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 46 毫秒
1.
This brief review, including new experimental results, is the summary of a talk at the RSC/SCI conference flavours & fragrances 2004 in Manchester, United Kingdom, 12-14 May, 2004. Musk odorants have been a classical domain for computer aided structure-odor relationship (SOR) studies, but, contrary to sandalwood or amber odorants, they belong to structurally very different substance classes, e.g., macrocycles, aromatic polycycles, and nitro arenes. Most SOR computer models are restricted to one class, excluding structural diversity to increase predictability. But even within a musk family, structural similarities are often due to a common synthetic access, and do not reflect binding requirements for the musk receptor. Beyond that, the importance of structural key features can be missed, which is discussed on the example of the (4S)-Me group of Galaxolide. By synthesis and olfactory evaluation of Galaxolide-like shaped macrobicycles as model compounds for conformationally constrained (12R)-12-methyltridecano-13-lactone, it was investigated how likely there is more than one musk receptor. Finally, the new family of so-called linear musks is discussed, especially with respect to the conformational importance of the gem-2',2'-dimethyl moiety in Helvetolide and the additional 2'-carbonyl group of Romandolide--structural features that strongly diminish the musk odor of macrocycles. On the example of 2-methyl-2-[(E)-1,2,4-trimethylpent-2-enyloxy]propyl esters, the 'brain-aided' design and conformational analysis of musk odorants is illustrated. The overview concludes with the synthesis, odor evaluation, and conformational discussion of the new musk odorant 2-(3,3-dimethylcyclohexyl)propanoic acid ethoxycarbonylmethyl ester.  相似文献   

2.
This review, including some new experimental results, is the summary of a talk at the ‘flavors & fragrances 2013’ conference in Leipzig, organized jointly by the GDCh, the Liebig‐Vereinigung, and the EuCheMS. After times of searching for natural odor principles and serendipitous discoveries by chemical inspiration, directed odorant design today offers the highest hit rates for the discovery of new odorants, although serendipity still plays a role. Keeping intact the electronic shape required for a certain olfactophore‐binding geometry, one can add or subtract structural elements, rigidify molecular structures, or introduce more structural flexibility. To find out which structural features are critical, the ‘seco‐approach’, in which different fragments are removed by cutting strategic bonds, is the most analytical. Following this approach, such ingredients as Serenolide, Sylkolide, and Pomarose were designed. Transferring this design principle from the family of damascones to that of the theaspiranes led to the discovery of Cassyrane, though completely different structural features turned out to be relevant. Application of the seco‐concept to a 3,7a‐substituted 2,6,7,7a‐tetrahydro‐1H‐inden‐5‐yl musk lead structure derived from carotol resulted in the discovery of a new family of dienone musks with novel structure? odor correlations. However, cutting the C(2)? O bond of Cassyrane and oxidizing the resulting secostructure to the 1,2,5,1″‐tetradehydro derivative links the family of dienone musks with that of blackcurrant odorants, but the resulting target structures turned out to be potent orris odorants. (3E,5E)‐5‐(tert‐Butyl)octadeca‐3,5‐dien‐2‐one even possesses the lowest odor threshold in the whole ionone family (0.036 ng/l air), which could be rationalized by a superposition analysis on (?)‐cisγ‐irone. In the course of the synthesis of these high‐impact orris odorants, we discovered that, depending on the reaction conditions, the dehydration step of the intermediate 5‐hydroxyalk‐3‐yn‐2‐ones was accompanied by a carbenium‐ion rearrangement. Depending on the substitution pattern, these rearrangement products and their derivatives possessed interesting musky‐woody olfactory properties reminiscent of Cashmeran, demonstrating that the same structural elements can code for completely different odors, i.e., cassis, musk, orris, violet, and Cashmeran‐type, depending only on their spatial arrangement.  相似文献   

