You are on page 1of 17

Springer

Handbook o
Odor
Buettner
Editor

123
Editor
Andrea Buettner
Fraunhofer Institute for Process Engineering and Packaging
Giggenhauser Str. 35
85354, Freising, Germany
andrea.buettner@ivv.fraunhofer.de

ISBN: 978-3-319-26930-6 e-ISBN: 978-3-319-26932-0


DOI 10.1007/978-3-319-26932-0
Springer Dordrecht Heidelberg London New York
Library of Congress Control Number: 2016953550

Springer International Publishing AG 2017


This work is subject to copyright. All rights are reserved by the Publisher,
whether the whole or part of the material is concerned, specifically
the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilm or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or
hereafter developed.
The use of general descriptive names, registered names, trademarks,
service marks, etc. in this publication does not imply, even in the absence
of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice
and information in this book are believed to be true and accurate at the date
of publication. Neither the publisher nor the authors or the editors give a
warranty, express or implied, with respect to the material contained herein
or for any errors or omissions that may have been made.

Production and typesetting: le-tex publishing services GmbH, Leipzig


Typography and layout: schreiberVIS, Seeheim
Illustrations: Hippmann GbR, Schwarzenbruck
Cover design: eStudio Calamar Steinen, Barcelona
Cover production: WMXDesign GmbH, Heidelberg
Printing and binding: Printer Trento s.r.l., Trento

Printed on acid free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham,
Switzerland
675

Matthias Laska
Human and A 32. Human and Animal
Olfactory Capabilities Compared

32.1 Olfactory Sensitivity ............................. 678


Humans are traditionally considered to have
a poorly developed sense of smell that is clearly 32.2 Olfactory Discrimination Ability ............ 681
inferior to that of nonhuman animals. This view,
however, is mainly based on an interpretation 32.3 Qualitative Comparisons of Olfactory
of neuroanatomical and recent genetic findings, Capabilities Between Species................ 684
32.3.1 Gathering Information
and not on physiological or behavioral evidence.
About the Chemical Environment 684
An increasing number of studies now suggest that
32.3.2 Foraging and Food Selection ...... 684
the human sense of smell is much better than
32.3.3 Spatial Orientation .................... 684
previously thought and that olfaction plays a sig- 32.3.4 Social Communication ............... 684
nificant role in regulating a wide variety of human 32.3.5 Reproduction ............................ 685
behaviors. This chapter, therefore, aims at sum- 32.3.6 Learning and Memory ................ 685
marizing the current knowledge about human
olfactory capabilities and compares them to those 32.4 General Conclusions............................. 685
of animals.
References................................................... 686

Comparing olfactory capabilities between species is not in their ability to distinguish between a given pair of
a trivial task. Several potentially confounding factors odorants from chance to perfect discrimination [32.4].
have to be taken into consideration when trying to make However, here too, more recent studies using state-of-
statements as to differences or similarities concerning the-art psychophysical methods (pioneered by Cain and
the efficiency of the sense of smell between species. coworkers) generally report a considerably lower range
First, we find a high variability in results between between the most and least sensitive subjects within
human studies. With regard to olfactory sensitivity, for a study population [32.5].
example, published mean threshold values for a given Third, it is inevitable to use different methods for

