Tonotopy

In physiology, tonotopy (from Greek tono=frequency and topos = place) is the spatial arrangement of where sounds of different frequency are processed in the brain. Tones close to each other in terms of frequency are represented in topologically neighbouring regions in the brain. Tonotopic maps are a particular case of topographic organization, similar to retinotopy in the visual system.

Tonotopy in the auditory system begins at the cochlea, the small snail-like structure in the inner ear that sends information about sound to the brain. Different regions of the basilar membrane in the organ of Corti, the sound-sensitive portion of the cochlea, vibrate at different sinusoidal frequencies due to variations in thickness and width along the length of the membrane. Nerves that transmit information from different regions of the basilar membrane therefore encode frequency tonotopically. This tonotopy then projects through the vestibulocochlear nerve and associated midbrain structures to the primary auditory cortex via the auditory radiation pathway. Throughout this radiation, organization is linear with relation to placement on the organ of Corti, in accordance to the best frequency response (that is, the frequency at which that neuron is most sensitive) of each neuron. However, binaural fusion in the superior olivary complex onward adds significant amounts of information encoded in the signal strength of each ganglion. Thus, the number of tonotopic maps varies between species and the degree of binaural synthesis and separation of sound intensities; in humans, six tonotopic maps have been identified in the primary auditory cortex.[1]

History

The earliest evidence for tonotopic organization in auditory cortex was indicated by Vladimir E. Larionov in an 1899 paper entitled "On the musical centers of the brain", which suggested that lesions in an S-shaped trajectory resulted in failure to respond to tones of different frequencies.[2] By the 1920s, cochlear cochlear anatomy had been described and the concept of tonotopicity had been introduced.[3] At this time, Hungarian biophysicist, Georg von Békésy began further exploration of tonotopy in the auditory cortex. Békésy measured the cochlear traveling wave by opening up the cochlea widely and using a strobe light and microscope to visually observe the motion on a wide variety of animals including guinea pig, chicken, mouse, rat, cow, elephant, and human temporal bone.[4] Importantly, Békésy found that different sound frequencies caused maximum wave amplitudes to occur at different places along the basilar membrane along the coil of the cochlea, which is the fundamental principal of tonotopy. Békésy was awarded the  Nobel Prize in Physiology or Medicine  for his work. In 1946, the first live demonstration of tonotopic organization in auditory cortex occurred at John Hopkins Hospital.[5] More recently, advances in technology have allowed researchers to map the tonotopic organization in healthy human subjects using electroencephalographic (EEG) and magnetoencephalographic (MEG) data. While most human studies agree on the existence of a tonotopic gradient map in which low frequencies are represented laterally and high frequencies are represented medially around Heschl's gyrus, a more detailed map in human auditory cortex is not yet firmly established due to methodological limitations[6]

Sensory mechanisms

Peripheral nervous system

Cochlea

Tonotopic organization in the cochlea forms throughout pre- and post-natal development through a series of changes that occur in response to auditory stimuli.[7] Research suggests that the pre-natal establishment of tonotopic organization is partially guided by synaptic reorganization; however, more recent studies have shown that the early changes and refinements occur at both the circuit and subcellular levels.[8] In mammals, after the inner ear is otherwise fully developed, the tonotopic map is then reorganized in order to accommodate higher and more specific frequencies.[9] Research has suggested that the receptor guanylyl cyclase Npr2 is vital for the precise and specific organization of this tonotopy.[10] Further experiments have demonstrated a conserved role of Sonic Hedgehog emanating from the notochord and floor plate in establishing tonotopic organization during early development.[11] It is this proper tonotopic organization of the hair cells in the cochlea that allows for correct perception of frequency as the proper pitch.[12]

Structural organization

In the cochlea, sound creates a traveling wave that moves from base to apex, increasing in amplitude as it moves along a tonotopic axis in the basilar membrane (BM).[13] This pressure wave travels along the BM of the cochlea until it reaches an area that corresponds to its maximum vibration frequency; this is then coded as pitch.[13] High frequency sounds stimulate neurons at the base of the structure and lower frequency sounds stimulate neurons at the apex.[13] This represents cochlear tonotopic organization. This occurs because the mechanical properties of the BM are graded along a tonotopic axis; this conveys distinct frequencies to hair cells (mechanosensory cells that amplify cochlear vibrations and send auditory information to the brain), establishing receptor potentials and, consequently frequency tuning.[13] For example, the BM increases in stiffness towards its base.

