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 629 毫秒
1.
Deep-water morphs of lake charr, Salvelinus namaycush, are found, with one exception, in four of the largest lakes in the world: lakes Superior and Mistassini (QC) and Great Bear and Slave lakes. This paper advances a hypothesis for resource polymorphisms involving two types of deep-water morph, one of which is characteristic of the humper and the other of the siscowet charrs of Lake Superior. My hypothesis states that, first, the humper, or a humper-like morph, diverged postglacially in sympatry from the ancestral common (shallow-water) lake charr and became a feeding specialist on Mysis relicta. Second, in at least two of the four lakes the siscowet, or a siscowet-like charr, diverged as a feeding specialist on postglacially derived forms of deep-water ciscoes. In Lake Superior a successional process may have resulted in dominance of the siscowet at the expense of the humper charr. I concur with a previous inference that the one occurrence of a deep-water charr in a small lake (the above exception) represents emigration from Lake Superior. I further infer that this event involved an early humper charr, which implies that this morphotype had differentiated in Lake Superior in less than 1,900 year. I suggest that innate differences in plasticity, breeding behavior and assortive mating, and philopatry account for why Arctic charr isolate readily in small lakes whereas lake charr do not. My hypothesis assumes divergence of deep-water morphs occurred postglacially, an idea consistent with genetic and biogeographical evidence.  相似文献   

2.
Climate change is expected to alter species distributions and habitat suitability across the globe. Understanding these shifting distributions is critical for adaptive resource management. The role of temperature in fish habitat and energetics is well established and can be used to evaluate climate change effects on habitat distributions and food web interactions. Lake Superior water temperatures are rising rapidly in response to climate change and this is likely influencing species distributions and interactions. We use a three-dimensional hydrodynamic model that captures temperature changes in Lake Superior over the last 3 decades to investigate shifts in habitat size and duration of preferred temperatures for four different fishes. We evaluated habitat changes in two native lake trout (Salvelinus namaycush) ecotypes, siscowet and lean lake trout, Chinook salmon (Oncorhynchus tshawytscha), and walleye (Sander vitreus). Between 1979 and 2006, days with available preferred thermal habitat increased at a mean rate of 6, 7, and 5 days per decade for lean lake trout, Chinook salmon, and walleye, respectively. Siscowet lake trout lost 3 days per decade. Consequently, preferred habitat spatial extents increased at a rate of 579, 495 and 419 km2 per year for the lean lake trout, Chinook salmon, and walleye while siscowet lost 161 km2 per year during the modeled period. Habitat increases could lead to increased growth and production for three of the four fishes. Consequently, greater habitat overlap may intensify interguild competition and food web interactions. Loss of cold-water habitat for siscowet, having the coldest thermal preference, could forecast potential changes from continued warming. Additionally, continued warming may render more suitable conditions for some invasive species.  相似文献   

3.
Humans have played a significant role in reducing levels of genetic diversity and differentiation of many teleost fishes, leading to homogenization across biological entities. We compare patterns of historical and contemporary genetic structure for three sympatric Great Lake??s lake trout (Salvelinus namaycush) morphs (lean, siscowet, and humper) that differ in patterns of habitat occupancy, susceptibility to overfishing and predation by invasive sea lamprey (Petromyzon marinus). Differential susceptibilities to overfishing and predation were expected to result in different impacts to levels of genetic diversity loss for each morphotype. Genetic data was collected for samples at three points in time: 1948 (pre-collapse), 1959 (collapse) and 1990s (current), corresponding to periods of intensive fishing, mortality due to lamprey and recovery, respectively. The lean morph preferentially targeted by the fishery and recognized as highly preyed upon by sea lamprey was more highly impacted genetically than other morphs, as evidenced by greater loss of genetic diversity first during the period of overfishing, then during the period of high sea lamprey abundance once the fishery collapsed. The siscowet morph also experienced genetic bottlenecks during the period of overfishing (pre-collapse period). Results indicate significant levels of genetic differentiation among morphs historically prior to declines in abundance and also among contemporary populations, suggesting that periods of population decline and resurgence in abundance and distribution did not result in loss of genetic distinctiveness among morphs.  相似文献   

