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 875 毫秒
1.
Six peptides have been isolated and characterized from the dorsal glands of the tree frog Litoria genimaculata. One of these is the known hypotensive peptide caerulein; the others have been named maculatins. The amino acid sequences of the maculatin peptides have been determined using a combination of fast atom bombardment mass spectrometry and automated Edman sequencing. Four of the maculatin peptides show antibiotic activity, with maculatin 1.1 [GLFGVLAKVAAHVVPAIAEHF(NH2;)] showing the most pronounced activity, particularly against Gram-positive organisms. Maculatin 1.1 resembles the known caerin 1 antibiotic peptides, except that four of the central amino acid residues (of the caerin 1 system) are missing in maculatin 1.1. A comparison of the antibiotic activity of maculatin 1.1 with those of caerin 1.1 is reported. ©1998 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Many species of amphibians in the wet tropics of Australia have experienced population declines linked with the emergence of a skin-invasive chytrid fungus, Batrachochytrium dendrobatidis. An innate defense, antimicrobial peptides produced by granular glands in the skin, may protect some species from disease. Here we present evidence that supports this hypothesis. We tested ten synthesized peptides produced by Australian species, and natural peptide mixtures from five Queensland rainforest species. Natural mixtures and most peptides tested in isolation inhibited growth of B. dendrobatidis in vitro. The three most active peptides (caerin 1.9, maculatin 1.1, and caerin 1.1) were found in the secretions of non-declining species (Litoria chloris, L. caerulea, and L. genimaculata). Although the possession of a potent isolated antimicrobial peptide does not guarantee protection from infection, non-declining species (L. lesueuri and L. genimaculata) inhabiting the rainforest of Queensland possess mixtures of peptides that may be more protective than those of the species occurring in the same habitat that have recently experienced population declines associated with chytridiomycosis (L. nannotis, L. rheocola, and Nyctimystes dayi). This study demonstrates that in vitro effectiveness of skin peptides correlates with the degree of decline in the face of an emerging pathogen. Further research is needed to assess whether this non-specific immune defense may be useful in predicting disease susceptibility in other species.  相似文献   

3.
Five healthy adult female first-generation hybrid tree frogs were produced by interspecific breeding of closely related tree frogs Litoria splendida and L. caerulea in a cage containing large numbers of males and females of both species. Phylogenetic analysis of mitochondrial DNA sequences established the female parent to be L. splendida. The peptide profile of the hybrid frogs included the neuropeptide caerulein, four antibiotics of the caerin 1 family and several neuronal nitric oxide synthase inhibitors of the caerin 1 and 2 classes of peptides. The skin secretions of the hybrids contained some peptides common to only one parent, some produced by both parental species, and four peptides expressed by the hybrids but not the parental species.  相似文献   

4.
It is generally accepted that high quality males are those that succeed in male–male competition: in either aggression or rivalry to attract and be selected by females. Previous studies of amphibians have suggested that the main characters influencing male mating success include variation in call characteristics (e.g. call rate, call intensity), calling behaviour, body condition, age and chorus tenure. In the present paper, several of the characters influencing female mate choice (male body size, body condition, call rate, call frequency and chorus tenure) are investigated in two closely related, explosive breeding frog species Litoria chloris and Litoria xanthomera. Smaller males of both species are shown to be more successful than larger males and this success is attributed to the increased chorus tenure of smaller males in L. xanthomera. This increased chorus tenure was attributed to the lower total energy used per call by a small male calling at a higher frequency. Whether increased chorus tenure explains female mate choice in L. chloris is uncertain but is highly probable given the strong similarity between the two species in both ecology and call characteristics.  相似文献   

5.
Nineteen citropin peptides are present in the secretion from the granular dorsal glands of the Blue Mountains tree-frog Litoria citropa; 15 of these peptides are also present in the secretion from the submental gland. Two major peptides, citropin 1.1 (GLFDVIKKVASVIGGL-NH2), citropin 1.2 (GLFDIIKKVASVVGGL-NH2) and a minor peptide, citropin 1.3 (GLFDIIKKVASVIGGL-NH2) are wide-spectrum antibacterial peptides. The amphibian has an endoprotease which deactivates these membrane-active peptides by removing residues from the N-terminal end: loss of three residues gives the most abundant degradation products. The solution structure of the basic peptide citropin 1.1 has been determined by NMR spectroscopy [in a solvent mixture of trifluoroethanol/water (1 : 1)] to be an amphipathic alpha-helix with well-defined hydrophobic and hydrophilic regions. The additional four peptides produced by the dorsal glands are structurally related to the antibacterial citropin 1 peptides but contain three more residues at their C-terminus [e.g. citropin 1.1.3 (GLFDVIKKVASVIGLASP-OH)]. These peptides show minimal antibacterial activity; their role in the amphibian skin is not known.  相似文献   