Mechanisms of cochlear tonotopy

Hair bundles, or the “mechanical antenna” of hair cells, are thought to be particularly important in cochlear tonotopy.[13] The morphology of hair bundles likely contributes to the BM gradient. Tonotopic position determines the structure of hair bundles in the cochlea.[14] The height of hair bundles increases from base to apex and the number of stereocilia decreases (i.e. hair cells located at the base of the cochlea contain more stereo cilia than those located at the apex).[14]

Furthermore, in the tip-link complex of cochlear hair cells, tonotopy is associated with gradients of intrinsic mechanical properties.[15] In the hair bundle, gating springs determine the open probability of mechanoelectrical ion transduction channels: at higher frequencies, these elastic springs are subject to higher stiffness and higher mechanical tension in tip-links of hair cells.[14]  This is emphasized by the division of labor between outer and inner hair cells, in which mechanical gradients for outer hair cells (responsible for amplification of lower frequency sounds) have higher stiffness and tension.[15]

Tonotopy also manifests in the electrophysical properties of transduction.[15] Sound energy is translated into neural signals through mechanoelectrical transduction. The magnitude of peak transduction current varies with tonotopic position. For example, currents are largest at high frequency positions such as the base of cochlea.[16] As noted above, basal cochlear hair cells have more stereocilia, thus providing more channels and larger currents.[16] Tonotopic position also determines the conductance of individual transduction channels. Individual channels at basal hair cells conduct more current than those at apical hair cells.[17]

Finally, sound amplification is greater in the basal than in the apical cochlear regions because outer hair cells express the motor protein prestin, which amplifies vibrations and increases sensitivity of outer hair cells to lower sounds.[13]

Central nervous system

Cortex

Audio frequency, otherwise known as the pitch, is currently the only characteristic of sound that is known with certainty to be topographically mapped in the central nervous system. However, other characteristics may form similar maps in the cortex such as sound intensity,[18][19] tuning bandwidth,[20] or modulation rate,[21][22][23] but these have not been as well studied.

In the midbrain, there exist two primary auditory pathways to the auditory cortex—the lemniscal classical auditory pathway and the extralemniscal non-classical auditory pathway.[24] The lemniscal classical auditory pathway is tonotopically organized and consists of the central nucleus of the inferior colliculusand the ventral medial geniculate bodyprojecting to primary areas in the auditory cortex. The non-primary auditory cortex receives inputs from the extralemniscal non-classical auditory pathway, which shows a diffuse frequency organization.[24]

The tonotopic organization of the auditory cortexhas been extensively examined and is therefore better understood compared to other areas of the auditory pathway.[24] Tonotopy of the auditory cortex has been observed in many animal species including birds, rodents, primates, and other mammals.[24] In mice, four subregions of the auditory cortex have been found to exhibit tonotopic organization. The classically divided A1 subregion has been found to in fact be two distinct tonopic regions—A1 and the dorsomedial field (DM).[25] Auditory cortex region A2 and the anterior audiotry field (AAF) both have tonotopic maps that runs dorsoventrally.[25] The other two regions of the mouse auditory cortex, the dorsoanterior field (DA) and the dorsoposterior field (DP) are non-tonotopic. While neurons in these non-tonotopic regions have a characteristic frequency, they are arranged randomly.[26]