3.
Odor Perception Phenotypes: Multiple, Specific Hyperosmias to Musks   总被引:1,自引:0,他引:1  
Olfactory detection thresholds for 11 structurally diverse muskodorants and one non-musk odorant were obtained from 32 subjects.Hierarchical cluster analysis produced four groups of subjects.One group (n = 12) was uniformly sensitive to all musks; another(n = 16) was uniformly insensitive. Two groups of subjects containedotherwise insensitive individuals who were exceptionally sensitiveto cyclopentadecanone and musk xylol (n = 2) and to delta9-hexadecenolactoneand tonalid (n = 2) respectively. We propose that the lattertwo groups are odor perception phenotypes (MSHM1 and MSHM2)that consist of multiple, specific hyperosmias to musk odorants.Chem. Senses 21: 411– 416, 1996. 1Present address: Synesthetics, Inc., Montclair, NJ 07043, USA  相似文献   

4.
Kraft P  Fráter G 《Chirality》2001,13(8):388-394
This brief review, the summary of a talk at the Symposium on Biological Chirality 2000 in Szeged, Hungary, illustrates what chiral recognition tells us about the molecular parameters of the musk odor sensation. While the enantioselectivity of odor perception is strong evidence for the key role of proteinogenic receptors in the molecular mechanism of olfaction, the quantitative and qualitative odor differences of enantiomers are often not very pronounced, as in the case of muscone (17/26). In those cases, however, where there is strong enantiodiscrimination, we find most intense musk odorants with very low odor thresholds, such as (-)-(12R)-12-methyl-9-oxa-14-tetradecanolide (35), (12R;9Z)-12-methyl-14-tetradec-9-enolide [(R)-Nirvanolide, 38], and (-)-(4S;7R)-1,3,4,6,7,8-hexahydro-4,6,6,7,8,8-hexamethylcyclopenta[g]-2-benzopyran [(-)-(4S;7R)-Galaxolide, 57], the latter being rather rigid. We thus can assume the geometry of the musk receptor to be fairly complementary to these compounds, which therefore can serve as templates for the design of new musk odorants.  相似文献   

5.
A method for the calculation of odor character from molecular structure   总被引:1,自引:0,他引:1  
The relationship between molecular structure and odor character is one of the most complex structure-activity problems in biology. Despite over a century of effort, it remains unsolved, and synthesis of new odorants still proceeds largely by trial and error. In previous work, I have argued that the reason for this failure lies in a mistaken assumption, namely that molecular shape determines odor character. Instead, I have taken up and extended an old idea (Dyson, 1938) according to which vertebrate olfactory receptors detect odorants by their molecular vibrations. I propose that the detection mechanism is inelastic electron tunnelling. If this is correct, there should be a correlation between the tunnelling vibrational spectra of odorants and their odor character. Here, using semi-empirical quantum chemistry methods and a simple calculation method for tunnelling mode intensities, I calculate the spectra of structurally diverse odorants belonging to various odor categories. With few exceptions, the calculated spectra of bitter almonds, musks, ambers, woods, sandalwoods and violets strongly correlate with odor character.  相似文献   

6.
A comprehensive method was developed for the simultaneous analysis in human breast milk of 12 synthetic musks, five nitro musks, six polycyclic muks and one macrocyclic musk; as well as one musk metabolite and triclosan. The target analytes were freeze dried and extracted using the accelerated solvent extraction (ASE) procedure. The extracts were further purified by gel permeation chromatography (GPC) and florisil solid-phase extraction (SPE) and then analyzed by gas chromatography tandem mass spectrometry (GC-MS/MS). Recoveries of the analytes based on the isotopic internal standard correction ranged from 82.4% to 112%, with relative standard derivations less than 20%. The method quantification limits (MQLs) were 0.6-5.4 ng/g lipid. The analytes were detected in human breast milk samples and ranged from 11.7 to 308.6 ng/g lipid.  相似文献   