4.
At least two phenotypes of lake charr, Salvelinus namaycush, coexist in Lake Superior. A lean morph frequents the shallow inshore waters (< 50m) and the fat morph (siscowet) occupies the deeper offshore waters (50–250 m). The objective of this study was to determine if the elevated lipid concentration of siscowets reduces the costs of swimming in deep water. First, we modelled the effects of body composition (lipids) on the costs of swimming by lake charr, and then compared these theoretical results with empirical evidence obtained from Cesium 137-based estimates of food consumption, gross energy conversion, and swimming costs (activity multiplier). The attributes of growth, energy content (kJg-1), lipid concentrations, and Cesium 137 concentration (Bqg-1) were obtained from multimesh gillnet catches in eastern Lake Superior (1998 and 1999). The model showed that siscowet (fat) lake charr expended less energy than lean lake charr moving through the water column. Empirical evidence derived from Cesium 137 analysis confirmed that the activity multipliers of siscowets (fat) were less than those for lean charr. These findings support the view that the restoration of the fish community of the predominately deep water of the Great Lakes might be facilitated by the introduction of the fat phenotype.  相似文献   

5.
The restoration and rehabilitation of the native fish communities is a long-term goal for the Laurentian Great Lakes. In Lake Superior, the ongoing restoration of the native lake trout populations is now regarded as one of the major success stories in fisheries management. However, populations of the deepwater morphotype (siscowet lake trout) have increased much more substantially than those of the nearshore morphotype (lean lake trout), and the ecosystem now contains an assemblage of exotic species such as sea lamprey, rainbow smelt, and Pacific salmon (chinook, coho, and steelhead). Those species play an important role in defining the constraints and opportunities for ecosystem management. We combined an equilibrium mass balance model (Ecopath) with a dynamic food web model (Ecosim) to evaluate the ecological consequences of future alternative management strategies and the interaction of two different sets of life history characteristics for fishes at the top of the food web. Relatively rapid turnover rates occur among the exotic forage fish, rainbow smelt, and its primary predators, exotic Pacific salmonids. Slower turnover rates occur among the native lake trout and burbot and their primary prey—lake herring, smelt, deepwater cisco, and sculpins. The abundance of forage fish is a key constraint for all salmonids in Lake Superior. Smelt and Mysis play a prominent role in sustaining the current trophic structure. Competition between the native lake trout and the exotic salmonids is asymmetric. Reductions in the salmon population yield only a modest benefit for the stocks of lake trout, whereas increased fishing of lake trout produces substantial potential increases in the yields of Pacific salmon to recreational fisheries. The deepwater or siscowet morphotype of lake trout has become very abundant. Although it plays a major role in the structure of the food web it offers little potential for the restoration of a valuable commercial or recreational fishery. Even if a combination of strong management actions is implemented, the populations of lean (nearshore) lake trout cannot be restored to pre-fishery and pre-lamprey levels. Thus, management strategy must accept the ecological constraints due in part to the presence of exotics and choose alternatives that sustain public interest in the resources while continuing the gradual progress toward restoration. Received 10 December 1999; accepted 13 June 2000.  相似文献   

6.
Dietary fish must be assessed for benefits and risks to formulate risk management strategies. This article demonstrates that Laurentian Great Lakes (GL) freshwater species are good sources of omega-3 fatty acids using new data from a small sample (n = 7) of Lake Superior siscowet lake trout (Salvelinus namaycush siscowet) and five other GL fish species’ data. For Lake Superior (LS) siscowets, the saturates, mono-unsaturates, and poly-unsaturates composed 20.1, 40.7, and 39.1% of total lipid weight, respectively. Omega-3 poly-unsaturates (PUFAs) in these fish were more than twice the omega-6 (omega 3/6 ratio = 2.4). The LS lake trout data were combined with earlier LS data collected during the 1980s for eight other species and from five species of Lake Erie fish. All the GL freshwater species were compared with seven other published marine and freshwater fish studies from other global regions. PUFAs were compared based on latitude and marine versus freshwater origin. Differences between marine and freshwater species in omega-3 fatty acid were less at higher latitudes. GL freshwater fish species can be a good source of beneficial fats like marine fish and must be accounted in effective risk communications involving persistent bioaccumulative toxicants in dietary fish.  相似文献   