Part D | 32
odorant may vary up to 6 orders of magnitude [32.1]. assessing olfactory performance with animals and hu-
This is probably, at least in part, due to different meth- mans, and also with different species of animals. This
ods used to determine such thresholds. However, there may affect the comparability of results. However, dif-
is a clear trend that more sophisticated psychophysi- ferences between methods employed with different
cal procedures employing signal detection methods and species of animals are usually necessary adaptations for
rigorous stimulus control yield lower threshold values meeting the physiological, anatomical, and behavioral
than more simple procedures [32.2]. needs and limitations of the study species to success-
Second, the vast majority of psychophysical studies fully cooperate in a behavioral test. (In other words,
on the human sense of smell only report mean values even if it was possible to use the same method with dif-
of performance (usually plus a measure of variation), ferent species, this would very likely put one species
but not the distribution of values or their range. This at an advantage over another species, thus invalidating
is all the more problematic as measures of olfactory any comparisons. Therefore, it is better to aim for op-
performance are notorious for a high interindividual timizing a testing method for each study species.) It is
variability. This is true for both studies on olfactory commonly agreed that operant conditioning procedures
sensitivity in which individual threshold values with are the gold standard among the methods employed
a given odorant have been reported to commonly vary with animals for assessing sensory performance [32.6],
up to 3 orders of magnitude within a study popula- and therefore only studies based on such procedures
tion [32.3], and for studies on olfactory discrimination are considered in this chapter. Figures 32.132.5 il-
performance in which subjects have been found to vary lustrate operant conditioning procedures for assessing
676 Part D Odorant Sensing and Physiological Effects

a) c) Fig. 32.1ad Olfactory conditioning


method used with spider monkeys.
(a) Portrait of a spider monkey
(Ateles geoffroyi). (b) The two-
choice apparatus used, viewed from
the animals side. It consists of
two manipulation boxes of which
one is baited with a Kelloggs
honeyloop while the other is empty,
depending on the odorant applied on
the absorbent paper strip attached
d) to the box. (d) A spider monkey
smelling at one of the absorbent
paper strips bearing an odorant used
either as a rewarded stimulus or as
b) an unrewarded stimulus. (c) A spider
monkey indicating its decision
for one of the two simultaneously
presented odorants by opening the
corresponding manipulation box
(courtesy of M. Laska)

a) c) Fig. 32.2ad Olfactory conditioning


method used with squirrel monkeys.
(a) Portrait of a squirrel monkey
(Saimiri sciureus). (b) Eppendorf
cups equipped with absorbent paper
strips. The Eppendorf cups serve as
Part D | 32

manipulation objects (artificial nuts)


that are either baited with a piece
of peanut or not, depending on the
odorant applied on the absorbent
paper strip. (d) Experimental setup.
d) An artificial nut tree is used to present
numerous artificial nuts, half of them
baited with a piece of peanut and
b) bearing an odorant used as rewarded
stimulus, and half of them empty
and bearing an odorant used as
unrewarded stimulus. (c) A squirrel
monkey inspecting an artificial nut
on a branch of the artificial nut tree
(courtesy of M. Laska)
Human and Animal Olfactory Capabilities Compared Human and Animal Olfactory Capabilities Compared 677

a) b) Fig. 32.3ad Olfactory conditioning


method used with Asian elephants
(Elephas maximus). (a) Elephant
sniffing at the left odor port. (b) Ele-
phant sniffing at the right odor port.
(d) Elephant indicating its decision
for one of the two simultaneously
presented odorants by placing her
trunk onto the grid on top of the
corresponding odor port. (c) Elephant
receiving a carrot as a food reward
after a correct choice (courtesy of
M. Laska)

c) d)

a) b) Fig. 32.4ad Olfactory conditioning


method used with South African fur
seals. (a) Schematic drawing of the

Part D | 32
OP2
experimental set up. C: container
OP1 bearing an odor stimulus; V: ventilator
V
for ingoing airflow; O: outlet for
C outgoing airflow; SB: stimulus box;
OP1: odor port 1; OP2: odor port
SB 2. (b) Portrait of a South African
fur seal (Arctocephalus pusillus).
(c) Simultaneous presentation of two
c) d) odor stimuli to a fur seal. (d) A fur
seal sniffing at one of the two odor
ports (courtesy of M. Laska)
678 Part D Odorant Sensing and Physiological Effects

a) b) Fig. 32.5ad Olfactory condition-


Initiation Decision ing procedure used with mice.
S+ odor (a) Schematic drawing of the exper-
Reward imental setup. (b) A mouse (Mus
Light barrier musculus) in front of the odor port.
(d) A mouse poking its head into the
Odor presentation odor port. (c) A mouse licking at the
Odor S odor water spout after a correct choice
applied
(courtesy of M. Laska)

