Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1936 2023-07-14 18:37:41 |
2 format correct Meta information modification 1936 2023-07-17 04:28:37 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Burnham, R. Acoustic Information Exchange. Encyclopedia. Available online: https://encyclopedia.pub/entry/46827 (accessed on 19 May 2024).
Burnham R. Acoustic Information Exchange. Encyclopedia. Available at: https://encyclopedia.pub/entry/46827. Accessed May 19, 2024.
Burnham, Rianna. "Acoustic Information Exchange" Encyclopedia, https://encyclopedia.pub/entry/46827 (accessed May 19, 2024).
Burnham, R. (2023, July 14). Acoustic Information Exchange. In Encyclopedia. https://encyclopedia.pub/entry/46827
Burnham, Rianna. "Acoustic Information Exchange." Encyclopedia. Web. 14 July, 2023.
Acoustic Information Exchange
Edit

The behavioural, physiological, and energetic repercussions for wildlife that result from changes in their soundscapes are increasingly being realized. To understand the effects of changing acoustic landscapes, people first must establish the importance of the acoustic sense for species to transfer information between the environment, con- and heterospecifics, and a receiver, and the functional role of calling in behaviours such as foraging, navigation, mate attraction, and weaning.

acoustic ecology soundscape vocal repertoire

1. The Use of Sound and the Acoustic Modality

The acoustic sense is used by many taxa in a wide range of social and behavioural contexts to send and receive information. The soundscape is the acoustic environment that an individual perceives and responds to. Acoustic cues from this sonic landscape aid navigation, prey detection and capture, and conspecific identification and localization. They also help to identify threats or the intrusion of another species, and are used in territory defense. The acoustic sense can be used to maintain social hierarchies and group cohesion, and aid in mate selection. Sound production shows similarity across taxa; it can engage the larynx to manipulate air flow, or use muscle-driven vibrations or drumming [1][2]. Signal modification is invoked via the vocal tract, tongue, beak, trunk, or alternative sound production spaces (e.g., [3][4]). The morphology of the animal can dictate the energy level of the sound, whereby larger individuals are typically thought to invoke longer, deeper, or louder signals. Indeed, an inverse relationship between the animal’s size and the peak frequency of the calls in their repertoire has been established for many taxa (see [5][6][7]). These morphological adaptations can, therefore, also influence mating signals, and give an indication of fitness or the prominence of a trait to potential mates (e.g., the morphological adaptation hypothesis (MAH) in birds [8][9] and insects [10]). Vocalizations and calling behaviours can also respond to changes in the acoustic environment. Altered ambient noise levels from natural or anthropogenic noise, or altered propagating conditions, can initiate adaptations to how, when, or where an individual calls. The process of sound reception is adapted to each species, and reflects both the medium in which they receive sound and the frequencies they are most sensitive to.
An animal’s vocal repertoire is adapted to maximize conspecific communication, or the exchange of information through acoustic means with others of its own species, sub-species, or group. An individual may utilize a spectrum of calls to retain this contact within a group or between individuals. Calls can be modified by the social, behavioural, or environmental context of the caller, as well as indicating an individual’s group membership, internal state, or the setting under which the call is being made. Courtship calls or song can, for example, play a role in species recognition and help define the acoustic niche definition of a group, predominantly arising from the male vocalizations (see for e.g., [11]). The call structure and diversity of the repertoire could also be an indicator of the size and social structuring of the population. The linguistic niche hypothesis [12] suggests that language complexity in humans reflects the socio-demographic variables of the population or sub-group. It proposes that the complexity of the inflections and lexical constructs used are a reflection of population size [12]. This hypothesis could also possibly have similar applicability to non-human animal communications.