7.
N-(Morpholinothiocarbonyl) benzamide (C(12)H(14)N(2)O(2)S) and N-(piperidylthiocarbonyl) benzamide (C(13)H(16)N(2)OS) and their Co(III) complexes have been synthesized and characterized by elemental analysis, FTIR and NMR methods. The complex Co(C(12)H(14)N(2)O(2)S)(3), crystallizes in the triclinic space group P1, with Z=2, and unit cell parameters, a=12.080(7)A, b=12.195(7)A, c=13.025(6)A, alpha=90.198(7) degrees, beta=95.721(7) degrees, gamma=106.426(9) degrees, V=1830.4(17)A(3). The antifungal activity against the major pathogens responsible for important plant diseases (Botrytis cinerea, Myrothecium and Verticillium dahliae dleb) of N-(piperidylthiocarbonyl) benzamide and its complex with cobalt (III) are studied and compared with N-(morpholinothiocarbonyl) benzamide.  相似文献   

8.
Ex vivo ?(13)C, (2)H? NMR spectroscopy allowed to estimate the relative sizes of neuronal and glial glutamate pools and the relative contributions of (1-(13)C) glucose and (2-(13)C, 2-(2)H(3)) acetate to the neuronal and glial tricarboxylic acid cycles of the adult rat brain. Rats were infused during 60 min in the right jugular vein with solutions containing (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose or (2-(13)C, 2-(2)H(3)) acetate only. At the end of the infusion the brains were frozen in situ and perchloric acid extracts were prepared and analyzed by high resolution (13)C NMR spectroscopy (90.5 MHz). The relative sizes of the neuronal and glial glutamate pools and the contributions of acetyl-CoA molecules derived from (2-(13)C, (2)H(3)) acetate or (1-(13)C) glucose entering the tricarboxylic acid cycles of both compartments, could be determined by the analysis of (2)H-(13)C multiplets and (2)H induced isotopic shifts observed in the C4 carbon resonances of glutamate and glutamine. During the infusions with (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose, the glial glutamate pool contributed 9% of total cerebral glutamate being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (4%), (2-(13)C) acetyl-CoA (3%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (2%). The neuronal glutamate pool accounted for 91% of the total cerebral glutamate being mainly originated from (2-(13)C) acetyl-CoA (86%) and (2-(13)C, 2-(2)H) acetyl-CoA (5%). During the infusions of (2-(13)C, 2-(2)H(3)) acetate only, the glial glutamate pool contributed 73% of the cerebral glutamate, being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (36%), (2-(13)C, 2-(2)H) acetyl-CoA (27%) and (2-(13)C) acetyl-CoA (10%). The neuronal pool contributed 27% of cerebral glutamate being formed from (2-(13)C) acetyl-CoA (11%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (16%). These results illustrate the potential of ?(13)C, (2)H? NMR spectroscopy as a novel approach to investigate substrate selection and metabolic compartmentation in the adult mammalian brain.  相似文献   

9.
Nan YH  Shin SY 《BMB reports》2011,44(11):747-752
To investigate the effects of disulphide bond position on the salt resistance and lipopolysaccharide (LPS)-neutralizing activity of α-helical homo-dimeric antimicrobial peptides (AMPs), we synthesized an α-helical model peptide (K6L4W1) and its homo-dimeric peptides (di-K(6)L(4)W(1)-N, di-K(6)L(4)W(1)-M, and di-K(6)L(4)W(1)-C) with a disulphide bond at the N-terminus, the central position, and the C-terminus of the molecules, respectively. Unlike (6)L(4)W(1) and di-K(6)L(4)W(1)-M, the antimicrobial activity of di-K(6)L(4)W(1)-N and di-K(6)L(4)W(1)-C was unaffected by 150 mM NaCl. Both di-K(6)L(4)W(1)-N and di-K(6)L(4)W(1)-C caused much greater inhibitory effects on nitric oxide (NO) release in LPS-induced mouse macrophage RAW 264.7 cells, compared to di-K(6)L(4)W(1)-M. Taken together, our results indicate that the presence of a disulphide bond at the N- or C-terminus of the molecule, rather than at the central position, is more effective when designing salt-resistant α-helical homo-dimeric AMPs with potent antimicrobial and LPS-neutralizing activities. [BMB reports 2011; 44(11): 747-752].  相似文献   