6.
Caerin 1.1 is a potent broad-spectrum antibacterial peptide isolated from a number of Australian frogs of the Litoria genus. In membrane-like media, this peptide adopts two alpha-helices, separated by a flexible hinge region bounded by Pro15 and Pro19. Previous studies have suggested that the hinge region is important for effective orientation of the two helices within the bacterial cell membrane, resulting in lysis via the carpet mechanism. To evaluate the importance of the two Pro residues, they were replaced with either Ala or Gly. The antibacterial activity of these two peptides was tested, and their three-dimensional structures were determined using two-dimensional NMR spectroscopy and restrained molecular dynamics calculations. The resulting structures indicate that the central hinge angle decreases significantly upon replacement of the Pro residues with Gly and to a further extent with Ala. This trend was mirrored by a corresponding decrease in antibiotic activity, further exemplifying the necessity of the hinge in caerin 1.1 and related peptides. In a broader context, the use of Pro, Gly, and Ala variants of caerin 1.1 has enabled the relationship between conformational flexibility and activity to be directly investigated in a systematic manner.  相似文献   

7.
Two neuropeptides have been isolated and identified from the secretions of the skin glands of the Stony Creek Frog Litoria lesueuri. The first of these, the known neuropeptide caerulein 1.1, is a common constituent of anuran skin secretions, and has the sequence pEQY(SO3)TGWMDF-NH2. This neuropeptide is smooth muscle active, an analgaesic more potent than morphine and is also thought to be a hormone. The second neuropeptide, a new peptide, has been named lesueurin and has the primary structure GLLDILKKVGKVA-NH2. Lesueurin shows no significant antibiotic or anticancer activity, but inhibits the formation of the ubiquitous chemical messenger nitric oxide from neuronal nitric oxide synthase (nNOS) at IC(50) (16.2 microm), and is the first amphibian peptide reported to show inhibition of nNOS. As a consequence of this activity, we have tested other peptides previously isolated from Australian amphibians for nNOS inhibition. There are three groups of peptides that inhibit nNOS (IC(50) at microm concentrations): these are (a) the citropin/aurein type peptides (of which lesueurin is a member), e.g. citropin 1.1 (GLFDVIKKVASVIGGL-NH(2)) (8.2 microm); (b) the frenatin type peptides, e.g. frenatin 3 (GLMSVLGHAVGNVLG GLFKPK-OH) (6.8 microm); and (c) the caerin 1 peptides, e.g. caerin 1.8 (GLFGVLGSIAKHLLPHVVPVIAEKL-NH(2)) (1.7 microm). From Lineweaver-Burk plots, the mechanism of inhibition is revealed as noncompetitive with respect to the nNOS substrate arginine. When the nNOS inhibition tests with the three peptides outlined above were carried out in the presence of increasing concentrations of Ca(2+) calmodulin, the inhibition dropped by approximately 50% in each case. In addition, these peptides also inhibit the activity of calcineurin, another enzyme that requires the presence of the regulatory protein Ca(2+) calmodulin. It is proposed that the amphibian peptides inhibit nNOS by interacting with Ca(2+)calmodulin, and as a consequence, blocks the attachment of this protein to the calmodulin domain of nNOS.  相似文献   

8.
The skin secretions of female and male Litoria splendida have been monitored monthly over a three-year period using HPLC and electrospray mass spectrometry. Two minor peptides are present only in the skin secretion of the male. The first of these is the female-attracting aquatic male sex pheromone that we have named splendipherin, a 25 amino acid peptide (GLVSSIGKALGGLLADVVKSKGQPA-OH). This pheromone constitutes about 1% of the total skin peptides during the breeding season (January to March), dropping to about 0.1% during the period June to November. Splendipherin attracts the female in water at a concentration of 10-11-10-9 M, and is species specific. The second peptide is a wide-spectrum antibiotic of the caerin 1 group, a 25 residue peptide (GLLSVLGSVAKHVLPHVVPVIAEKL-NH2) named caerin 1.10. The neuropeptides of L. splendida are also seasonally variable, the change identical for both the female and male. During the period October to March, the sole neuropeptide present in skin secretions is caerulein [pEQDY(SO3)TGWMDF-NH2]; this is active on smooth muscle and is also an analgaesic. During the southern winter (June to September), more than half of the caerulein is hydrolysed to [pEQDYTGWMDF-NH2], a peptide that shows no smooth muscle activity. In place of caerulein, a new peptide, Phe8 caerulein [pEQDY(SO3)TGWFDF-NH2], becomes a major component of the skin secretion. Perhaps this seasonal change is involved in thermoregulation, that is, with the initiation and maintenance of the inactive (hibernation) phase of the animal.  相似文献   