c) d)

olfactory performance in different species of mam- mechanisms underlying possible differences or simi-
mals. larities in olfactory efficiency between species [32.7].
Finally, animal studies usually only employ a low Second, between-species comparisons of olfactory per-
number of individuals, sometimes only one or two an- formance allow us to test hypotheses about the evo-
imals per species, thus making the use of mean values lution of sensory systems and the selective pressures
arguable, and statements as to how representative the acting on them [32.8]. Finally, the integration of ani-
findings are for the whole species difficult. mal and human studies on olfactory performance may
Despite all these difficulties, there are several good help us to better understand medically relevant phe-
reasons that make it worthwhile to compare olfactory nomena such as aging processes or neurodegenerative
capabilities between humans and animals: first, such diseases, which are often accompanied by a loss of ol-
comparisons allow us to study the neural and/or genetic factory function [32.9, 10].
Part D | 32.1

32.1 Olfactory Sensitivity


The sensitivity of the sense of smell is usually as- With the exception of four odorants tested with rats
sessed by determining olfactory detection thresholds, (one explosive substance, and three explosive taggants,
that is, the lowest concentration of a given odorant that is, substances added to explosives to identify their
that a human subject or an animal is able to de- sources) and one odorant tested with mice (pyridazine),
tect. Human olfactory detection thresholds have been all 138 odorants for which olfactory detection thresh-
reported for a total of  3300 odorants [32.1]. In con- olds have been determined with nonhuman mammals
trast, the total number of odorants tested in animals have also been tested with human subjects, allowing for
is much lower. Table 32.1 summarizes the species direct comparisons of their performance.
of mammals and the number of odorants for which In the following figures (as well as in all other
olfactory detection thresholds using operant condition- comparisons that follow), I compare the lowest mean
ing procedures have been published. The total number threshold values reported in human subjects to the
of species is 17 (please note that there exist about lowest individual threshold values reported in a given
 5500 species of mammals), and the highest num- animal species. The rationale for this is as follows:
ber of odorants tested with a given species is 81.
These 17 species represent 7 of the 29 orders of mam- 1. Human studies only rarely report the range of
mals. threshold values, but usually a mean value.
Human and Animal Olfactory Capabilities Compared 32.1 Olfactory Sensitivity 679

Table 32.1 Animal species and number of odorants for which olfactory detection threshold values using operant condi-
tioning procedures have been published
No. Common name Scientific name Mammalian order Number of odorants tested
1 Spider monkey Ateles geoffroyi Primates 81
2 Mouse Mus musculus Rodentia 72
3 Squirrel monkey Saimiri sciureus Primates 61
4 Pigtail macaque Macaca nemestrina Primates 60
5 Rat Rattus norvegicus Rodentia 45
6 Short-tailed fruit bat Carollia perspicillata Chiroptera 18
7 Dog Canis lupus familiaris Carnivora 15
8 Common vampire bat Desmodus rotundus Chiroptera 15
9 Common mouse-eared bat Myotis myotis Chiroptera 13
10 Sea otter Enhydra lutris Carnivora 7
11 Pig Sus scrofa domestica Artiodactyla 5
12 Hedgehog Erinaceus europaeus Insectivora 4
13 Great fruit-eating bat Artibeus lituratus Chiroptera 3
14 Pale spear-nosed bat Phyllostomus discolor Chiroptera 3
15 Common shrew Sorex araneus Insectivora 3
16 Rabbit Oryctolagus cuniculus Lagomorpha 1
17 Harbor seal Phoca vitulina Carnivora 1

Fig. 32.6 Comparison of the olfactory


0 15
6
4 3 detection threshold values (expressed as vapor
4
1 10 8 Low sensitivity phase concentrations) of human subjects
6 13 3
15 1 2
6 5
3 10 for aliphatic n-carboxylic acids and those
2 10 6 1
8
4 10
8 15 14 2 2
of other mammalian species. Data points of
12 the human subjects (brown circles) represent
3 2
Threshold (log ppm)