10.
The carboxyl-terminal domain of RNA polymerase II, which is rich in phosphorylation sites, contains 17--52 tandem repeats with the consensus sequence of the heptapeptide, YSPTSPS. The repeat unit of the heptapeptide has two SPXX motifs showing potential beta-turns, SPTS and SPSY. NMR studies were performed in water at pH 4.0 for two cyclic peptides containing one and two repeat units, cyclo-[C(1)R(2)D(3)Y(4)S(5)P(6)T(7)S(8)P(9)S(10)Y(11)S(12)R(13)D(14)C(15)] (peptide 1) and cyclo-[C(1)R(2)D(3)Y(4)S(5)P(6)T(7)S(8)P(9)S(10)Y(11)S(12)P(13)T(14)S(15)P(16)N(17)Y(18)S(19)R(20)D(21)C(22)] (peptide 2), which are cyclized with a disulfide bridge of two Cys residues at the N- and C-termini. SP in 1 and 2 are predominantly in trans form. The following NMR parameters were detected: (1) lower temperature coefficients of amide proton chemical shifts of T7 and S8 in 1, and Tx (T7 or T14), Sx (S8 or S15), Tz (T14 or T7) and Sz (S15 or S8) in 2, (2) significantly large deviation of H(alpha) chemical shifts from its random coil value (Delta H(alpha)) of Pro preceding the Thr (P6 in 1, and Px and Pz in 2), (3) relatively large (3)J(HNH alpha) coupling constants (>8.7 Hz) of T7 in 1 and Tx and Tz in 2, and (4) NOE (d(NN) (i, i+1)) connectivities between the amide protons of T7-S8 and S10-Y11 in 1, and Tx-Sx, S10-Y11, Tz-Sz, and N17-Y18 in 2, although two Pro-Thr-Ser segments in 2 (each of these are annotated by 'x' and 'z') in the first and second repeat units were not distinguishable. Comparison of the NMR parameters between the cyclic peptides and the corresponding linear peptides indicates that cyclization promotes structural stabilization in water. The present NMR data were consistent with the presence of a beta-turn at both SPTS and SPSY: S(5)P(6)T(7)S(8) and S(8)P(9)S(10)Y(11) in 1, and SPxTxSx, SPzTzSz, SP(9)S(10)Y(11), SP(16)N(17)Y(18) in 2. However, the structure of the SPTS segment is more stable than that of the SPSY segment. Conformations consistent with NMR parameters including NOE distances were obtained through molecular dynamics and energy minimization methods. These calculations yielded two stable conformers for the SPTS segment. One of the two corresponds to a type I beta-turn.  相似文献   

11.
Five new organotin(IV) molecules with the heterocyclic thioamides; 2-mercaptobenzothiazole (Hmbzt), 5-chloro-2-mercaptobenzothiazole (Hcmbzt), 3-methyl-2-mercaptobenzothiazole (mmbzt) and 2-mercaptonicotinic acid (H(2)mna) of formulae [(n-C(4)H(9))(2)Sn(mbzt)(2)] (1), [(C(6)H(5))(2)Sn(mbzt)(2)] (2), [(CH(3))(2)Sn(cmbzt)(2)].1.7(H(2)O)] (3), [(n-C(4)H(9))(2)SnCl(2)(mmbzt)(2).(CH(2)Cl(2))] (4) and [[(C(6)H(5))(3)Sn](2)(mna).[(CH(3))(2)CO]] (5) have been synthesized and characterized by elemental analysis, 1H-, 13C-NMR, FT-IR and M?ssbauer spectroscopic techniques. Crystal structures of molecules 1, 3 and 5 have been determined by X-ray diffraction at 173(1) K (1 and 5) and 293(2) K (3). Compound 1 C(22)H(26)N(2)S(4)Sn, is monoclinic, space group C2/c, a=44.018(2), b=8.8864(5), c=12.8633(7) A, beta=104.195(5) degrees, Z=8. Compound 3 is also monoclinic, space group P2(1)/c and a=17.128(2) A, b=17.919(2) A, c=7.3580(10) A, beta=98.290(10) degrees, Z=4. In both molecules 1 and 3, two carbon atoms from aryl groups, two sulfur and two nitrogen atoms from thione ligands form a distorted octahedral geometry around tin(IV) with trans-C(2), cis-N(2), cis-S(2) configurations. Compound 5 C(45)H(39)NO(3)SSn(2) is monoclinic, space group P2(1)/n, a=9.1148(2) A, b=29.2819(6), c=15.5556(4) A, beta=106.2851(9) degrees, Z=4. Complex 5 contains two [(C(6)H(5))(3)Sn(IV)] moieties linked by a double deprotonated 2-mercaptonicotinic acid (H(2)mna). Both tin(IV) ions are five coordinated. This complex is the an example of a pentacoordinated Ph(3)SnXY system with an axial-equatorial arrangement of the phenyl groups at Sn(1) atom. Compounds 1, 3 and 5 were tested for in vitro cytotoxicity against the cancer cell line of sarcoma cells (mesenchymal tissue) from the Wistar rat, polycyclic aromatic hydrocarbons (benzo[a]pyrene) carcinogenesis. Compound 5 exhibits strong cytotoxic activity, while complexes 1 and 3 show less cytotoxic activity.  相似文献   