Acoustic signal use can also be informed by the environment that the sounds are emitted into. Pure tone signals, with narrow frequency bandwidths, show greater reverberation. This allows longer, louder transmission—for example, by birds in dense forests [4][13][14]. This differs from frequency-modulated notes or tones that rapidly sweep through a range of frequencies—for instance, in bats’ probing ‘chirps’, which return as a single pulse echo after a time delay [1]. This is in accordance with the Acoustic Adaptation Hypothesis (AAH), whereby the acoustic properties of the environment in which the calls are produced influence the use of call types and the structure of these calls. Signals are selected to maximize efficacy in calling and minimize degradation of the call content as it is transmitted (see for e.g., the meta-analysis by [15] and review in [16]). This hypothesis suggests that species in environments where call propagation might be more dampened or obstructed, or habitats described as ‘closed’, use calls that differ in their frequency extents and peak frequency than those in more ‘open’ areas. Typically, calls in closed habitats are adapted for longer-range propagation [17][18]. Similar to this hypothesis, the sensory drive hypothesis also suggests how perceived differences in the acoustic environment, or the individual’s soundscape, can change their signaling traits and behaviours [19][20]. This hypothesis furthers the AAH by suggesting that calling behaviours are adapted to overcome a distortion or a source of acoustic masking. However, the strength of the relationship between the habitat or soundscape structure and call structure may be obscured by the influence of morphological, physiological, or social variables acting on the caller, also shaping the signals used (e.g., [21] and references therein).
Acoustic environments are dynamic; individuals may use compensatory responses in signal production to overcome noise additions to the ambient sound field. This adaptation in calling in response to the perceived soundscape is in accordance with the Lombard Effect [22], which is typically described as an involuntary increase in vocal amplitude. Lombard-like responses have also been seen to alter the frequency, duration, and repetition rate of calls (see for e.g., [23]), but animal responses to changing acoustic environments are not limited to these adaptations. The Lombard Effect is physiological [24]; changes in humans’ speech due to the Lombard Effect have been noted to differ from ‘loud speech’, when a person simply speaks louder, but the mechanisms in non-human animal taxa are mostly unknown [25]. To understand the impact of noise on wildlife, a description of vocal repertoire acquisition, functional use, and complexity is presented here. There is a provisional discussion of the implications of noise.

2. Acoustic Information Exchange

Although, for many species, call structure and application may seem simple, there can be great complexity in the call parameters and use. Vocal behaviours have innate components, but have aspects that are shaped by the experience of the individual. Repertoires are constrained by phylogeny and morphology (MAH), but call use is reinforced through learning and social interaction. Evidence of this social strengthening of call use and repertoire development arises from individuals that have been removed from their mothers or natal group, whereby call types appear present initially and then are lost (e.g., from gray whales (Eschrichtius robustus) [26]). Periods of ‘babbling’ have been noted in several species across taxa. This occurs during an individual’s first few months of life, when vocal learning occurs, and the adult vocalizations are being acquired (e.g., [27][28][29][30]). Signal units within calls, and their sequence, form, and syntax, may also be a product of learning [31][32]. Characteristics such as frequency sweeps, the onset of calling, and temporal variations in call pattern, amplitude, frequency, length, and repetition are rehearsed. This period of ‘babbling’ may also be marked by the use of adaptive mother–calf calls (‘motherese’ in gray whales [33][34] and bats [35]) as learning occurs.
Species with the capacity for vocal learning acquire their acoustic repertoire by imitation and mimicry [36], with phases of practice and refinement [37][38][39][40][41]. Species capable of this type of call acquisition include songbirds, parrots, hummingbirds, bats, elephants, pinnipeds, and cetaceans [36]. Vocal learning is the acquiring of calls and vocal patterns via a social channel, where a conspecific teacher monitors progress and provides feedback. Following this socially directed learning, the behaviours should persist in the absence of the demonstrator or teacher [42]. Learning could be vertical, whereby the information flows from a parent or more experienced elder to the individual (downwards transmission). It can also be horizontal, which represents peer-directed social learning, which occurs between individuals in the same population group or generation [43]. This can pass on group-specific social traditions in calling as well as the repertoire itself. Learning occurs predominantly during the weaning phase, especially for species with more limited parental investment, whereby the young acquire adult calls and stimulate vocal development. However, it can continue throughout the individual’s lifetime. It can aid the spread of novel behaviours in a population or group (e.g., humpback whales (Megaptera novaeangliae) [44]). Imitation can help with the recognition of individuals and reinforces group cohesion. This then aids in the identification and sharing of resources, mate finding, or within-group recognition. This is especially beneficial in the adaptation to elevated ambient noise [31] or increasing the complexity of sounds and signals used by a group [45].
Deciphering the information coded into calls has been a central area for study in animal communication [46][47][48]. Vocalizations can relay information on the internal and external environment of the caller. The stability of the call structures and their use, and the way in which notes can be formed into patterns, forms the basis of categorizing each species’ repertoire by function. This can begin to be interpreted from the temporal aspects of calling—for example, the season—as well as the social context or behavioural, emotional, or physiological state of the signaler when calling [49].