9.
10.
Previously Os, a 22 amino acid sequence of a defensin from the soft tick Ornithodoros savignyi, was found to kill Gram‐positive and Gram‐negative bacteria at low micromolar concentrations. In this study, we evaluated synthetic peptide analogues of Os for antibacterial activity with an aim to identify minimalized active peptide sequences and in so doing obtain a better understanding of the structural requirements for activity. Out of eight partially overlapping sequences of 10 to 12 residues, only Os(3–12) and Os(11–22) exhibit activity when screened against Gram‐positive and Gram‐negative bacteria. Carboxyamidation of both peptides increased membrane‐mediated activity, although carboxyamidation of Os(11–22) negatively impacted on activity against Staphylococcus aureus. The amidated peptides, Os(3–12)NH2 and Os(11–22)NH2, have minimum bactericidal concentrations of 3.3 μM against Escherichia coli. Killing was reached within 10 minutes for Os(3–12)NH2 and only during the second hour for Os(11–22)NH2. In an E. coli membrane liposome system, both Os and Os(3–12)NH2 were identified as membrane disrupting while Os(11–22)NH2 was less active, indicating that in addition to membrane permeabilization, other targets may be involved in bacterial killing. In contrast to Os, the membrane disruptive effect of Os(3–12)NH2 did not diminish in the presence of salt. Neither Os nor its amidated derivatives caused human erythrocyte haemolysis. The contrasting killing kinetics and effects of amidation together with structural and liposome leakage data suggest that the 3–12 fragment relies on a membrane disruptive mechanism while the 11–22 fragment involves additional target mechanisms. The salt‐resistant potency of Os(3–12)NH2 identifies it as a promising candidate for further development.  相似文献   

11.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Eight peptides with differential growth–inhibitory activity against the gram-positive bacterium Staphylococcus aureus, the gram-negative bacterium Escherichia coli and the yeast, Candida albicans were isolated from an extract of the skin of the North American pig frog Rana grylio. The primary structures of these antimicrobial peptides were different from previously characterized antimicrobial peptides from Ranid frogs but on the basis of sequence similarities, the peptides may be classified as belonged to four previously characterized peptide families: the ranatuerin-1, ranatuerin-2 and ranalexin families, first identified in the North American bullfrog, Rana catesbeiana, and the temporin family first identified in the European common frog Rana temporaria. Peptides belonging to the brevinin-1, brevinin-2, esculentin-1, and esculentin-2 families, previously isolated from the skins of other species of Ranid frogs, were not identified in the extracts. The ranatuerin-1 and ranalexin peptides showed broadest spectrum of antimicrobial activity whereas the temporins were active only against S. aureus. Synthetic replicates of temporin-1Gb (SILPTIVSFLSKFL.NH2) and temporin-1Gd (FILPLIASFLSKFL.NH2) produced concentration-dependent relaxation of preconstricted vascular rings from the rat thoracic aorta (EC50=2.4±0.1 μM for temporin-1Gb and 2.3±0.2 μM for temporin-1Gd). The antimicrobial peptides that were isolated in extracts of the skin R. grylio were present in the same molecular forms in electrically-stimulated skin secretions of the animal demonstrating that the peptides are stored in the granular glands of the skin in their fully processed forms.  相似文献   