12.
We report different analytical methods used to study the effects of 3\'-azido-3\'-deoxythymidine, aspirin, taxol, cisplatin, atrazine, 2,4-dichlorophenoxyacetic, biogenic polyamines, chlorophyll, chlorophyllin, poly(ethylene glycol), vanadyl cation, vanadate anion, cobalt-hexamine cation, and As2O3, on the stability and secondary structure of human serum albumin (HSA) in aqueous solution, using capillary electrophoresis, Fourier transform infrared, ultraviolet visible, and circular dichroism (CD) spectroscopic methods. The concentrations of HSA used were 4% to 2% or 0.6 to 0.3 mM, while different ligand concentrations were 1 microM to 1 mM. Structural data showed drugs are mostly located along the polypeptide chains with both specific and nonspecific interactions. The stability of drug-protein complexes were in the order K(VO(2+)) 1.2 x 10(8) M(-1) > K(AZT) 1.9 x 10(6) M(-)1 > K(PEG) 4.1 x 10(5) M(-1) > K(atrazine) 3.5 x 10(4) M(-1) > K(chlorophyll) 2.9 x 10(4) M(-1) > K2,4-D 2.5 x 10(4) M-1 > K(spermine) 1.7 x 10(4) M(-1) > K(taxol) 1.43 x 10(4) M(-1) > K(Co(3+)) > 1.1 x 10(4) M(-1) > K(aspirin) 1.04 x 10(4)i(-1) > K(chlorophyllin) 7.0 x 10(3) M(-1) > K(VO(3)(-)) 6.0 x 103 M(-1) > K(spermidine) 5.4 x 10(3) M(-1) > K(putrescine) 3.9 x 10(3) M(-1) > K(As(2)O(3)) 2.2 x 10(3) M(-1)> K(cisplatin) 1.2 x 10(2) M(-1). The protein conformation was altered (infrared and CD results) with major reduction of alpha-helix from 60 to 55% (free HSA) to 49 to 40% and increase of beta-structure from 22 to 15% (free HSA) to 33 to 23% in the drug-protein complexes. The alterations of protein secondary structure are attributed to a partial unfolding of HSA on drug complexation.  相似文献   

13.
The binding properties of a spacer-linked synthetic Sd(a) tetrasaccharide beta-D-GalpNAc-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (1), two tetrasaccharide mimics beta-D-Galp-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (2) and beta-D-GlcpNAc-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (3), and two trisaccharide mimics beta-D-GalpNAc-(1-->4)-3-O-(SO(3)H)-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (4) and beta-D-GalpNAc-(1-->4)-3-O-(CH(2)COOH)-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (5) with lectins from Dolichos biflorus (DBL), Maackia amurensis (MAL), Phaseolus limensis (PLL), Ptilota plumosa (PPL), Ricinus communis 120 (RCL120) and Triticum vulgaris (wheat germ agglutinin, WGA) have been investigated by surface plasmon resonance (SPR) detection. MAL, PPL, RCL120 and WGA did not display any binding activity with compounds 1-5. However, DBL and PLL, both exhibiting GalNAc-specificity, showed strong binding activity with compounds 1, 4 and 5, and 1, 3, 4 and 5, respectively. The results demonstrate that SPR is a very useful analysis system for identifying biologically relevant oligosaccharide mimics of the Sd(a) determinant.  相似文献   

