Vibration theory of olfaction

Last updated

The vibration theory of smell proposes that a molecule's smell character is due to its vibrational frequency in the infrared range. This controversial theory is an alternative to the more widely accepted docking theory of olfaction (formerly termed the shape theory of olfaction), which proposes that a molecule's smell character is due to a range of weak non-covalent interactions between its protein odorant receptor (found in the nasal epithelium), such as electrostatic and Van der Waals interactions as well as H-bonding, dipole attraction, pi-stacking, metal ion, Cation–pi interaction, and hydrophobic effects, in addition to the molecule's conformation. [1] [2] [3]

Contents

Introduction

The current vibration theory has recently been called the "swipe card" model, in contrast with "lock and key" models based on shape theory. [4] As proposed by Luca Turin, the odorant molecule must first fit in the receptor's binding site.[ citation needed ] Then it must have a vibrational energy mode compatible with the difference in energies between two energy levels on the receptor, so electrons can travel through the molecule via inelastic electron tunneling, triggering the signal transduction pathway. [5] The vibration theory is discussed in a popular but controversial book by Chandler Burr. [6] [7]

The odor character is encoded in the ratio of activities of receptors tuned to different vibration frequencies, in the same way that color is encoded in the ratio of activities of cone cell receptors tuned to different frequencies of light. An important difference, though, is that the odorant has to be able to become resident in the receptor for a response to be generated. The time an odorant resides in a receptor depends on how strongly it binds, which in turn determines the strength of the response; the odor intensity is thus governed by a similar mechanism to the "lock and key" model. [5] For a pure vibrational theory, the differing odors of enantiomers, which possess identical vibrations, cannot be explained. However, once the link between receptor response and duration of the residence of the odorant in the receptor is recognised, differences in odor between enantiomers can be understood: molecules with different handedness may spend different amounts of time in a given receptor, and so initiate responses of different intensities.

Seeing as there are some aroma molecules of different shapes that smell the same (eg. benzaldehyde, that gives the same scent to both almonds and/or cyanide), the shape "lock and key" model is not quite sufficient to explain what is going on. Experiments with olfaction, taking quantum mechanics into consideration, suggest that ultimately both theories might work in harmony - first the scent molecules need to fit, as in the docking theory of olfaction model, but then the molecular vibrations of the chemical/atom bonds take over. So in essence your sense of smell could be much more like your sense of hearing, where your nose could be 'listening' to the acoustic/vibrational bonds of aroma molecules.

Some studies support vibration theory while others challenge its findings.

Major proponents and history

The theory was first proposed by Malcolm Dyson in 1928 [8] and expanded by Robert H. Wright in 1954, after which it was largely abandoned in favor of the competing shape theory. A 1996 paper by Luca Turin revived the theory by proposing a mechanism, speculating that the G-protein-coupled receptors discovered by Linda Buck and Richard Axel were actually measuring molecular vibrations using inelastic electron tunneling as Turin claimed, rather than responding to molecular keys fitting molecular locks, working by shape alone. [5] [9] In 2007 a Physical Review Letters paper by Marshall Stoneham and colleagues at University College London and Imperial College London showed that Turin's proposed mechanism was consistent with known physics and coined the expression "swipe card model" to describe it. [10] A PNAS paper in 2011 by Turin, Efthimios Skoulakis, and colleagues at MIT and the Alexander Fleming Biomedical Sciences Research Center reported fly behavioral experiments consistent with a vibrational theory of smell. [11] The theory remains controversial. [3] [12] [13] [14] [15] [16] [17] [18]

Support

Isotope effects

A major prediction of Turin's theory is the isotope effect: that the normal and deuterated versions of a compound should smell different, although they have the same shape. A 2001 study by Haffenden et al. showed humans able to distinguish benzaldehyde from its deuterated version. [19] [20] However, this study has been criticized for lacking double-blind controls to eliminate bias and because it used an anomalous version of the duo-trio test. [21] In another study, tests with animals have shown fish and insects able to distinguish isotopes by smell. [22] [23] [24] [25]