13.
Spatial and temporal partitioning of resources underlies the coexistence of species with similar niches. In communities of frogs and toads, the phenology of advertisement calling provides insights into temporal partitioning of reproductive effort and its implications for community dynamics. This study assessed the phenology of advertisement calling in an anuran community from Melbourne, in southern Australia. We collated data from 1432 surveys of 253 sites and used logistic regression to quantify seasonality in the nightly probability of calling and the influence of meteorological variables on this probability for six species of frogs. We found limited overlap in the predicted seasonal peaks of calling among these species. Those shown to have overlapping calling peaks are unlikely to be in direct competition, due to differences in larval ecology (Crinia signifera and Litoria ewingii) or differences in calling behavior and acoustics (Limnodynastes dumerilii and Litoria raniformis). In contrast, closely related and ecologically similar species (Crinia signfera and Crinia parinsignifera; Litoria ewingii and Litoria verreauxii) appear to have staggered seasonal peaks of calling. In combination with interspecific variation in the meteorological correlates of calling, these results may be indicative of temporal partitioning of reproductive activity to facilitate coexistence, as has been reported for tropical and temperate anurans from other parts of the globe.  相似文献   

14.
Environmental factors play an integral role, either directly or indirectly, in structuring faunal assemblages. Water chemistry, predation, hydroperiod and competition influence tadpole assemblages within waterbodies. We surveyed aquatic predators, habitat refugia, water height and water chemistry variables (pH, salinity and turbidity) at 37 waterbodies over an intensive 22‐day field survey to determine which environmental factors influence the relative abundance and occupancy of two habitat specialist anuran tadpole species in naturally acidic, oligotrophic waterbodies within eastern Australian wallum communities. The majority of tadpoles found were of Litoria olongburensis (wallum sedge frog) and Crinia tinnula (wallum froglet) species, both habitat specialists that are associated with wallum waterbodies and listed as Vulnerable under the IUCN Red List. Tadpoles of two other species (Litoria fallax (eastern sedge frog), and Litoria cooloolensis (cooloola sedge frog)) were recorded from two waterbodies. Tadpoles of Litoria gracilenta (graceful treefrog) were recorded from one waterbody. Relative abundance and occupancy of L. olongburensis tadpoles were associated with pH and water depth. Additionally, L. olongburensis tadpole relative abundance was negatively associated with turbidity. Waterbody occupancy by C. tinnula tadpoles was negatively associated with predatory fish and water depth and positively associated with turbidity. Variables associated with relative abundance of C. tinnula tadpoles were inconclusive and further survey work is required to identify these environmental factors. Our results show that the ecology of specialist and non‐specialist tadpole species associated with ‘unique’ (e.g. wallum) waterbodies is complex and species specific, with specialist species likely dominating unique habitats.  相似文献   

15.
The dynamics of ammonium (NH4 +) and nitrate (NO3 -) concentrations in the soil solution is an important determinant of the species composition of natural vegetation. A mathematical model of uptake, assimilation and translocation of NH4 + and NO3 - is presented to assess the performance of species with respect to NO3 -/NH4 + feeding characterised by physiologically defined parameters. Nitrate efflux is explicitly considered. The capacities of NO3 -, [U NM], and NH4 + influx, [U AM], and NO3 - reduction, [A NM], appear sufficient to characterise whole-plant N metabolism including NO3 - translocation. The parameter space made up by these parameters is represented by 276 parameter combinations (`species'). Simulated total net N uptake rate and C costs for uptake and assimilation per mole total net N taken up are used to decide on how a species profits or suffers from NO3 -+NH4 + mixtures relative to pure N forms with similar total N concentration for external concentrations up to 1.6 mM. Five response categories were identified and contrasted with categories defined by Bogner (1968) on the basis of experimental results on forest plants. The largest category comprises species that respond positively to NO3 - and positively or indifferently to NH4 +. These species have intermediate to high [U NM] and [A NM] and variable [U AM] and correspond to woodland edge species and forest plants on rich soil including typical `nitrophilic' species. This category fades into a group of species that respond positively to NO3 - and negatively to NH4 +. These species have high [U NM] and low [U AM] and [A NM]; several species from oak-hornbeam woodland (Carpinion) belong to this group. Many parameter combinations were found that responded positively to NH4 + and indifferently to NO3 -: low to medium [U NM], medium to high [U AM] and variable [A NM]. This category includes all heathland species. No species were found which responded negatively to NO3 -. The physiological background of differences between the categories is explained with respect to the equilibrium NO3 - concentration in roots, influx, efflux, translocation and assimilation of NO3 - and uptake and assimilation of NH4 +. The relationship between NO3 - accumulation capacity and morphology is discussed. Some slow-growing species with high [U NM] and low [A NM] use NO3 - mainly as an osmotic solute. Respiratory costs in roots of inherently slow-growing species are discussed with respect to patterns in NH4 + and NO3 - availabilities of their habitat. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