Studies using non-human primates have generated a hierarchical model of auditory cortical organization consisting of an elongated core consisting of three back-to-back tonotopic fields—the primary auditory field A1, the rostral field R, and the rostral temporal field RT. These regions are surrounded by belt fields (secondary) regions and higher-order parabelt fields.[27] A1 exhibits a frequency gradient from high to low in the posterior-to-anterior direction; R exhibits a reversed gradient with characteristic frequencies from low to high in the posterior-to-anterior direction. RT has a less clearly organized gradient from high back to low frequencies.[24] These primary tonotopic patterns continuously extend into the surrounding belt areas.[28]

Tonotopic organization in the human auditory cortex has been studied using a variety of non-invasive imaging techniques including magneto- and electroencephalography (MEG/EEG), positron emission tomography (PET), and functional magnetic resonance imaging (fMRI).[29] The primary tonotopic map in the human auditory cortex is along Heschl's gyrus(HG). However, various researchers have reached conflicting conclusions about the direction of frequency gradient along HG. Some experiments found that tonotopic progression ran parallel along HG while others found that the frequency gradient ran perpendicularly across HG in a diagonal direction, forming an angled V-shaped pair of gradients.[24]

In mice

One of the well-established methods of studying tonotopic patterning in the auditory cortex during development is tone-rearing.[30][31] In mouse Primary Auditory Cortex (A1), different neurons respond to different ranges of frequencies with one particular frequency eliciting the largest response – this is known as the "best frequency" for a given neuron.[30] Exposing mouse pups to one particular frequency during the auditory critical period (postnatal day 12 to 15)[30] will shift the "best frequencies" of neurons in A1 towards the exposed frequency tone.[30]

These frequency shifts in response to environmental stimuli have been shown to improve performance in perceptual behavior tasks in adult mice that were tone-reared during auditory critical period.[32][33] Adult learning and critical period sensory manipulations induce comparable shifts in cortical topographies, and by definition adult learning results in increased perceptual abilities.[34] The tonotopic development of A1 in mouse pups is therefore an important factor in understanding the neurological basis of auditory learning.

Other species also show similar tonotopic development during critical periods. Rat tonotopic develop is nearly identical to mouse, but the critical period is shifted slightly earlier,[31] and barn owls show an analogous auditory development in Interaural Time Differences (ITD).[35]

Plasticity of auditory critical period

The auditory critical period of rats, which lasts from postnatal day 11 (P11) to P13[31] can be extended through deprivation experiments such as white noise-rearing.[36] It has been shown that subsets of the tonotopic map in A1 can be held in a plastic state indefinitely by exposing the rats to white noise consisting of frequencies within a particular range determined by the experimenter.[30][31] For example, exposing a rat during auditory critical period to white noise that includes tone frequencies between 7 kHz and 10 kHz will keep the corresponding neurons in a plastic state far past the typical critical period–one study has retained this plastic state until the rats were 90 days old.[30] Recent studies have also found that release of the neurotransmitter norepinephrine is required for critical period plasticity in the auditory cortex, however intrinsic tonotopic patterning of the auditory cortical circuitry occurs independently from norepinephrine release.[37] A recent toxicity study showed that in-utero and postnatal exposure to polychlorinated biphenyl (PCB) altered overall primary auditory cortex (A1) organization, including tonotopy and A1 topography. Early PCB exposure also changed the balance of excitatory and inhibitory inputs, which altered the ability of the auditory cortex to plastically reorganize after changes in the acoustic environment, thereby altering the critical period of auditory plasticity.[38]

Adult plasticity

Studies in mature A1 have focused on neuromodulatory influences and have found that direct and indirect vagus nerve stimulation, which triggers neuromodulator release, promotes adult auditory plasticity.[39] Cholinergic signaling has been shown to engage 5-HT3AR cell activity across cortical areas and facilitate adult auditory plasticity.[40] Furthermore, behavioral training using rewarding or aversive stimuli, commonly known to engage cholinergic afferents and 5-HT3AR cells, has also been shown to alter and shift adult tonotopic maps.[41]

