-containing species to gain further insights into the characteristic fragmentation pattern. These fragmentations, in good agreement with true MS(3) measurements obtained using an ion trap mass spectrometer, highlighted characteristic monoalkylated lysine (immonium-NH(3)) and protonated modified caprolactam ions at m/z 168 and 213, respectively. In contrast, no fragment ion derived from a modified lysine residue (immonium or caprolactam) was observed when dialkylation occurs on Lys(29). However, a fragment ion corresponding to a protonated dihexylamine was observed at m/z 186. This loss, characteristic of dialkylated lysine fragmentation, was also observed upon dialkylation of N(alpha)-acetyllysine with either hexanal or pentanal. On the other hand, acetylation and malondialdehyde-modification of the N(alpha)-acetyllysine side chain led mainly to the corresponding modified (immonium-NH(3)) fragment ions at m/z 126 and 138, respectively. Finally, it was demonstrated that precursor ion scanning for both m/z 168 and 213 ions led to specific and sensitive identification of peptides containing hexanal-modified lysine residues within an unfractionated tryptic digest of hexanal-modified apomyoglobin. Thus, Lys(42), Lys(45), Lys(62), Lys(63), Lys(77), Lys(87), Lys(96), Lys(98), Lys(145) and Lys(147) were found to be modified upon reaction with hexanal.  相似文献   

13.
Resonance assignments of the 1H spectrum of insulin are the basis on which to investigate its solution conformation by using NMR method. Owing to the complicated aggregation behaviour of the molecule to give broadened n. m. r. lines, only limited resonance assignments have been reported. S-sulfonated A and B chains of insulin gave 1H spectra with good resolutions. Based on the 500 MHz absolute 2D-COSY spectrum and 400 MHz phase sensitive DQF-COSY, Relayed-COSY and NOESY spectra of B chain recorded in D2O, all of the spin system identifications of the non-labile protons in the S-sulfonated B chain of insulin were reported including the specific resonance assignments of eight residues: B3Asn, B9Ser, B16Tyr, B22Arg, B26Tyr, B27Thr, B28Pro and B29Lys. The pK values of B16 and B26 tyrosine are 10.65 and 10.60 respectively from pH titration.  相似文献   

14.
Weakly polar interactions between the side-chain aromatic rings and hydrogens of backbone amides (Ar-HN) are found in unique conformational regions. To characterize these conformational regions and to elucidate factors that determine the conformation of the Ar-HN interactions, four 4-ns molecular dynamics simulations were performed using four different low-energy conformations obtained from simulated annealing and one extended conformation of the model tripeptide Ac-Phe-Gly-Gly-NH-CH(3) as starting structures. The Ar(i)-HN(i+1) interactions were 4 times more frequent than were Ar(i)-HN(i+2) interactions. Half of the conformations with Ar(i)-HN(i+2) interactions also contained an Ar(i)-HN(i+1) interaction. The solvent access surface area of the Phe side chain and of the amide groups of Phe1, Gly2, and Gly3 involved in Ar-HN interactions was significantly smaller than in residues not involved in such interactions. The number of hydrogen bonds between the solvent and Phe1, Gly2, and Gly3 amide groups was also lower in conformations with Ar-HN interactions. For each trajectory, structures that contained Ar(i)-HN(i), Ar(i)-HN(i+1), and Ar(i)-HN(i+2) interactions were clustered on the basis of similarity of selected torsion angles. Attraction energies between the aromatic ring and the backbone amide in representative conformations of the clusters ranged from -1.98 to -9.24 kJ mol(-1) when an Ar-HN interaction was present. The most representative conformations from the largest clusters matched well with the conformations from the Protein Data Bank of Phe-Gly-Gly protein fragments containing Ar-HN interactions.  相似文献   

15.
A novel selectively deprotectable triazacyclophane scaffold was used for the design and split-mix synthesis of two libraries of solid-phase bound tripodal synthetic receptors possessing three different amino acid or peptidic arms. In the synthesis of the first library, the two outer arms consisted of amino acid Ala, Arg, Asp, Gln, Gly, Lys, Phe, Ser, Tyr, or Val and the middle arm consisted of amino acid Asn, Glu, His, Leu, or Pro. The second library contained amino acid and/or (di)peptide arms. The arms were different in all library members. The first outer arm consisted of amino acid(s) Ala, Arg, Gln, Phe, or Ser, the second outer arm consisted of amino acid(s) Asp, Gly, Lys, Tyr, or Val, and the middle arm consisted of amino acid(s) Asn, Glu, His, Leu, or Pro, leading to a 27 000 member library of synthetic tripodal receptor molecules. In on-bead screening experiments, a remarkable selectivity of some library members for Fe(3+) was observed and decoding of their structures by Edman degradation revealed consensus sequences with structural resemblance to non-heme iron proteins.  相似文献   

16.
The purification of a trypsin inhibitor from Ascaris lumbricoides var. suum is described. The electrophoretically pure preparation which inhibits trypsin in a specific manner is a relatively small peptide containing 5 Asp, 4 Thr, 1 Ser, 11 Glu, 6 Pro, 6 Gly, 5 Ala, 2 Val, 10 (Cys)1/2, 3 Ile, 2 Phe, 7 Lys, 3 Arg and 1 Try.  相似文献   