16.
Host defense peptides of 35 species of Australian frogs from the hylids Cyclorana and Litoria, and the myobatrachids Crinia, Limnodynastes and Uperoleia have been identified. The biological activities of the majority of these peptides have been determined and include hormones, neuropeptides, opioids, immunomodulators, membrane active peptides [including antimicrobial, anticancer, antiviral (enveloped viruses like HIV and Herpes) and antifungal peptides], neuronal nitric oxide synthase inhibitors, pheromones and individual peptides with other specific activities. The host defense peptide skin profile can be diagnostic at both the species and higher taxonomic levels; for example, species of Crinia, Litoria and Uperoleia each produce quite different types of peptides. Species of Cyclorana and Limnodynastes are more difficult to characterize by skin peptides alone: species of both genera produce similar peptides with no apparent activity. The skin peptide profiles of frogs from the genera Crinia, Litoria and Uperoleia may be used together with morphological and cognate methods, to differentiate between sub-species and even different population clusters of the same species. Nucleotide sequencing of cDNAs of precursors (pre-pro peptides) of bioactive peptides from the skin glands of various species of the genus Litoria show that the majority of these peptides originated from a single ancestor gene before the break away of Australia from Gondwana. The exceptions are the caerulein neuropeptides {e.g. caerulein [pEQDY(SO3H)TGWMDF(NH2)]} which have a different origin to that of other Litoria peptides. Disulfide containing peptides from skin glands of species of Crinia show a different evolutionary route to peptides from species of Litoria.  相似文献   

17.
The caerin 1 peptides are among the most powerful of the broad-spectrum antibiotic amphibian peptides. Caerin 1.1 has previously been shown to form an amphipathic helix-bend-helix structure in aqueous trifluoroethanol (H. Wong, J. H. Bowie, and J. A. Carver European Journal Biochemistry, 1997, Vol. 247, pp. 545-557) and structure-activity relationship studies indicate that both helices are required for activity, as well as flexibility in the bend region connecting the two. The structure of caerin 1.1 in dodecylphosphocholine micelles was investigated and shown to be very similar to that determined in aqueous trifluoroethanol. Caerin 1.4, which is identical to caerin 1.1, but with serine residues replacing Val5 and Gly7, is less active than caerin 1.1 against most bacterial species but has improved activity against Escherichia coli and Micrococcus luteus. The solution NMR structure of caerin 1.4 was determined in both aqueous trifluoroethanol and dodecylphosphocholine micelles, and was shown to be similar to caerin 1.1. It was concluded that differences in the hydrophobicity and hydrophilic angle of the first helix are probably responsible for the different spectra of antibacterial activity. The similarity of the structures calculated in aqueous trifluoroethanol and dodecylphosphocholine micelles suggests that, for caerin 1.1 and 1.4, these solvent systems are equally as good at representing a membrane environment.  相似文献   

18.
The granular dorsal glands of the giant tree frog Litoria infrafrenata contain five peptides including caerulein (a known neuropeptide), and four new peptides named frenatins 1 (MH+ = 1140 Da), 2 (1423), 3 (2180) and 4 (2493). The amino acid sequences of the frenatins are detailed: their structures do not correspond to those of peptides isolated from other amphibians or animals. Frenatin 3, Gly-Leu-Met-Ser-Val-Leu-Gly-His-Ala-Val-Gly-Asn-Val-Leu-Gly- Gly-Leu-Phe-Lys-Pro-Lys-Ser-(OH), has wide spectrum antimicrobial properties.  相似文献   

19.
Tachykinin perform multiple physiological functions such as smoothing muscle contraction, vasodilation, inflammation, the processing of nerve signal, neuroprotection and neurodegeneration. Two novel tachykinin‐like peptides named tachykinin‐DR1 and ‐DR2 were identified from skin secretions of Danio rerio in current work. Their amino acid sequences were determined as SKSQHFHGLM‐NH2 and NKGEIFVGLM‐NH2, respectively. They share a conserved FXGLM‐NH2C‐terminal consensus motif. By cDNA cloning, the precursor encoding both tachykinin‐DR1 and ‐DR2 was screened from the skin cDNA library of D. rerio. Tachykinin‐DR1 and ‐DR2 share the same precursor, which is composed of 108 amino acid (aa) residues. Regarding the biological activity, tachykinin‐DRs could induce the contraction of isolated strips of guinea pig ileum just like other tackykinins. To our best knowledge, this is the first report of tachykinin from fish skin. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
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