Signalling and Communicating

In animal communications, a sender produces a signal to be perceived and understood, and elicit a response in a receiver [50]. If the signal is an auto-communication [51], such as echolocation in bats and toothed whales, or electrolocation by some fish, it is the interpretation of the echo of the caller’s own signal that carries the information.
Signals project information, without expectation of an acoustic response, although the information conveyed could influence the behaviour of the receiver. Signals also share information about the presence of the signaler or their state of arousal, motivation, or emotion (e.g., [52][53][54]), or could be a display of physical characteristics (e.g., [55][56]). They can convey information about age, group membership, individuality, and fitness (e.g., elephants [3]; bats [57][58]), or the context in which the call is being made. It may be possible, for example, for the receiver to determine whether the caller is in an antagonistic or threatening situation, alone or isolated, or feeling playful/affiliative or aggressive (e.g., mammals [59][60][61][62][63]; birds [64][65]). In addition, the ordering and emphasis of the call components may represent the urgency of the response or priority of actions needed from the receivers. This might range from a warning from a signaler to listeners (e.g., Richardson’s ground squirrel (Spermophilus richardsonii) [66]) to mobbing behaviours (e.g., Carolina chickadees (Poecile carolinensis) [67]). Affiliative calls could be used to aggregate conspecifics or direct social behaviours, such as flight calls in migrating birds (e.g., [68]). They could also direct conspecifics to prey resources (e.g., [69]). These signals may be used to propagate information over great distances, and are structured to be minimally influenced or degraded by the acoustic environment [48].
Communicative calls come with the expectation of an acoustic response from the receiver, as well as possibly modifying behaviour. Calling has been described as ‘maintaining the social life’ of birds [70], with this likely to also be true for other taxa. If vocalizations are part of an interactive exchange or chorus, it is rare for calls to be unanswered [71][72][73]. The initial signaler elicits a response from a receiver, with the context of the sender and receiver, and the interaction between the two, being core to the communication. Modification to calls, such as the amplitude and speed, may be made based on the intended target and their distance from the sender (e.g., zebra finches (Taeniopygia guttata) [70][74]). Vocal communication takes the form of a back-and-forth, give-and-take exchange of information between conspecifics through acoustic means, even from early infancy. Contact calls between conspecifics combine patterns of frequency modulations, rhythmic call series, and amplitude parameters to confirm species, group, or colony membership or encode individual identity [74][75][76].

References

  1. Rossing, T.D. Springer Handbook of Acoustics; Springer Science and Business Media LLC: New York, NY, USA, 2007.
  2. Ladich, F. Effects of noise on sound detection and acoustic communication in fishes. In Animal Communication and Noise; Brumm, H., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 65–90.
  3. Stoeger, A.S.; Heilmann, G.; Zeppelzauer, M.; Ganswindt, A.; Hensman, S.; Charlton, B.D. Visualizing sound emission of elephant vocalisations: Evidence for two rumble production types. PLoS ONE 2012, 7, e48907.
  4. Farina, A. Soundscape Ecology, Principles, Patterns, Methods and Applications; Springer Science+Business Media: Dordrecht, The Netherlands, 2014.
  5. Fletcher, N.H. A simple frequency-scaling rule for animal communication. J. Acoust. Soc. Am. 2004, 115, 2334–2338.
  6. Gillooly, J.F.; Ophir, A.G. The energetic basis of acoustic communication. Proc. R. Soc. B Boil. Sci. 2010, 277, 1325–1331.
  7. Tervo, O.M.; Christoffersen, M.F.; Simon, M.; Miller, L.A.; Jensen, F.H.; Parks, S.E.; Madsen, P.T. High Source Levels and Small Active Space of High-Pitched Song in Bowhead Whales (Balaena mysticetus). PLoS ONE 2012, 7, e52072.
  8. Seddon, N. 2005 Ecological adaptation and species recognition drives vocal evolution in neotropical suboscine birds. Evolution 2005, 59, 200–215.
  9. Ballentine, B. Morphological adaptation influences the evolution of a mating signal. Evolution 2006, 60, 1936–1944.
  10. Arnqvist, G.; Rowe, L. Sexual conflict and arms races between the sexes: A morphological adaptation for control of mating in a female insect. Proc. R. Soc. B Boil. Sci. 1995, 261, 123–127.