17.
To discover novel inhibitors that target the influenza polymerase basic protein 2 (PB2) cap-binding domain (CBD), commercial ChemBridge compound libraries containing 384,796 compounds were screened using a cascade docking of LibDock–LigandFit–GOLD, and 60 compounds were selected for testing with cytopathic effect (CPE) inhibition assays and surface plasmon resonance (SPR) assay. Ten compounds were identified to rescue cells from H1N1 virus-mediated death at non-cytotoxic concentrations with EC50 values ranging from 0.30 to 67.65 μM and could bind to the PB2 CBD of H1N1 with Kd values ranging from 0.21 to 6.77 μM. Among these, four compounds (11D4, 12C5, 21A5, and 21B1) showed inhibition of a broad spectrum of influenza virus strains, including oseltamivir-resistant ones, the PR/8-R292K mutant (H1N1, recombinant oseltamivir-resistant strain), the PR/8-I38T mutant (H1N1, recombinant baloxavir-resistant strain), and the influenza B/Lee/40 virus strain. These compounds have novel chemical scaffolds and relatively small molecular weights and are suitable for optimization as lead compounds. Based on sequence and structure comparisons of PB2 CBDs of various influenza virus subtypes, we propose that the Phe323/Gln325, Asn429/Ser431, and Arg355/Gly357 mutations, particularly the Arg355/Gly357 mutation, have a marked impact on the selectivities of PB2 CBD-targeted inhibitors of influenza A and influenza B.  相似文献   

18.
A potentiometric method was used to determine the stability constants for the various complexes of copper(II) with carbamoylcholine chloride (C) drug as a ligand in the presence of some biorelevant amino acid constituents like glycine (Gly), alanine (Ala), valine (Val), proline (Pro), β-phenylalanine (Phe), S-methylcysteine (Met), threonine (Thr), ornithine (Orn), lysine (Lys), histidine (Hisd), histamine (Hist), and imidazole (Imz) as ligands (L). Stability constants of complexes were determined at 25°C and I = 0.10 mol/L NaNO3. The relative stability of each ternary complex was compared with that of the corresponding binary complexes in terms of Δlog K and % R.S. values. Cu(II) complexes of drug C were synthesized in 1:1 and 1:1:1 M ratios of copper to drug [Cu(C)(NO3)2] (1) and copper to drug to glycine[Cu(C)(Gly)(NO3)].NO3 (2), respectively. Glycine ternary complex with drug and copper [Cu(C)(Gly)(NO3)].NO3 was considered as representative amino acid. The complexes 1 and 2 were isolated and characterized using various physicochemical and spectral techniques. Both complexes 1 and 2 were found to have magnetic moments corresponding to one unpaired electron. The possible square planar and square-pyramidal geometries of the copper (II) complexes were assigned on the basis of electron paramagnetic resonance (EPR), thermogravimetric analysis (TGA), scanning electron microscopy (SEM), X-ray powder diffraction (XRPD), ultraviolet–visible (UV–Vis) and infrared (IR) spectral studies, and the discrete Fourier transform method from DMOL3 calculations. Antioxidant activities of all the synthesized compounds were also investigated.  相似文献   

19.
In order to find informations on the native structure of the Leu-Enkephalin opiate peptide, the parent peptide and its two thioamide analogs (Thio-Gly2)Leu-Enkephalin and (Thio-Gly3)Leu-Enkephalin were studied by the theoretical method PEPSEA. This comparative conformational analysis showed that the active conformation is a β turn structure centered on Gly3 and Phe4. Moreover, this study showed also that the more active analog (Thio-Gly2)Leu-Enk has a lower tendency to adopt this structure. Consequently, its high activity can only be explained by its long lifetime due to its resistance to enzymatic hydrolysis, following the substitution of the amide linkage by the thioamide one. The weakly active analog (Thio-Gly3)Leu-Enk does not adopt this structure and prefers instead a β turn structure centered on Gly2 and Gly3. This study also confirmed the importance of the distances between the Tyr and Phe residues at positions 1 and 4, and that of the terminal Tyrosine N-H group which must be free of any intramolecular hydrogen bond in order to be available in the molecular recognition process.  相似文献   

20.
The question whether the emitter of yellow‐green firefly bioluminescence is the enol or keto‐constrained form of oxyluciferin (OxyLH2) still has no definitive answer from experiment or theory. In this study, Arg220, His247, adenosine monophosphate (AMP), Water324, Phe249, Gly343, and Ser349, which make the dominant contributions to color tuning of the fluorescence, are selected to simulate the luciferase (Luc) environment and thus elucidate the origin of firefly bioluminescence. Their respective and compositive effects on OxyLH2 are considered and the electronic absorption and emission spectra are investigated with B3LYP, B3PW91, and PBE1KCIS methods. Comparing the respective effects in the gas and aqueous phases revealed that the emission transition is prohibited in the gas phase but allowed in the aqueous phase. For the compositive effects, the optimized geometry shows that OxyLH2 exists in the keto(?1) form when Arg220, His247, AMP, Water324, Phe249, Gly343, and Ser349 are all included in the model. Furthermore, the emission maximum wavelength of keto(?1)+Arg+His+AMP+H2O+Phe+Gly+Ser is close to the experimental value (560 nm). We conclude that the keto(?1) form of OxyLH2 is a possible emitter which can produce yellow‐green bioluminescence because of the compositive effects of Arg220, His247, AMP, Water324, Phe249, Gly343, and Ser349 in the luciferase environment. Moreover, AMP may be involved in enolization of the keto(?1) form of OxyLH2. Water324 is indispensable with respect to the environmental factors around luciferin (LH2).  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