See also

References

  1. Talavage TM, Sereno MI, Melcher JR, Ledden PJ, Rosen BR, Dale AM (March 2004). "Tonotopic organization in human auditory cortex revealed by progressions of frequency sensitivity" (PDF). Journal of Neurophysiology. 91 (3): 1282–96. doi:10.1152/jn.01125.2002. PMID 14614108.
  2. Popper AN, Fay RR (2012-12-06). Comparative Studies of Hearing in Vertebrates. New York, NY. ISBN 978-1461380740. OCLC 1058153919.
  3. Stevens SS (September 1972). "Georg von Békésy". Physics Today. 25 (9): 78–81. Bibcode:1972PhT....25i..78S. doi:10.1063/1.3071029.
  4. von Békésy G, Wever EG (1960). Experiments in hearing. Wever, Ernest Glen, 1902-. New York: McGraw-Hill. ISBN 978-0070043244. OCLC 14607524.
  5. Walzl EM, Woolsey CN (October 1946). "Effects of cochlear lesions on click responses in the auditory cortex of the cat". Bulletin of the Johns Hopkins Hospital. 79 (4): 309–19. PMID 20280876.
  6. Langers DR, van Dijk P (September 2012). "Mapping the tonotopic organization in human auditory cortex with minimally salient acoustic stimulation". Cerebral Cortex. 22 (9): 2024–38. doi:10.1093/cercor/bhr282. PMC 3412441. PMID 21980020.
  7. Mann ZF, Kelley MW (June 2011). "Development of tonotopy in the auditory periphery". Hearing Research. 276 (1–2): 2–15. doi:10.1016/j.heares.2011.01.011. PMID 21276841.
  8. Kandler K, Clause A, Noh J (June 2009). "Tonotopic reorganization of developing auditory brainstem circuits". Nature Neuroscience. 12 (6): 711–7. doi:10.1038/nn.2332. PMC 2780022. PMID 19471270.
  9. "Development alterations in the frequency map of the mammalian cochlea". American Journal of Otolaryngology. 11 (3): 207. May 1990. doi:10.1016/0196-0709(90)90041-s. ISSN 0196-0709.
  10. Lu CC, Cao XJ, Wright S, Ma L, Oertel D, Goodrich LV (December 2014). "Mutation of Npr2 leads to blurred tonotopic organization of central auditory circuits in mice". PLoS Genetics. 10 (12): e1004823. doi:10.1371/journal.pgen.1004823. PMC 4256264. PMID 25473838.
  11. Son EJ, Ma JH, Ankamreddy H, Shin JO, Choi JY, Wu DK, Bok J (March 2015). "Conserved role of Sonic Hedgehog in tonotopic organization of the avian basilar papilla and mammalian cochlea". Proceedings of the National Academy of Sciences of the United States of America. 112 (12): 3746–51. Bibcode:2015PNAS..112.3746S. doi:10.1073/pnas.1417856112. PMC 4378437. PMID 25775517.
  12. Oxenham AJ, Bernstein JG, Penagos H (February 2004). "Correct tonotopic representation is necessary for complex pitch perception". Proceedings of the National Academy of Sciences of the United States of America. 101 (5): 1421–5. doi:10.1073/pnas.0306958101. PMC 337068. PMID 14718671.
  13. Dallos P (1996), "Overview: Cochlear Neurobiology", in Dallos P, Popper AN, Fay RR (eds.), The Cochlea, Springer Handbook of Auditory Research, 8, Springer New York, pp. 1–43, doi:10.1007/978-1-4612-0757-3_1, ISBN 9781461207573
  14. LeMasurier M, Gillespie PG (November 2005). "Hair-cell mechanotransduction and cochlear amplification". Neuron. 48 (3): 403–15. doi:10.1016/j.neuron.2005.10.017. PMID 16269359.
  15. Tobin M, Chaiyasitdhi A, Michel V, Michalski N, Martin P (April 2019). "Stiffness and tension gradients of the hair cell's tip-link complex in the mammalian cochlea". eLife. 8: e43473. doi:10.7554/eLife.43473. PMC 6464607. PMID 30932811.