  11. Hoikkala, A.; Klappert, K.; Mazzi, D. Factors affecting male song evolution in Drosophilia montana. Curr. Top. Dev. Biol. 2005, 67, 225–250.
  12. Dale, R.; Lupyan, G. Understanding the origins of morphological diversity: The linguistic niche hypothesis. Adv. Complex Syst. 2012, 15, 1150017.
  13. Slabberkoorn, H.; Smith, T.B. Bird song, ecology and speciation. Philos. Trans. R. Soc. Lond. B 2002, 357, 493–503.
  14. Heimann, D. Numerical simulations of wind and sound propagation through an idealized stand of trees. Acta Acust. 2003, 89, 779–788.
  15. Boncoraglio, G.; Saino, N. Habitat structure and the evolution of bird song: A meta-analysis of the evidence for the acoustic adaptation hypothesis. Funct. Ecol. 2007, 21, 134–142.
  16. Ey, E.; Fischer, J. The “acoustic adaptation hypothesis”: A review of the evidence from birds, anurans and mammals. Bioacoustics 2009, 19, 21–48.
  17. Morton, E.S. Ecological Sources of Selection on Avian Sounds. Am. Nat. 1975, 109, 17–34.
  18. Hansen, P. Vocal learning: Its role in adapting sound structures to long-distance propagation, and a hypothesis on its evolution. Anim. Behav. 1979, 27, 1270–1271.
  19. Endler, J.A. Signals, signal conditions, and the direction of evolution. Am. Nat. 1992, 139, 125–153.
  20. Schaefer, H.; Ruxton, G. Signal diversity, sexual selection, and speciation. Annu. Rev. Ecol. Evol. Syst. 2015, 46, 573–592.
  21. Arasco, A.G.; Manser, M.; Watson, S.K.; Kyabulima, S.; Radford, A.N.; Cant, M.A.; Garcia, M. Testing the acoustic adaptation hypothesis with vocalizations from three mongoose species. Anim. Behav. 2022, 187, 71–95.
  22. Lombard, E. Le signe de l’elevation de la voix (“The sign of the rise in the voice”) annals maladiers oreille. Larynx Nez Pharynx 1911, 37, 101–119.
  23. Stowe, L.M.; Golob, E.J. Evidence that the Lombard effect is frequency-specific in humans. J. Acoust. Soc. Am. 2013, 134, 640–647.
  24. Junqua, J.-C. The influence of acoustics on speech production: A noise-induced stress phenomenon known as the Lombard reflex. Speech Commun. 1996, 20, 13–22.
  25. Zollinger, S.A.; Brumm, H. The Lombard effect. Curr. Biol. 2011, 21, R614.
  26. Wisdom, S.; Bowles, A.E.; Anderson, K.E. Development of behaviour and sound repertoire of a rehabilitating gray whale calf. Aquat. Mamm. 2001, 27, 239–255.
  27. Lenneberg, E.H. Biological Foundations of Language; Wiley: New York, NY, USA, 1967.
  28. Marler, P.; Peters, S. Structural changes in song ontogeny in the swamp sparrow Melospiza georgiana. Auk 1981, 99, 446–458.
  29. McCowan, B.; Reiss, D. Vocal learning in captive bottlenose dolphins: A comparison with humans and nonhuman animals. In Social Influences on Vocal Development; Snowdon, C.T., Hausberger, M., Eds.; Cambridge University Press: Cambridge, UK, 1997; pp. 178–207.
  30. Lipkind, D.; Marcus, G.F.; Bemis, D.K.; Sasahara, K.; Jacoby, N.; Takahasi, M.; Suzuki, K.; Feher, O.; Ravbar, P.; Okanoya, K.; et al. Stepwise acquisition of vocal combinatorial capacity in songbirds and human infants. Nature 2013, 498, 104–108.
  31. Janik, V.M.; Slater, P.J. Vocal Learning in Mammals. Adv. Study Behav. 1997, 26, 59–99.
  32. Podos, J.; Nowicki, S.; Peters, S. Permissiveness in the learning and development of song syntax in swamp sparrows. Anim. Behav. 1999, 58, 93–103.
  33. Ollervides, F.J. Gray Whales and Boat Traffic: Movement, Vocal, and Behavioral Responses in Bahia Magdalena, Mexico. Ph.D. Thesis, Texas A and M University, Dallas, TX, USA, 2001.