14.
The G-protein G(salpha) exists in three isoforms, the G(salpha) splice variants G(salphashort) (G(salphaS)) and G(salphalong) (G(salphaL)), and the G-protein G(alphaolf) that is not only involved in olfactory signaling but also in extrapyramidal motor regulation. Studies with beta(2)-adrenoceptor (beta(2)AR)-G(salpha) fusion proteins showed that G(salpha) proteins activate adenylyl cyclase (AC) in the order of efficacy G(salphaS)>G(salphaL) approximately G(alphaolf) and that G(salpha) proteins confer the hallmarks of constitutive activity to the beta(2)AR in the order of efficacy G(salphaL)>G(alphaolf)>G(salphaS). However, it is unclear whether such differences between G(salpha) proteins also exist in the nonfused state. In the present study, we co-expressed the beta(2)AR and dopamine D(1)-receptor (D(1)R) with G(salpha) proteins at different ratios in Sf9 insect cells. In agreement with the fusion protein studies, nonfused G(alphaolf) was less efficient than nonfused G(salphaS) and G(salphaL) at activating AC, but otherwise, we did not observe differences between the three G(salpha) isoforms. Thus, it is much easier to dissect differences between G(salpha) isoforms using beta(2)AR-G(salpha) fusion proteins than nonfused G(salpha) isoforms.  相似文献   

15.
UV and NMR spectroscopy were employed to study the tautomerism, acid-base properties and conformation of the exocyclic N(4)-OH group in 1-methyl-N(4)-hydroxycytosine (1-mOH(4)C), and its methyl derivatives, viz. the fixed imino forms (1,3-m(2)OH(4)C and 1,3,5-m(3)OH(4)C), the fixed amino form (1,N(4)-m(2)OH(4)C), and analogues sterically constrained to the form syn (1,5-m(2)OH(4)C) or anti (1,3-m(2)OH(4)C) with respect to the ring N(3). Relative to 1,N(4)-m(2)OH(4)C, UV spectroscopy showed that the other analogues were predominantly imino and that all analogues formed a structurally common cation in acid medium, with results pointing to approximately 90% population of the imino species for 1-mOH(4)C and 1,5-m(2)OH(4)C, further supported by NMR spectroscopy. Both exhibited two sequential dissociations in alkaline medium, the first due to N(4)-OH, followed by the N(3)-H. (1)H and (13)C NMR spectroscopy showed 1-mOH(4)C in the conformation syn. With 1,3,5-m(3)OH(4)C, an ;overcrowded' planar molecule with steric constraints to both the syn and anti conformations, a syn-anti equilibrium is observed, with a preference of approximately 75% for the anti rotamer, independently of the polarity of the medium. Exchange between the rotamers is slow on the NMR time-scale, with a minimal barrier to exchange exceeding 100 kJ/mol. In low-polar media, the analogues associate as dimers via O(4)-Hcdots, three dots, centeredO(2) or O(4)-Hcdots, three dots, centeredN(4) hydrogen bonds, with association constants at ambient temperature of 4.6 (1,3-m(2)OH(4)C), 12.8 (anti 1,3,5-m(3)OH(4)C), 36 (1,5-m(2)OH(4)C), 109 (syn 1,3,5-m(3)OH(4)C) M(-1). Implications of the overall findings to the promutagenic activities of OH(4)C and OMe(4)C are examined.  相似文献   

16.
The binding of carbon dioxide by horse haemoglobin   总被引:15,自引:7,他引:8  
1. Three modified horse haemoglobins have been prepared: (i) alpha(c) (2)beta(c) (2), in which both the alpha-amino groups of the alpha- and beta-chains have reacted with cyanate, (ii) alpha(c) (2)beta(2), in which the alpha-amino groups of the alpha-chains have reacted with cyanate, and (iii) alpha(2)beta(c) (2), in which the two alpha-amino groups of the beta-chain have reacted with cyanate. 2. The values of n (the Hill constant) for alpha(c) (2)beta(c) (2), alpha(2)beta(c) (2) and alpha(c) (2)beta(2) were (respectively) 2.5, 2.0 and 2.6, indicating the presence of co-operative interactions between the haem groups for all derivatives. 3. In the alkaline pH range (about pH8.0) all the derivatives show the same charge as normal haemoglobin whereas in the acid pH range (about pH6.0) alpha(c) (2)beta(c) (2) differs by four protonic charges and alpha(c) (2)beta(2), alpha(2)beta(c) (2) by two protonic charges from normal haemoglobin, indicating that the expected number of ionizing groups have been removed. 4. alpha(c) (2)beta(2) and alpha(c) (2)beta(c) (2) show a 25% decrease in the alkaline Bohr effect, in contrast with alpha(2)beta(c) (2), which has the same Bohr effect as normal haemoglobin. 5. The deoxy form of alpha(c) (2)beta(c) (2) does not bind more CO(2) than the oxy form of alpha(c) (2)beta(c) (2), whereas alpha(c) (2)beta(2) and alpha(2)beta(c) (2) show intermediate binding. 6. The results reported confirm the hypothesis that, under physiological conditions, haemoglobin binds CO(2) through the four terminal alpha-amino groups and that the two terminal alpha-amino groups of alpha-chains are involved in the Bohr effect.  相似文献   