  16. He DZ, Jia S, Dallos P (June 2004). "Mechanoelectrical transduction of adult outer hair cells studied in a gerbil hemicochlea". Nature. 429 (6993): 766–70. Bibcode:2004Natur.429..766H. doi:10.1038/nature02591. PMID 15201911.
  17. Ricci AJ, Crawford AC, Fettiplace R (December 2003). "Tonotopic variation in the conductance of the hair cell mechanotransducer channel". Neuron. 40 (5): 983–90. doi:10.1016/S0896-6273(03)00721-9. PMID 14659096.
  18. Bilecen D, Seifritz E, Scheffler K, Henning J, Schulte AC (October 2002). "Amplitopicity of the human auditory cortex: an fMRI study". NeuroImage. 17 (2): 710–8. doi:10.1016/s1053-8119(02)91133-x. PMID 12377146.
  19. Pantev C, Hoke M, Lehnertz K, Lütkenhöner B (March 1989). "Neuromagnetic evidence of an amplitopic organization of the human auditory cortex". Electroencephalography and Clinical Neurophysiology. 72 (3): 225–31. doi:10.1016/0013-4694(89)90247-2. PMID 2465125.
  20. Seifritz E, Di Salle F, Esposito F, Herdener M, Neuhoff JG, Scheffler K (February 2006). "Enhancing BOLD response in the auditory system by neurophysiologically tuned fMRI sequence". NeuroImage. 29 (3): 1013–22. doi:10.1016/j.neuroimage.2005.08.029. PMID 16253522.
  21. Langner G, Sams M, Heil P, Schulze H (December 1997). "Frequency and periodicity are represented in orthogonal maps in the human auditory cortex: evidence from magnetoencephalography". Journal of Comparative Physiology A. 181 (6): 665–76. doi:10.1007/s003590050148. PMID 9449825.
  22. Herdener M, Esposito F, Scheffler K, Schneider P, Logothetis NK, Uludag K, Kayser C (November 2013). "Spatial representations of temporal and spectral sound cues in human auditory cortex". Cortex; A Journal Devoted to the Study of the Nervous System and Behavior. 49 (10): 2822–33. doi:10.1016/j.cortex.2013.04.003. PMID 23706955.
  23. Barton B, Venezia JH, Saberi K, Hickok G, Brewer AA (December 2012). "Orthogonal acoustic dimensions define auditory field maps in human cortex". Proceedings of the National Academy of Sciences of the United States of America. 109 (50): 20738–43. Bibcode:2012PNAS..10920738B. doi:10.1073/pnas.1213381109. PMC 3528571. PMID 23188798.
  24. Saenz M, Langers DR (January 2014). "Tonotopic mapping of human auditory cortex". Hearing Research. 307: 42–52. doi:10.1016/j.heares.2013.07.016. PMID 23916753.
  25. Tsukano H, Horie M, Bo T, Uchimura A, Hishida R, Kudoh M, Takahashi K, Takebayashi H, Shibuki K (April 2015). "Delineation of a frequency-organized region isolated from the mouse primary auditory cortex". Journal of Neurophysiology. 113 (7): 2900–20. doi:10.1152/jn.00932.2014. PMC 4416634. PMID 25695649.
  26. Guo W, Chambers AR, Darrow KN, Hancock KE, Shinn-Cunningham BG, Polley DB (July 2012). "Robustness of cortical topography across fields, laminae, anesthetic states, and neurophysiological signal types". The Journal of Neuroscience. 32 (27): 9159–72. doi:10.1523/jneurosci.0065-12.2012. PMC 3402176. PMID 22764225.
  27. Hackett TA, Preuss TM, Kaas JH (December 2001). "Architectonic identification of the core region in auditory cortex of macaques, chimpanzees, and humans". The Journal of Comparative Neurology. 441 (3): 197–222. doi:10.1002/cne.1407. PMID 11745645.