3 1
2
1
4
12 3 the lowest mean threshold values reported in
4 12 7
the literature, and data points of all animal
2 12
4 1 species (numbered circles) represent the
5
2 lowest threshold values of individual animals
6 reported in the literature. Numbers in circles
High sensitivity

7
refer to the numbers assigned to each species
in Table 32.1 (human data: [32.1]; spider
7
8

Part D | 32.1
7 monkey data: [32.11]; mouse data: [32.12];
7 7 squirrel monkey data: [32.13]; pigtail macaque
9 7

7
data: [32.11]; rat data: [32.14]; short-tailed
10 fruit bat data: [32.15]; dog data: [32.16, 17];
Ethanoic n-Butanoic n-Hexanoic n-Octanoic common vampire bat data: [32.18]; sea otter
acid acid acid acid
data: [32.19]; hedgehog data: [32.20]; great
n-Propanoic n-Pentanoic n-Heptanoic
acid acid acid fruit-eating bat data: [32.18]; pale spear-nosed
bat data: [32.18]; common shrew data: [32.21])

2. Animal studies usually only employ a low num- those from other mammalian species. With the notable
ber of individuals (in some cases only one animal), exception of the dog, human subjects have a higher
making the use of mean values arguable. sensitivity, that is, lower olfactory detection thresh-
3. Comparing the threshold value of the very best in- olds, than the majority of mammalian species tested
dividual animal with the mean threshold value of with this class of odorants. (Please note that all data
a group of human subjects minimizes the risk of un- points for the dog, except the one for n-heptanoic
intentionally favoring humans over animals. acid, are from a study by Neuhaus [32.16] who em-
ployed only one animal. Later studies by other re-
Figure 32.6 compares human olfactory detection searchers [32.17, 57] employed several dogs as well
threshold values for aliphatic n-carboxylic acids to as more rigorous stimulus control, and, interestingly,
680 Part D Odorant Sensing and Physiological Effects

Fig. 32.7 Comparison of the olfactory


3 9
5
9
detection threshold values (expressed as vapor
9

Low sensitivity
2 3 4 phase concentrations) of human subjects
for aliphatic alcohols and those of other
1
8 4 5 mammalian species. Data points of the
3
0 8 5
3 4 2 human subjects (brown circles) represent the
1
Threshold (log ppm)

8 1 lowest mean threshold values reported in


4 3
1 8 the literature, and data points of all animal
5 5 3
3
2 2
6
5 species (numbered circles) represent the
1 4 1 4
11 2 3 4 5 lowest threshold values of individual animals
3 reported in the literature. Numbers in circles
1 1

High sensitivity
4
refer to the numbers assigned to each species
in Table 32.1 (human data: [32.1]; spider
5 monkey data: [32.22]; mouse data: [32.2325];
11 squirrel monkey data: [32.26]; pigtail macaque
6
data: [32.26]; rat data: [32.27, 28]; short-tailed
2 fruit bat data: [32.15]; common vampire
7
Ethanol 1-Butanol 1-Hexanol 1-Octanol bat data: [32.18]; common mouse-eared bat
1-Propanol 1-Pentanol 1-Heptanol data: [32.29]; pig data: [32.30, 31])

Fig. 32.8 Comparison of the olfactory


2
4
Low sensitivity detection threshold values (expressed as vapor
1 3 phase concentrations) of human subjects for
aliphatic acetic esters and those of other
0
5 4 mammalian species. Data points of the
1
16
1 8 3
5
human subjects (brown circles) represent the
Threshold (log ppm)

1 3 1 10 4 6 lowest mean threshold values reported in


2 6
4 5
the literature, and data points of all animal
6 4 6 2
3 2 11 6 species (numbered circles) represent the
4
4 lowest threshold values of individual animals
4 1 3
1 1 3
reported in the literature. Numbers in circles
1 3
High sensitivity