Deuteration changes the heats of adsorption and the boiling and freezing points of molecules (boiling points: 100.0 °C for H2O vs. 101.42 °C for D2O; melting points: 0.0 °C for H2O, 3.82 °C for D2O), pKa (i.e., dissociation constant: 9.71×10−15 for H2O vs. 1.95×10−15 for D2O, cf. Heavy water) and the strength of hydrogen bonding. Such isotope effects are exceedingly common, and so it is well known that deuterium substitution will indeed change the binding constants of molecules to protein receptors. [26] Any binding interaction of an odorant molecule with an olfactory receptor will therefore be likely to show some isotope effect upon deuteration, and the observation of an isotope effect in no way argues exclusively for a vibrational theory of olfaction.

A study published in 2011 by Franco, Turin, Mershin and Skoulakis shows both that flies can smell deuterium, and that to flies, a carbon-deuterium bond smells like a nitrile, which has a similar vibration. The study reports that drosophila melanogaster (fruit fly), which is ordinarily attracted to acetophenone, spontaneously dislikes deuterated acetophenone. This dislike increases with the number of deuteriums. (Flies genetically altered to lack smell receptors could not tell the difference.) Flies could also be trained by electric shocks either to avoid the deuterated molecule or to prefer it to the normal one. When these trained flies were then presented with a completely new and unrelated choice of normal vs. deuterated odorants, they avoided or preferred deuterium as with the previous pair. This suggested that flies were able to smell deuterium regardless of the rest of the molecule. To determine whether this deuterium smell was actually due to vibrations of the carbon-deuterium (C-D) bond or to some unforeseen effect of isotopes, the researchers looked to nitriles, which have a similar vibration to the C-D bond. Flies trained to avoid deuterium and asked to choose between a nitrile and its non-nitrile counterpart did avoid the nitrile, lending support to the idea that the flies are smelling vibrations. [25] Further isotope smell studies are under way in fruit flies and dogs. [27]

Explaining differences in stereoisomer scents

Carvone presented a perplexing situation to vibration theory. Carvone has two isomers, which have identical vibrations, yet one smells like mint and the other like caraway (for which the compound is named).

An experiment by Turin filmed by the 1995 BBC Horizon documentary "A Code in the Nose" consisted of mixing the mint isomer with butanone, on the theory that the shape of the G-protein-coupled receptor prevented the carbonyl group in the mint isomer from being detected by the "biological spectroscope". The experiment succeeded with the trained perfumers used as subjects, who perceived that a mixture of 60% butanone and 40% mint carvone smelled like caraway.

The sulfurous smell of boranes

According to Turin's original paper in the journal Chemical Senses, the well documented smell of borane compounds is sulfurous, though these molecules contain no sulfur. He proposes to explain this by the similarity in frequency between the vibration of the B-H bond and the S-H bond. [5] However, it has been pointed out that for o-carborane, which has a very strong B−H stretch at 2575 cm−1, the "onion-like odor of crude commercial o-carborane is replaced by a pleasant camphoraceous odor on careful purification, reflecting the method for commercial preparation of o-carborane from reactions promoted by onion-smelling diethyl sulfide, which is removed on purification." [3]

Consistency with physics

Biophysical simulations published in Physical Review Letters in 2006 suggest that Turin's proposal is viable from a physics standpoint. [10] [28] However, Block et al. in their 2015 paper in Proceedings of the National Academy of Sciences indicate that their theoretical analysis shows that "the proposed electron transfer mechanism of the vibrational frequencies of odorants [10] could be easily suppressed by quantum effects of nonodorant molecular vibrational modes". [17]

Correlating odor to vibration

A 2004 paper published in the journal Organic Biomolecular Chemistry by Takane and Mitchell shows that odor descriptions in the olfaction literature correlate with EVA descriptors, which loosely correspond to the vibrational spectrum, better than with descriptors based on the two dimensional connectivity of the molecule. The study did not consider molecular shape. [29]

Lack of antagonists

Turin points out that traditional lock-and-key receptor interactions deal with agonists, which increase the receptor's time spent in the active state, and antagonists, which increase the time spent in the inactive state. In other words, some ligands tend to turn the receptor on and some tend to turn it off. As an argument against the traditional lock-and-key theory of smell, very few olfactory antagonists have been found.