7.
Structures of two strong benzenoid musks, 3,5-di-t-butylacetophenoneand 5-acetyl-1,1,2,3,3,6-hexamethylindan, were optimized byPM3 (MOPAC Ver.5). Based on these structures, a three-dimensionalstructure model for benzenoid musks was constructed. This structuremodel is expressed as a box of about 12 x 11 x 6.5 Å Whichcircumscribes musk molecules. The box contains (i) a smallerbox of about 7.5 x 11 x 6.5 Å Which approximates the hydrophobicbulky moiety of benzenoid musks; (ii) the position of the benzenering located in the smaller box; and (iii) the binding pointon an odor receptor assumed to interact with an oxygen atomof the functional group in the benzenoid musks. The validity of this structure model was examined by applyingit to structures of 40 benzenoids (30 musks and 10 odorless).As a result, 37 out of the 40 benzenoids were correctly discriminatedto be odoriferous or odorless by referring them to this structuremodel.  相似文献   

8.
The syntheses and olfactory evaluations of eight new macrocyclic musks with a 1,6-dioxa structure (1a-d and 1'a-d) as well as of twelve optically active 3-methyl macrolides (5a-c and 5'a-c) are reported. These macrocycles were synthesized via intramolecular metathesis mediated by the Grubbs catalyst. Despite the absence of a C=O function, the 1,6-dioxa compounds, both unsaturated (1) and saturated (1'), possess musky odors similar to those of macrocyclic ketones and lactones. Especially 16-membered rings were found to display an intense and pleasant musk character. However, in the case of optically active 3-methyl macrolides (5, 5'), only the (R)-configured 15- and 16-membered rings had intense and pleasant musk notes.  相似文献   

9.
This brief review, including new experimental results, is the summary of a talk at the GDCh conference ‘flavors & fragrances 2013’ in Leipzig, Germany, 11th–13th September, 2013. Musk odorants are indispensable in perfumery to lend sensuality to fine fragrances, a nourishing effect to cosmetics, and a comforting feeling to laundry. We have recently found serendipitously a new oxy‐oxonia‐Cope rearrangement. In this account, we review the background of oxonia‐sigmatropic rearrangements and the discovery of this novel reaction. Special attention is focused on the versatile lactone and lactam formation reactions via [n+4] ring enlargement and the macrocyclization in the synthesis of new macrocyclic musks. The synthesized structures provide new insights into the structure? odor relationships of musks.  相似文献   

10.
11.
12.
D'ANDREA  CARL L. 《Chemical senses》1975,1(3):359-369
The odor intensities of six commercial musks: celestolide, muskolactone,galaxolide, musk 36A®, tonalid®, and musk ambrette*were matched by the method of equal odor intensity. Galaxolidewas arbitrarily used as the reference odorant in order to generatea scale for the determination of relative odor intensities.The odor intensities are expressed in terms of Stevens' psychophysicalpower function with the exponent value, b, ranging from 0.43to 0.73.  相似文献   

13.
Revial et al. (1982) reported the experimental findings obtainedin cell units stimulated with two new sets of odorants. Thepresent paper deals with mathematical processing of these electrophysiologicaldata. The multidimensionality of the odour space was confirmed.Camphor, isoborneol and cineole appeared to represent a ‘camphorgroup’ markedly discriminated from other odorants. A‘terpenegroup’ including terpinene, dimethylstyrene, limoneneand cymene and related to menthane and menthene was observed.A series of 8 cycloketones was found to have a linear patternin the factorial planes and to be subdivided into three subsets.Cyclooctadecanone and cyclotetradecanone displayed correlationsand spatial relationships with musk ketone while cycloundecanone,cyclodecanone and especially cyclononanone were more relatedto the‘camphor group’. Separate processing of theunitary responses of 12 receptor cells recorded from a singlepreparation led to the building of a factorial space resemblingthat constructed with the full collection of data. The incidenceof odorant concentration on similarity evaluation is discussedon the basis of receptor responses to two different concentrationsof some of the odorants.  相似文献   