17.
Shi YB  Fang JL  Liu XY  Du L  Tang WX 《Biopolymers》2002,65(2):81-88
The secondary structures of porcine brain Cu(4)Zn(3)-metallothionein (MT)-III and Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I from rabbit livers in the solid state are investigated by Fourier transform IR spectroscopy (FTIR) and Fourier transform Raman spectroscopy (FT-Raman). The Cu(4)Zn(3)MT-III contains 26-28% beta-turns and half-turns, 13-14% 3(10)-helices, 47-49% random coils, and 11-12% beta-extended chains. The structural comparison of porcine brain Cu(4)Zn(3)MT-III with rabbit liver Cd(5)Zn(2)MT-I (II) and Zn(7)MT-I shows that the contents of the random coil structure are obviously increased. The results indicate that the insert of an acidic hexapeptide in the alpha domain of Cu(4)Zn(3)MT-III possibly forms an alpha helix. However, because the bands assigned to the alpha-helix and random coil structures are overlapped in the spectra, the content of random coil structures in Cu(4)Zn(3)MT-III is therefore higher than those in Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I.  相似文献   

18.
载脂蛋白E基因多态性与2型 糖尿病脑梗死的关系   总被引:1,自引:1,他引:0  
研究载脂蛋白E (ApoE) 基因多态性与中国东北汉族2型糖尿病合并脑梗死的关系。采用聚合酶链反应-限制性片段长度多态性(PCR-RFLP)技术,检测了208例个体的ApoE基因多态性,其中对照(CON)组69例,2型糖尿病无大血管病变(T2DM)组67例和2型糖尿病合并脑梗死(T2DMCI)组72例;同时测定了其中70例T2DMCI患者的血脂含量。CON组等位基因频率为:ε2 9.6%、ε3 82.4%、ε4 8.1%,基因型频率为:ε2ε3 13.2%、ε3ε367.6%、ε3ε416.2%;T2DM组等位基因频率为:ε2 10.5%、ε3 84.3%、ε4 5.2%,基因型频率为:ε2ε3 19.4%、ε3ε370.1%、ε3ε49%;T2DMCI组等位基因频率为:ε2 11.8%、ε3 84.7%、ε4 3.5%,基因型频率为:ε2ε3 15.2%、ε3ε375%、ε3ε44.2%。3组间等位基因和基因型频率的差异经检验无统计学意义。T2DMCI患者中各基因型之间的血浆总胆固醇 (TC)、甘油三酯 (TG)、高密度脂蛋白胆固醇(HDL-C)和低密度脂蛋白胆固醇(LDL-C)水平亦无显著性差异。在中国东北汉族人群中,未发现ApoE基因多态性与T2DMCI之间存在关联,亦未发现ApoE基因多态性与T2DMCI患者的TC、TG、 HDL-C和 LDL-C水平之间存在关联。Abstract: In order to explore the association of apolipoprotein E (ApoE) gene polymorphism with cerebral infarction in type 2 diabetic patients of Han nationality in Northeast China , the genotypes of ApoE gene were analyzed by polymerase chain reaction –restriction fragment length polymorphism (PCR-RFLP) in the 208 cases, including 69 cases in control (CON) group and 67 in type 2 diabetes mellitus (T2DM) group as well as 72 in type 2 diabetes mellitus with cerebral infarction (T2DMCI) group. Plasma lipid content in T2DMCI was also detected for 70 cases. The distribution of genotypes in ApoE gene,ε2ε3、ε3ε3 as well asε3ε4 was no significant difference in three groups (ε2ε3 : 13.2%、ε3ε3 : 67.6%、ε3ε4 : 16.2%in CON group;ε2ε3 : 19.4%、ε3ε3: : 70.1%ε3ε4 : 9%in T2DM group;ε2ε3 : 15.2%、ε3ε3 : 75%、ε3ε4 : 4.2%in T2DMCI group).The allele frequencies ofε2、ε3 andε4 were not significantly different in the three groups, either (ε2 : 9.6%、ε3 : 82.4%、ε4 : 8.1%in CON group; ε2 :10.5%、ε3 :84.3%、ε4 : 5.2%in T2DM group; ε2 :11.8%、ε3 :84.7%、ε4 : 3.5%in T2DMCI group). The levels of total cholesterol (TC), tryglyceride (TG), high density lipoprotein-cholesterol (HDL-C) and low density lipoprotein-cholesterol (LDL-C) were not significantly different among the different genotypes in T2DMCI group. The study confirmed that the polymorphisms of ApoE gene are neither associated with the T2DMCI, nor with the levels of plasma lipid in T2DMCI.  相似文献   