In 2004, a Japanese research group published that an oxidation product of isoeugenol is able to antagonize, or prevent, mice olfactory receptor response to isoeugenol. [30]

Additional challenges to the docking theory of olfaction

Challenges

Three predictions by Luca Turin on the nature of smell, using concepts of vibration theory, were addressed by experimental tests published in Nature Neuroscience in 2004 by Vosshall and Keller. [21] The study failed to support the prediction that isotopes should smell different, with untrained human subjects unable to distinguish acetophenone from its deuterated counterpart. [10] [28] [34] This study also pointed to experimental design flaws in the earlier study by Haffenden. [19] In addition, Turin's description of the odor of long-chain aldehydes as alternately (1) dominantly waxy and faintly citrus and (2) dominantly citrus and faintly waxy was not supported by tests on untrained subjects, despite anecdotal support from fragrance industry professionals who work regularly with these materials. Vosshall and Keller also presented a mixture of guaiacol and benzaldehyde to subjects, to test Turin's theory that the mixture should smell of vanillin. Vosshall and Keller's data did not support Turin's prediction. However, Vosshall says these tests do not disprove the vibration theory. [35]

In response to the 2011 PNAS study on flies, Vosshall acknowledged that flies could smell isotopes but called the conclusion that smell was based on vibrations an "overinterpretation" and expressed skepticism about using flies to test a mechanism originally ascribed to human receptors. [27] For the theory to be confirmed, Vosshall stated there must be further studies on mammalian receptors. [36] Bill Hansson, an insect olfaction specialist, raised the question of whether deuterium could affect hydrogen bonds between the odorant and receptor. [37]

In 2013, Turin and coworkers confirmed Vosshall and Keller's experiments showing that even trained human subjects were unable to distinguish acetophenone from its deuterated counterpart. [38] At the same time Turin and coworkers reported that human volunteers were able to distinguish cyclopentadecanone from its fully deuterated analog. To account for the different results seen with acetophenone and cyclopentadecanone, Turin and coworkers assert that "there must be many C-H bonds before they are detectable by smell. In contrast to acetophenone which contains only 8 hydrogens, cyclopentadecanone has 28. This results in more than 3 times the number of vibrational modes involving hydrogens than in acetophenone, and this is likely essential for detecting the difference between isotopomers." [38] [39] Turin and coworkers provide no quantum mechanical justification for this latter assertion. Note that the correct term for compounds differing in the number of isotopic substitutions is isotopologue; isotopomers differ only in the position of the substitutions.

Vosshall, in commenting on Turin's work, notes that "the olfactory membranes are loaded with enzymes that can metabolise odorants, changing their chemical identity and perceived odour. Deuterated molecules would be poor substrates for such enzymes, leading to a chemical difference in what the subjects are testing. Ultimately, any attempt to prove the vibrational theory of olfaction should concentrate on actual mechanisms at the level of the receptor, not on indirect psychophysical testing." [15] Richard Axel co-recipient of the 2004 Nobel prize for physiology for his work on olfaction, expresses a similar sentiment, indicating that Turin's work "would not resolve the debate – only a microscopic look at the receptors in the nose would finally show what is at work. Until somebody really sits down and seriously addresses the mechanism and not inferences from the mechanism... it doesn't seem a useful endeavour to use behavioural responses as an argument". [13]