14.
王永奇  盛岩  刘文华  李斐然  唐婕  孟秀祥 《生态学报》2015,35(15):4986-4992
林麝是濒危资源动物,林麝驯养是保育濒危林麝资源及可持续利用麝香的有效方式。基于对2001至2012年间的陕西凤县林麝驯养的监测和调查,分析了其种群动态、性比和年龄结构。结果表明,陕西凤县的林麝驯养在近10余年获得了快速发展,全县共有150余个麝场,麝场数呈指数式增长,增长率达27.33%,但其平均驯养规模无明显增长,平均存栏种群仅为16.38头。凤县的驯养麝种群总体增长近似指数式增长,增长率达27.22%,目前存栏种群已达3712头。区分性别和年龄,各亚群的增长均呈指数式增长,幼年麝的增长率(30.30%)高于成年麝(27.16%),雄麝的增长率(28.30%)高于雌麝(27.78%)。在2001至2012年间,幼麝种群的雌雄性比((102.64±3.15)%,n=12)和成年麝种群的雌雄性比((100.85±2.585)%,n=10)均显著偏雌(P0.01),但幼麝、成年麝种群间的性比差异不显著(P0.05)。在2005年及2010—2012年间,幼麝(0.5岁龄)占种群的比例为31.91%,亚成体麝(1.5岁龄)占种群的比例为21.11%,成麝(2.5—12.5岁龄)比例为42.72%,老年麝(13.5岁龄及以上)仅占种群的4.26%。合并年龄分析,育龄前个体(幼麝和亚成体麝)的平均比例为53.02%,表明凤县驯养林麝属快速增长种群,其增长潜力较大。在林麝驯养实践中,管理部门可制定准入制度或适当重组现有麝场,促进较大的驯养种群构建,并建立通畅的麝香交易渠道,控制林麝种源的过热交易,以利于林麝驯养种群的性比平衡及可持续的繁育、增长。  相似文献   

15.
During the past years biocatalytic production of fine chemicals has been expanding rapidly. Flavours and fragrances belong to many different structural classes and therefore represent a challenging target for academic and industrial research. Here, we present a condensed overview of the potential offered by biocatalysis for the synthesis of natural and natural-identical odorants, highlighting relevant biotransformations using microorganisms and isolated enzymes. The industrial processes based on biocatalytic methods are discussed in terms of their advantages over classical chemical synthesis and extraction from natural sources. Recent applications of the biocatalytic approach to the preparation of the most important fine odorants are comprehensively covered.  相似文献   

16.
Humans are able to detect and discriminate myriads of odorants using only several hundred olfactory receptors (ORs) classified in two major phylogenetic classes representing ORs from aquatic (class I) and terrestrial animals (class II). Olfactory perception results in a combinatorial code, in which one OR recognizes multiple odorants and different odorants are recognized by different combinations of ORs. Moreover, recent data suggest that odorants could also behave as antagonists for other ORs, thus making the combinatorial coding more complex. Here we describe the odorant repertoires of two human ORs belonging to class I and class II, respectively. For this purpose, we set up an assay based on calcium imaging in which 100 odorants were screened using air-phase odorant stimulation at physiological doses. We showed that the human class I OR52D1 is functional, exhibiting a narrow repertoire related to that of its orthologous murine OR, demonstrating than this human class I OR is not an evolutionary relic. The class II OR1G1 was revealed to be broadly tuned towards odorants of 9-10 carbon chain length, with diverse functional groups. The existence of antagonist odorants for the class II OR was also demonstrated. They are structurally related to the agonists, with shorter carbon chain length.  相似文献   

17.
Odorant-binding proteins (OBPs) are small abundant soluble proteins belonging to the lipocalin superfamily, which are thought to carry hydrophobic odorants through aqueous mucus towards olfactory receptors. Human variant hOBP-2A has been demonstrated to bind numerous odorants of different chemical classes with a higher affinity for aldehydes and fatty acids. Three lysyl residues of the binding pocket (Lys62, Lys82 and Lys112) have been suggested as candidates for playing such a role. Here, using site-directed mutagenesis and fluorescent probe displacements, we show that Lys112 is the major determinant for governing hOBP-2A specificity towards aldehydes and small carboxylic acids.  相似文献   