  34. Charles, S.M. Social Context of Gray Whale Eschrichtius robustus Sound Activity. Master’s Thesis, Texas A and M University, Dallas, TX, USA, 2011.
  35. Fernandez, A.A.; Knörnschild, M. Pup directed vocalisations of adult females and males in a vocal learning bat. Front. Ecol. Evol. 2020, 8, 265.
  36. Jarvis, E.D. Selection for and against vocal learning in birds and mammals. Ornithol. Sci. 2006, 5, 5–14.
  37. Ramus, F.; Hauser, M.D.; Miller, C.; Morris, D.; Mehler, J. Language discrimination by human newborns and by cotton-top tamarin monkeys. Science 2000, 288, 349–351.
  38. Toro, J.M.; Trobalon, J.B.; Sebastian-Galles, N. The use of prosodic cues in language discrimination tasks by rats. Anim. Cogn. 2003, 6, 131–136.
  39. Toro, J.M.; Trobalon, J.B.; Sebastian-Galles, N. Effects of backward speech and speaker variability in language discrimination by rats. J. Exp. Psychol. Anim. B 2005, 31, 95–100.
  40. Tincoff, R.; Hauser, M.; Tsao, F.; Spaepen, G.; Ramus, F.; Mehler, J. The role of speech rhythm in language discrimination: Further tests with a non-human primate. Dev. Sci. 2005, 8, 26–35.
  41. Naoi, N.; Watanabe, S.; Maekawa, K.; Hibiya, J. Prosody discrimination by songbirds (Padda oryzivora). PLoS ONE 2012, 7, e47446.
  42. Wynne, C.D.L.; Udell, M.A.R. Animal Cognition: Evolution, Behavior and Cognition, 2nd revised ed.; Palgrave Macmillan: Basingstoke, UK, 2013.
  43. Fox, E.A.; Sitompul, A.F.; Van Schaik, C.P. Intelligent tool use in wild Sumatran orangutans. In The Mentalities of Gorillas and Orangutans: Comparative Perspectives; Parker, S.T., Mitchell Miles, H.L., Eds.; Cambridge University Press: Cambridge, UK, 1999; pp. 99–116.
  44. Guinee, L.N.; Chu, K.; Dorsey, E.M. Changes over time in the songs of known individual humpback whales (Megaptera novaeangliae). In Communication and Behavior of Whales; Payne, R., Ed.; Westview Press: Boulder, CO, USA, 1999; pp. 59–80.
  45. Janik, V.M. Pitfalls in the categorization of behaviour: A comparisons of dolphin whistle classification methods. Anim. Behav. 2013, 57, 133–143.
  46. Becker, P.H. The coding of species-specific characteristics in bird sounds. In Ecology and Evolution of Acoustic Communication in Birds; Kroodsma, D.E., Miller, E.H., Eds.; Cornell University Press: Ithaca, NY, USA, 1999; pp. 136–159.
  47. Kroodsma, D.E.; Byers, B.E. The Function(s) of Bird Song. Am. Zool. 1991, 31, 318–328.
  48. Bradbury, J.W.; Vehrencamp, S.L. Principle of Animal Communication; Sinauer Associates: Sunderland, MA, USA, 1998.
  49. Curé, C.; Mathevon, N.; Mundry, R.; Aubin, T. Acoustic cues used for species recognition can differ between sexes and sibling species: Evidence in shearwaters. Anim. Behav. 2012, 84, 239–250.
  50. Maynard-Smith, J.; Harper, D. Animal Signals; Oxford University Press: Oxford, UK, 2003.
  51. Jones, T.K.; Allen, K.A.; Moss, C.F. Communication with self, friends and foes in active sensing animals. J. Exp. Biol 2021, 224, jeb242637.
  52. Zahavi, A. The pattern of vocal signals and the information they convey. Behaviour 1982, 80, 1–8.
  53. Ehret, G. Infant rodent ultrasounds? A gate to the understanding of sound communication. Behav. Genet. 2005, 35, 19–29.
  54. Brudzynski, S.M. Ultrasonic calls of rats as indicator variables of negative or positive states: Acetylcholine–dopamine interaction and acoustic coding. Behav. Brain Res. 2007, 182, 261–273.