In response to the 2013 paper on cyclopentadecanone, [38] Block et al. [17] report that the human musk-recognizing receptor, OR5AN1, identified using a heterologous olfactory receptor expression system and robustly responding to cyclopentadecanone and muscone (which has 30 hydrogens), fails to distinguish isotopologues of these compounds in vitro. Furthermore, the mouse (methylthio)methanethiol-recognizing receptor, MOR244-3, as well as other selected human and mouse olfactory receptors, responded similarly to normal, deuterated, and carbon-13 isotopologues of their respective ligands, paralleling results found with the musk receptor OR5AN1. Based on these findings, the authors conclude that the proposed vibration theory does not apply to the human musk receptor OR5AN1, mouse thiol receptor MOR244-3, or other olfactory receptors examined. Additionally, theoretical analysis by the authors shows that the proposed electron transfer mechanism of the vibrational frequencies of odorants could be easily suppressed by quantum effects of nonodorant molecular vibrational modes. The authors conclude: "These and other concerns about electron transfer at olfactory receptors, together with our extensive experimental data, argue against the plausibility of the vibration theory."

In commenting on this work, Vosshall writes "In PNAS, Block et al.... shift the "shape vs. vibration" debate from olfactory psychophysics to the biophysics of the ORs themselves. The authors mount a sophisticated multidisciplinary attack on the central tenets of the vibration theory using synthetic organic chemistry, heterologous expression of olfactory receptors, and theoretical considerations to find no evidence to support the vibration theory of smell." [1] While Turin comments that Block used "cells in a dish rather than within whole organisms" and that "expressing an olfactory receptor in human embryonic kidney cells doesn't adequately reconstitute the complex nature of olfaction...", Vosshall responds "Embryonic kidney cells are not identical to the cells in the nose ... but if you are looking at receptors, it's the best system in the world." [40] In a Letter to the Editor of PNAS, Turin et al. [41] raise concerns about Block et al. [17] and Block et al. respond. [42]

Recently, Saberi and Allaei have suggested that a functional relationship exists between molecular volume and the olfactory neural response. The molecular volume is an important factor, but it is not the only factor that determines the response of ONRs. The binding affinity of an odorant-receptor pair is affected by their relative sizes. The maximum affinity can be attained when the molecular volume of an odorant matches the volume of the binding pocket. [43] A recent study [44] describes the responses of primary olfactory neurons in tissue culture to isotopes and finds that a small fraction of the population (<1%) clearly discriminates between isotopes, some even giving an all-or-or -none response to H or D isotopologues of octanal. The authors attribute this to differences in hydrophobicity between normal and deuterated odorants.

See also

Related Research Articles

<span class="mw-page-title-main">Olfactory nerve</span> Cranial nerve I, for smelling

The olfactory nerve, also known as the first cranial nerve, cranial nerve I, or simply CN I, is a cranial nerve that contains sensory nerve fibers relating to the sense of smell.

<span class="mw-page-title-main">Olfactory receptor neuron</span> Transduction nerve cell within the olfactory system

An olfactory receptor neuron (ORN), also called an olfactory sensory neuron (OSN), is a sensory neuron within the olfactory system.

<span class="mw-page-title-main">Olfactory receptor</span> Chemoreceptors expressed in cell membranes of olfactory receptor neurons

Olfactory receptors (ORs), also known as odorant receptors, are chemoreceptors expressed in the cell membranes of olfactory receptor neurons and are responsible for the detection of odorants which give rise to the sense of smell. Activated olfactory receptors trigger nerve impulses which transmit information about odor to the brain. In vertebrates, these receptors are members of the class A rhodopsin-like family of G protein-coupled receptors (GPCRs). The olfactory receptors form a multigene family consisting of around 400 genes in humans and 1400 genes in mice. In insects, olfactory receptors are members of an unrelated group of ligand-gated ion channels.

<span class="mw-page-title-main">Aroma compound</span> Chemical compound that has a smell or odor

An aroma compound, also known as an odorant, aroma, fragrance or flavoring, is a chemical compound that has a smell or odor. For an individual chemical or class of chemical compounds to impart a smell or fragrance, it must be sufficiently volatile for transmission via the air to the olfactory system in the upper part of the nose. As examples, various fragrant fruits have diverse aroma compounds, particularly strawberries which are commercially cultivated to have appealing aromas, and contain several hundred aroma compounds.