  28. Kusmierek P, Rauschecker JP (September 2009). "Functional specialization of medial auditory belt cortex in the alert rhesus monkey". Journal of Neurophysiology. 102 (3): 1606–22. doi:10.1152/jn.00167.2009. PMC 2746772. PMID 19571201.
  29. van Dijk, P., & Langers, D. R. M. (2013). "Mapping Tonotopy in Human Auditory Cortex" In B. C. J. Moore, R. D. Patterson, I. M. Winter, R. P. Carlyon, & H. E. Gockel (Eds.), Basic Aspects of Hearing (Vol. 787, pp. 419–425). https://doi.org/10.1007/978-1-4614-1590-9_46
  30. Barkat TR, Polley DB, Hensch TK (July 2011). "A critical period for auditory thalamocortical connectivity". Nature Neuroscience. 14 (9): 1189–94. doi:10.1038/nn.2882. PMC 3419581. PMID 21804538.
  31. de Villers-Sidani E, Chang EF, Bao S, Merzenich MM (January 2007). "Critical period window for spectral tuning defined in the primary auditory cortex (A1) in the rat" (PDF). The Journal of Neuroscience. 27 (1): 180–9. doi:10.1523/JNEUROSCI.3227-06.2007. PMC 6672294. PMID 17202485.
  32. Han YK, Köver H, Insanally MN, Semerdjian JH, Bao S (September 2007). "Early experience impairs perceptual discrimination". Nature Neuroscience. 10 (9): 1191–7. doi:10.1038/nn1941. PMID 17660815.
  33. Sarro EC, Sanes DH (April 2011). "The cost and benefit of juvenile training on adult perceptual skill". The Journal of Neuroscience. 31 (14): 5383–91. doi:10.1523/JNEUROSCI.6137-10.2011. PMC 3090646. PMID 21471373.
  34. Polley DB, Steinberg EE, Merzenich MM (May 2006). "Perceptual learning directs auditory cortical map reorganization through top-down influences". The Journal of Neuroscience. 26 (18): 4970–82. doi:10.1523/JNEUROSCI.3771-05.2006. PMC 6674159. PMID 16672673.
  35. Knudsen EI (1998). "Capacity for Plasticity in the Adult Owl Auditory System Expanded by Juvenile Experience". Science. 279 (5356): 1531–1533. Bibcode:1998Sci...279.1531K. doi:10.1126/SCIENCE.279.5356.1531. PMID 9488651.
  36. Chang EF, Merzenich MM (April 2003). "Environmental noise retards auditory cortical development". Science. 300 (5618): 498–502. Bibcode:2003Sci...300..498C. doi:10.1126/SCIENCE.1082163. PMID 12702879.
  37. Shepard, K., Liles, L., Weinshenker, D., & Liu, R. (2015). Norepinephrine is necessary for experience-dependent plasticity in the developing mouse auditory cortex. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience,35(6), 2432-7.
  38. T. Kenet, R. C. Froemke, C. E. Schreiner, I. N. Pessah, & M. M. Merzenich. (2007). Perinatal exposure to a noncoplanar polychlorinated biphenyl alters tonotopy, receptive fields, and plasticity in rat primary auditory cortex. Proceedings of the National Academy of Sciences, 104(18), 7646-7651.
  39. Navzer D. Engineer, Jonathan R. Riley, Jonathan D. Seale, Will A. Vrana, Jai A. Shetake, Sindhu P. Sudanagunta, . . . Michael P. Kilgard. (2011). Reversing pathological neural activity using targeted plasticity. Nature, 470(7332), 101-4.
  40. Takesian, A. E., Bogart, L. J., Lichtman, J. W., & Hensch, T. K. (2018). Inhibitory circuit gating of auditory critical-period plasticity. Nature neuroscience, 21(2), 218–227. doi:10.1038/s41593-017-0064-2
  41. Polley, D., Steinberg, E., & Merzenich, M. (2006). Perceptual learning directs auditory cortical map reorganization through top-down influences. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience, 26(18), 4970-4982.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.