  55. Harris, T.R.; Fitch, W.T.; Goldstein, L.M.; Fashing, P.J. Black and White Colobus Monkey (Colobus guereza) roars as a source of both honest and exaggerated information about body mass. Ethology 2006, 112, 911–920.
  56. Pfefferle, D.; Fischer, J. Sounds and size: Identification of acoustic variables that reflect body size in hamadryas baboons, Papio hamadryas. Anim. Behav. 2006, 72, 43–51.
  57. Barclay, R.M.R. Interindividual use of echolocation calls: Eavesdropping by bats. Behav. Ecol. Sociobiol. 1982, 10, 271–275.
  58. Masters, W.; Raver, K.A.; Kazial, K.A. Sonar signals of big brown bats, Eptesicus fuscus, contain information about individual identity, age and family affiliation. Anim. Behav. 1995, 50, 1243–1260.
  59. Seyfarth, R.M.; Cheney, D.L.; Marler, P. Vervet monkey alarm calls: Semantic communication in a free-ranging primate. Anim. Behav. 1980, 28, 1070–1094.
  60. Yin, S.; McCowan, B. Barking in domestic dogs: Context specificity and individual identification. Anim. Behav. 2004, 68, 343–355.
  61. Yeon, S.C. The vocal communication of canines. J. Vet. Behav. 2007, 2, 141–144.
  62. Handelman, B. Canine Behavior: A Photo Illustrated Handbook; Dogwise Publishing: Wenatchee, WA, USA, 2012; ISBN 0976511827.
  63. Pongrácz, P.; Molnár, C.; Miklósi, Á.; Csányi, V. Human listeners are able to classify dog (Canis familiaris) barks recorded in different situations. J. Comp. Psychol. 2005, 119, 136–144.
  64. Ficken, M.S.; Popp, J. A comparative analysis of passerine mobbing calls. Auk 1996, 113, 370–380.
  65. Evans, C.S.; Evans, L.; Marler, P. On the meaning of alarm calls: Functional reference in an avian vocal system. Anim. Behav. 1993, 46, 23–38.
  66. Swan, D.C.; Hare, J.F. The first cut is the deepest: Primary syllables of Richardson’s ground squirrel, Spermophilus richardsonii, repeated calls alert receivers. Anim. Behav. 2008, 76, 47–54.
  67. Freeberg, T.M. Complexity in the chick-a-dee call of Carolina chickadees (Poecile carolinensis): Associations of context and signaler behavior to call structure. Auk 2008, 125, 896–907.
  68. Farnsworth, A.; Lovette, I.J. Evolution of nocturnal flight calls in migrating wood-warblers: Apparent lack of morphological constraints. J. Avian Biol. 2005, 36, 337–347.
  69. Evans, C.S.; Marler, P. Food calling and audience effects in male chickens, Gallus gallus: Their relationships to food availability, courtship and social facilitation. Anim. Behav. 1994, 47, 1159–1170.
  70. Marler, P. Bird calls: Their potential for behavioral neurobiology. Ann. N. Y. Acad. Sci. 2004, 1016, 31–44.
  71. Poole, J.H. Sex differences in the behaviour of African elephants. In The Differences between the Sexes; Short, R.V., Balaban, E., Eds.; Cambridge University Press: Cambridge, UK, 1994; pp. 331–346.
  72. Langbauer, W.R. Elephant communication. Zoo Biol. 2000, 19, 425–445.
  73. Maciej, P.; Fischer, J.; Hammerschmidt, K. Transmission characteristics of primate vocalizations: Implications for acoustic analyses. PLoS ONE 2011, 6, e23015.
  74. Lakshminarayanan, K.; Ben Shalom, D.; van Wassenhove, V.; Orbelo, D.; Houde, J.; Poeppel, D. The effect of spectral manipulations on the identification of affective and linguistic prosody. Brain Lang. 2003, 84, 250–263.
  75. Rickheit, G.; Herrmann, T.; Deutsch, W. Psycholinguistik. Ein Internationales Handbuch/Psycholingistic. An International Handbook; Walter de Gruyter: Berlin, Germany, 2003.
  76. Sidtis, D.; Kreiman, J. In the beginning was the familiar voice: Personally familiar voices in the evolutionary and contemporary biology of communication. Integr. Psychol. Behav. Sci. 2011, 46, 146–159.
More
Information
Subjects: Ecology
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 233
Revisions: 2 times (View History)
Update Date: 17 Jul 2023
1000/1000