<span class="mw-page-title-main">Androstenone</span> Chemical compound

Androstenone (5α-androst-16-en-3-one) is a 16-androstene class steroidal pheromone. It is found in boar's saliva, celery cytoplasm, and truffle fungus. Androstenone was the first mammalian pheromone to be identified. It is found in high concentrations in the saliva of male pigs, and, when inhaled by a female pig that is in heat, results in the female assuming the mating stance. Androstenone is the active ingredient in 'Boarmate', a commercial product made by DuPont sold to pig farmers to test sows for timing of artificial insemination.

<span class="mw-page-title-main">Docking theory of olfaction</span>

The docking theory of olfaction proposes that the smell of an odorant molecule is due to a range of weak non-covalent interactions between the odorant [a ligand] and one or more G protein-coupled odorant receptors. These include intermolecular forces, such as dipole-dipole and Van der Waals interactions, as well as hydrogen bonding. More specific proposed interactions include metal-ion, ion-ion, cation-pi and pi-stacking. Interactions can be influenced by the hydrophobic effect. Conformational changes can also have a significant impact on interactions with receptors, as ligands have been shown to interact with ligands without being in their conformation of lowest energy.

<span class="mw-page-title-main">Muscone</span> Chemical compound

Muscone is a macrocyclic ketone, an organic compound that is the primary contributor to the odor of musk.

The antennal lobe is the primary olfactory brain area in insects. The antennal lobe is a sphere-shaped deutocerebral neuropil in the brain that receives input from the olfactory sensory neurons in the antennae and mouthparts. Functionally, it shares some similarities with the olfactory bulb in vertebrates. The anatomy and physiology function of the insect brain can be studied by dissecting open the insect brain and imaging or carrying out in vivo electrophysiological recordings from it.

Luca Turin is a biophysicist and writer with a long-standing interest in bioelectronics, the sense of smell, perfumery, and the fragrance industry.

Olfactory fatigue, also known as odor fatigue, olfactory adaptation, and noseblindness, is the temporary, normal inability to distinguish a particular odor after a prolonged exposure to that airborne compound. For example, when entering a restaurant initially the odor of food is often perceived as being very strong, but after time the awareness of the odor normally fades to the point where the smell is not perceptible or is much weaker. After leaving the area of high odor, the sensitivity is restored with time. Anosmia is the permanent loss of the sense of smell, and is different from olfactory fatigue.

Hyperosmia is an increased olfactory acuity, usually caused by a lower threshold for odor. This perceptual disorder arises when there is an abnormally increased signal at any point between the olfactory receptors and the olfactory cortex. The causes of hyperosmia may be genetic, hormonal, environmental or the result of benzodiazepine withdrawal syndrome.

Quantum biology is the study of applications of quantum mechanics and theoretical chemistry to aspects of biology that cannot be accurately described by the classical laws of physics. An understanding of fundamental quantum interactions is important because they determine the properties of the next level of organization in biological systems.

Dysosmia is a disorder described as any qualitative alteration or distortion of the perception of smell. Qualitative alterations differ from quantitative alterations, which include anosmia and hyposmia. Dysosmia can be classified as either parosmia or phantosmia. Parosmia is a distortion in the perception of an odorant. Odorants smell different from what one remembers. Phantosmia is the perception of an odor when no odorant is present. The cause of dysosmia still remains a theory. It is typically considered a neurological disorder and clinical associations with the disorder have been made. Most cases are described as idiopathic and the main antecedents related to parosmia are URTIs, head trauma, and nasal and paranasal sinus disease. Dysosmia tends to go away on its own but there are options for treatment for patients that want immediate relief.

<span class="mw-page-title-main">OR1A1</span> Protein-coding gene in the species Homo sapiens

Olfactory receptor 1A1 is a protein that in humans is encoded by the OR1A1 gene.

<span class="mw-page-title-main">OR13H1</span> Protein-coding gene in the species Homo sapiens

Olfactory receptor 13H1 is a protein that in humans is encoded by the OR13H1 gene.

<span class="mw-page-title-main">Sense of smell</span> Sense that detects smells

The sense of smell, or olfaction, is the special sense through which smells are perceived. The sense of smell has many functions, including detecting desirable foods, hazards, and pheromones, and plays a role in taste.