19.
Stereoselective, total synthesis of O-alpha-D-galactopyranosyl-(1----4) -O-beta-D-galactopyranosyl-(1----4)-O-beta-D-glucopyranosyl-(1----1)-N -tetracosanoyl-[2S,3R,4E (and 4Z)]-sphingenine and O-alpha-D -galactopyranosyl-(1----3)-O-beta-D-galactopyranosyl-(1----4)-O-beta-D -glucopyranosyl-(1----1)-N-tetracosanoyl-(2S,3R,4E)-sphin gen ine was achieved by using O-(2,3,4,6-tetra-O-acetyl-alpha-D-galactopyranosyl) -(1----4)-O-(2,3,6-tri-O-acetyl-beta-D-galactopyranosyl)-(1----4)-2,3,6- tri-O-acetyl-alpha-D-glucopyranosyl trichloroacetimidate, O-(2,3,4,6-tetra-O-acetyl-alpha-D-galactopyranosyl) -(1----4)-O-(2,3,6-tri-O-acetyl-beta-D-galactopyranosyl)-(1----4)-2,3,6- tri-O-acetyl-alpha (and beta)-D-glucopyranosyl fluoride, and O-(2,3,4,6-tetra-O-acetyl-alpha-D -galactopyranosyl)-(1----3)-O-(2,3,6-tri-O-acetyl-beta-D-galactopyran osyl)-(1----4)-2,3,6-tri-O-acetyl-alpha-D-glucopyranosyl trichloroacetimidate.  相似文献   

20.
alpha-Melanocyte stimulating hormone (alphaMSH), Ac-Ser(1)-Tyr(2)-Ser(3)-Met(4)-Glu(5)-His(6)-Phe(7)-Arg(8)-Trp(9)-Gly(10)-Lys(11)-Pro(12)-Val(13)-NH(2), is an endogenous agonist for the melanocortin receptor 1 (MC1R), the receptor found in the skin, several types of immune cells, and other peripheral sites. Three-dimensional models of complexes of this receptor with alphaMSH and its synthetic analog NDP-alphaMSH, Ac-Ser(1)-Tyr(2)-Ser(3)-Nle(4)-Glu(5)-His(6)-D-Phe(7)-Arg(8)-Trp(9)-Gly(10)-Lys(11)-Pro(12)-Val(13)-NH(2), have been previously proposed. In those models, the 6-9 segment of the ligand was considered essential for the ligand-receptor interactions. In this study, we probed the role of Trp(9) of NDP-alphaMSH in interactions with hMC1bR. Analogs of NDP-alphaMSH with various amino acids in place of Trp(9) were synthesized and tested in vitro in receptor affinity binding and cAMP functional assays at human melanocortin receptors 1b, 3, 4, and 5 (hMC1b,3-5R). Several new compounds displayed high agonist potency at hMC1bR (EC(50) = 0.5-5 nM) and receptor subtype selectivity greater than 2000-fold versus hMC3-5R. The Trp(9) residue of NDP-alphaMSH was determined to be not essential for molecular recognition at hMC1bR.  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