18.
Phosphoinositide 3-kinase (PI3K) signaling has been implicated in mediating inhibitory odorant input to mammalian olfactory receptor neurons (ORNs). To better understand the breadth of such inhibition in odor coding, we screened a panel of odorants representing different chemical classes, as well as odorants known to occur in a natural odor object (tomato), for their ability to rapidly activate PI3K-dependent inhibitory signaling. Odorants were screened on dissociated native rat ORNs before and after pre-incubation with the PI3K-isoform specific blockers AS252424 and TGX221. Many different odorants increased their excitatory strength for particular ORNs following PI3K blockade in a manner consistent with activating PI3K-dependent inhibitory signaling in those cells. The PI3K-dependent inhibitory odorants overlapped with conventional excitatory odorants, but did not share the same bias, indicating partial partitioning of the odor space. Finding that PI3K-dependent inhibition can be activated by a wide range of otherwise conventional excitatory odorants strongly implies PI3K-dependent inhibition provides a broad basis for opponent coding in mammalian ORNs.  相似文献   

19.
Receptor-ligand interaction models are generally based on a 'lock and key' concept. How far this holds true for olfactory receptors and odor molecules is currently uncertain. Here, we have investigated the response of a human olfactory receptor, OR1D2, to a broad array of odorants and found that there is no simple, direct correlation between a molecule's ability to activate this receptor and the odor impression elicited in the brain. In a parallel study on specific anosmia, we have found no evidence for odor-specific anosmia to either musk or amber, but rather to specific molecules within these categories. Cluster analysis confirmed that there is no simple correlation between molecular structure and impaired perception in either odor type. There are some differences in patterns of impairment between the two odor types and some evidence to suggest that subjects with specific anosmia to a given substance can identify its presence in a mixture. Taken together, our results show that simplistic 'lock and key' models of olfaction based on a concept of odor-quality-tuned receptors are inadequate, irrespective of the nature of the lock-key interaction. Receptor activation is only one step in a long chain of events leading from inhalation of odorants to perception of odor in the higher brain, and, therefore, although structure-odor correlations are useful tools for the design of novel odorants, caution should be exercised when extrapolating them to models of olfactory perception. Those seeking to understand the odorant-receptor interaction should use receptor activation rather than odor as input data.  相似文献   

20.
Using a conditioning paradigm, we assessed the olfactory sensitivity of six CD-1 mice (Mus musculus) for six sulfur-containing odorants known to be components of the odors of natural predators of the mouse. With all six odorants, the mice discriminated concentrations <0.1 ppm (parts per million) from the solvent, and with five of the six odorants the best-scoring animals were even able to detect concentrations <1 ppt (parts per trillion). Four female spider monkeys (Ateles geoffroyi) and twelve human subjects (Homo sapiens) tested in parallel were found to detect the same six odorants at concentrations <0.01 ppm, and with four of the six odorants the best-scoring animals and subjects even detected concentrations <10 ppt. With all three species, the threshold values obtained here are generally lower than (or in the lower range of) those reported for other chemical classes tested previously, suggesting that sulfur-containing odorants may play a special role in olfaction. Across-species comparisons showed that the mice were significantly more sensitive than the human subjects and the spider monkeys with four of the six predator odorants. However, the human subjects were significantly more sensitive than the mice with the remaining two odorants. Human subjects and spider monkeys significantly differed in their sensitivity with only two of the six odorants. These comparisons lend further support to the notion that the number of functional olfactory receptor genes or the relative or absolute size of the olfactory bulbs are poor predictors of a species’ olfactory sensitivity. Analysis of odor structure–activity relationships showed that in both mice and human subjects the type of alkyl rest attached to a thietane and the type of oxygen moiety attached to a thiol significantly affected olfactory sensitivity.  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