Olfactory memory refers to the recollection of odors. Studies have found various characteristics of common memories of odor memory including persistence and high resistance to interference. Explicit memory is typically the form focused on in the studies of olfactory memory, though implicit forms of memory certainly supply distinct contributions to the understanding of odors and memories of them. Research has demonstrated that the changes to the olfactory bulb and main olfactory system following birth are extremely important and influential for maternal behavior. Mammalian olfactory cues play an important role in the coordination of the mother infant bond, and the following normal development of the offspring. Maternal breast odors are individually distinctive, and provide a basis for recognition of the mother by her offspring.

<span class="mw-page-title-main">Leslie B. Vosshall</span> American neurobiologist

Leslie Birgit Vosshall is an American neurobiologist and currently a Howard Hughes Medical Institute (HHMI) investigator and the Robin Chemers Neustein Professor of Neurogenetics and Behavior at The Rockefeller University. In 2022 she was appointed Chief Scientific Officer and vice president of HHMI. She is also the director of the Kavli Neural Systems Institute at The Rockefeller University. Vosshall, a member of the National Academy of Sciences, is known for her contributions to the field of olfaction, particularly for the discovery and subsequent characterization of the insect olfactory receptor family, and the genetic basis of chemosensory behavior in mosquitoes. She has also extended her research into the study of human olfaction, revealing parts of human genetic olfactory architecture, and finding variations in odorant receptors that determine individuals’ abilities to detect odors.

Or83b, also known as Orco, is an odorant receptor and the corresponding gene that encodes it. The odorant receptor Or83b is not exclusively expressed in insects. Though its actual function is still a mystery, the broadly expressed Or83b has been conserved across highly divergent insect populations across 250 million years of evolution.

<span class="mw-page-title-main">Insect olfaction</span> Function of chemical receptors

Insect olfaction refers to the function of chemical receptors that enable insects to detect and identify volatile compounds for foraging, predator avoidance, finding mating partners and locating oviposition habitats. Thus, it is the most important sensation for insects. Most important insect behaviors must be timed perfectly which is dependent on what they smell and when they smell it. For example, olfaction is essential for locating host plants and hunting prey in many species of insects, such as the moth Deilephila elpenor and the wasp Polybia sericea, respectively.

References

  1. 1 2 Vosshall LB (2015). "Laying a controversial smell theory to rest". Proc. Natl. Acad. Sci. USA. 112 (21): 6525–6526. Bibcode:2015PNAS..112.6525V. doi: 10.1073/pnas.1507103112 . PMC   4450429 . PMID   26015552.
  2. Horsfield, A. P.; Haase, A.; Turin, L. (2017). "Molecular recognition in olfaction". Advances in Physics: X. 2 (3): 937–977. doi: 10.1080/23746149.2017.1378594 . hdl: 10044/1/50604 .
  3. 1 2 3 4 Block, E. (2018). "Molecular basis of mammalian odor discrimination: A status report". Journal of Agricultural and Food Chemistry. 66 (51): 13346–13366. doi:10.1021/acs.jafc.8b04471. PMID   30453735. S2CID   53873781.
  4. Ball, Philip (7 December 2006). "Access : Rogue theory of smell gets a boost : Nature News". Nature: news061204–10. doi:10.1038/news061204-10 . Retrieved 2008-04-11.
  5. 1 2 3 4 5 Turin L (1996). "A spectroscopic mechanism for primary olfactory reception" (PDF). Chem. Senses. 21 (6): 773–91. doi: 10.1093/chemse/21.6.773 . PMID   8985605.
  6. Burr, C. (2003). The Emperor of Scent: A Story of Perfume, Obsession, and the Last Mystery of the Senses . Random House. ISBN   978-0-375-50797-7.
  7. Gilbert, A. (2003). "The Emperor's new theory". Nature Neuroscience. 6 (4): 335. doi: 10.1038/nn0403-335 .
  8. Dyson GM (1928). "Some aspects of the vibration theory of odor". Perfumery and Essential Oil Record. 19: 456–459.
  9. "TED Talks: Luca Turin | The science of scent".
  10. 1 2 3 4 Brookes JC, Hartoutsiou F, Horsfield AP, Stoneham AM (2007). "Could humans recognize odor by phonon assisted tunneling?". Phys. Rev. Lett. 98 (3): 038101. arXiv: physics/0611205 . Bibcode:2007PhRvL..98c8101B. doi:10.1103/PhysRevLett.98.038101. PMID   17358733. S2CID   1519986.
  11. Ball, Philip (14 February 2011). "Flies Sniff Out Heavy Hydrogen". Nature. doi:10.1038/news.2011.39 . Retrieved 16 February 2011.
  12. Eric Block (5 June 2015). "What's that smell? A controversial theory of olfaction deemed implausible". The Conversation.
  13. 1 2 "'Quantum smell' idea gains ground". BBC News. 28 January 2013.
  14. Mark Anderson. "Study Bolsters Quantum Vibration Scent Theory". scientificamerican.com.
  15. 1 2 "Controversial theory of smell given a boost". rsc.org.
  16. "In science, no work is completed until it has been picked to pieces". newstatesman.com. 10 June 2021.
  17. 1 2 3 4 5 Block E, et al. (2015). "Implausibility of the Vibrational Theory of Olfaction". Proc. Natl. Acad. Sci. USA. 112 (21): E2766–E2774. Bibcode:2015PNAS..112E2766B. doi: 10.1073/pnas.1503054112 . PMC   4450420 . PMID   25901328.
  18. "Quantum mechanics may explain how humans smell".
  19. 1 2 Haffenden LJ, Yaylayan VA, Fortin J (2001). "Investigation of vibrational theory of olfaction with variously labelled benzaldehydes". Food Chem. 73 (1): 67–72. doi:10.1016/S0308-8146(00)00287-9.
  20. "David MacKay: Smells: Summary" . Retrieved 2008-04-11.
  21. 1 2 Keller A, Vosshall LB (2004). "A psychophysical test of the vibration theory of olfaction". Nature Neuroscience. 7 (4): 337–338. doi:10.1038/nn1215. PMID   15034588. S2CID   1073550.
  22. Havens BR, Melone CD (1995). "The application of deuterated sex pheromone mimics of the American cockroach (Periplaneta americana, L.), to the study of wright's vibrational theory of olfaction" (PDF). Dev. Food. Sci. Developments in Food Science. 37 (1): 497–524. doi:10.1016/S0167-4501(06)80176-7. ISBN   9780444820136.
  23. Hara J (1977). "Olfactory discrimination between glycine and deuterated glycine by fish". Experientia. 33 (5): 618–9. doi:10.1007/BF01946534. PMID   862794. S2CID   5203445.
  24. Flies sniff out heavy hydrogen Nature 14 February 2011.
  25. 1 2 Franco, M. I.; Turin, L.; Mershin, A. & Skoulakis, E. M. C. (2011). "Molecular vibration-sensing component in Drosophila melanogaster olfaction". Proceedings of the National Academy of Sciences of the USA. 108 (9): 3797–3802. Bibcode:2011PNAS..108.3797F. doi: 10.1073/pnas.1012293108 . PMC   3048096 . PMID   21321219.
  26. Schramm, V. L. (2007). "Binding isotope effects: boon and bane". Curr. Opin. Chem. Biol. 11 (5): 529–536. doi:10.1016/j.cbpa.2007.07.013. PMC   2066183 . PMID   17869163.
  27. 1 2 Courtland, Rachel (14 February 2011). "Fly sniffs molecule's quantum vibrations". New Scientist. Retrieved 16 February 2011.
  28. 1 2 "Rogue Odour Theory Could Be Right" . Retrieved 2008-04-11.
  29. Takane SY, Mitchell JB (2004). "A structure-odour relationship study using EVA descriptors and hierarchical clustering". Org. Biomol. Chem. 2 (22): 3250–5. doi:10.1039/B409802A. PMID   15534702.
  30. Oka Y, Nakamura A, Watanabe H, Touhara K (2004). "An odorant derivative as an antagonist for an olfactory receptor". Chem. Senses. 29 (9): 815–22. doi:10.1093/chemse/bjh247. PMID   15574817.
  31. Doszczak, L; Kraft, P; Weber, H-P; Bertermann, R; Triller, A; Hatt, H; Reinhold Tacke, R (2007). "Prediction of Perception: Probing the hOR17-4 Olfactory Receptor Model with Silicon Analogues of Bourgeonal and Lilial". Angew. Chem. Int. Ed. 46 (18): 3367–3371. doi:10.1002/anie.200605002. PMID   17397127.
  32. Pirrung, MC; Ghorai, S; Ibarra-Rivera, TR (2009). "Multicomponent Reactions of Convertible Isonitriles". J. Org. Chem. 74 (11): 4110–4117. doi:10.1021/jo900414n. PMID   19408909.
  33. Nishide, K; Miyamoto, T; Kumar, K; Ohsugi, S-I; Node, M (2002). "Synthetic equivalents of benzenethiol and benzyl mercaptan having faint smell: odor reducing effect of trialkylsilyl group". Tetrahedron Lett. 43 (47): 8569–8573. doi:10.1016/s0040-4039(02)02052-x.
  34. "Testing a radical theory". Nat. Neurosci. 7 (4): 315. 2004. doi: 10.1038/nn0404-315 . PMID   15048113.
  35. "Putting a smell theory to the sniff test", Renee Twombly; Rockefeller Scientist, March 26, 2004
  36. Rinaldi, Andrea (2011). "Do Vibrating Molecules Give Us Our Sense of Smell?". Science Now. Retrieved 2011-02-17.
  37. Ball, Philip (2011). "Flies Sniff out heavy hydrogen". Nature. doi:10.1038/news.2011.39 . Retrieved 2011-02-17.
  38. 1 2 3 Gane, S; Georganakis, D; Maniati, K; Vamvakias, M; Ragoussis, N; Skoulakis, EMC; Turin, L (2013). "Molecular vibration-sensing component in human olfaction". PLOS ONE. 8 (1): e55780. Bibcode:2013PLoSO...855780G. doi: 10.1371/journal.pone.0055780 . PMC   3555824 . PMID   23372854.
  39. New study strengthens olfactory vibration-sensing theory Bob Yirka, Physorg, January 29, 2013
  40. Everts S (2015). "Receptor Research Reignites A Smelly Debate". Chem. Eng. News. 93 (18): 29–30.
  41. Turin, L; Gane, S; Georganakis, D; Maniati, K; Skoulakis, EMC (2015). "Plausibility of the Vibrational Theory of Olfaction". Proc. Natl. Acad. Sci. USA. 112 (25): E3154. Bibcode:2015PNAS..112E3154T. doi: 10.1073/pnas.1508035112 . PMC   4485082 . PMID   26045494.
  42. Block E; Jang, S; Matsunami, H; Batista, VS; Zhuang, H (2015). "Reply to Turin et al.: Vibrational theory of olfaction is implausible". Proc. Natl. Acad. Sci. USA. 112 (25): E3155. Bibcode:2015PNAS..112E3155B. doi: 10.1073/pnas.1508443112 . PMC   4485112 . PMID   26045493.
  43. Saberi M, Seyed-allaei (2016). "Odorant receptors of Drosophila are sensitive to the molecular volume of odorants". Scientific Reports. 6: 25103. Bibcode:2016NatSR...625103S. doi:10.1038/srep25103. PMC   4844992 . PMID   27112241.
  44. Na, Mihwa; Liu, Min Ting; Nguyen, Minh Q.; Ryan, Kevin (2019-01-16). "Single-Neuron Comparison of the Olfactory Receptor Response to Deuterated and Nondeuterated Odorants". ACS Chemical Neuroscience. 10 (1): 552–562. doi: 10.1021/acschemneuro.8b00416 . PMID   30343564.