5 3 refer to the numbers assigned to each species


2 5 in Table 32.1 (human data: [32.1]; spider
6
Part D | 32.1

7 monkey data: [32.32]; mouse data: [32.3335];


squirrel monkey data: [32.36]; pigtail macaque
7
data: [32.36]; rat data: [32.3739]; short-
8 tailed fruit bat data: [32.15]; dog data: [32.40];
Ethyl n-Butyl n-Hexyl iso-Pentyl common vampire bat data: [32.18]; sea
acetate acetate acetate acetate
otter data: [32.19]; pig data: [32.31]; rabbit
n-Propyl n-Pentyl iso-Butyl
acetate acetate acetate data: [32.41])

obtained markedly higher threshold values with these tion thresholds, than most of the other species tested.
odorants compared to those reported by Neuhaus.) The It is interesting to note that humans outperform the rat,
mouse, another mammal with a reputation for a keen another mammal believed to have a highly developed
sense of smell, is more sensitive than humans with sense of smell, with all seven 1-alcohols. Similarly, hu-
only three of the seven n-carboxylic acids whereas hu- mans generally have lower olfactory detection thresh-
mans outperform this rodent with four of these seven olds than the bats and nonhuman primates tested with
odorants. these odorants. The pig, in contrast, is clearly more sen-
Figure 32.7 compares human olfactory detection sitive than humans with the two alcohols tested in this
threshold values for aliphatic 1-alcohols to those from species.
other mammalian species. Here, too, human subjects Figure 32.8 compares human olfactory detection
have a higher sensitivity, that is, lower olfactory detec- threshold values for aliphatic acetic esters to those from
Human and Animal Olfactory Capabilities Compared 32.2 Olfactory Discrimination Ability 681

Animals Humans
Fig. 32.9 Comparison of all olfactory detection
more sensitive more sensitive threshold values between human subjects and
Spider monkey 18 57 animal species. Depicted are the number of odor-
Mouse 30 35 ants for which either human subjects or a given
Squirrel monkey 11 50 species of mammal are more sensitive (human
Pigtail macaque 6 54 data: [32.1]; spider monkey data: [32.11, 22, 25,
Rat 10 31
32, 4249]; mouse data: [32.12, 23, 25, 3335,
Short-tailed fruit bat 1 17
4244, 50, 51]; squirrel monkey data: [32.13, 26,
Dog 10 5
Common vampire bat 1 14
36, 4548]; pigtail macaque data: [32.11, 26, 36,
Common mouse-eared bat 0 13 4548]; rat data: [32.14, 27, 28, 3739, 45]; short-
Sea otter 0 7 tailed fruit bat data: [32.15]; dog data: [32.16, 17,
Pig 2 3 40, 5254]; common vampire bat data: [32.18];
Hedgehog 1 3 common mouse-eared bat data: [32.29]; sea otter
Great fruit-eating bat 1 2 data: [32.19]; pig data: [32.30, 31, 55]; hedgehog
Pale spear-nosed bat 1 2 data: [32.20]; great fruit-eating bat data: [32.18];
Common shrew 0 3
pale spear-nosed bat data: [32.18]; common shrew
Rabbit 0 1
data: [32.21]; rabbit data: [32.41]; harbor seal
Harbor seal 1 0
data: [32.56])
40 30 20 10 0 10 20 30 40 50 60
Number of odorants

other mammalian species. With only two exceptions lower olfactory detection thresholds, that is, a higher
(spider monkeys and squirrel monkeys with n-butyl ac- sensitivity with the majority of odorants tested so far
etate), human subjects have lower olfactory detection compared to all other mammal species tested so far.
thresholds, that is, a higher sensitivity for these odorants This includes species traditionally considered to have
than all other mammal species tested, including dogs, a highly developed sense of smell, such as mice, rats,
mice, and rats. With only few exceptions, humans are hedgehogs, shrews, pigs, and rabbits.
also more sensitive for aliphatic acetic esters than spi- It is interesting to note that humans outperform
der monkeys, squirrel monkeys, and pigtail macaques. even the dog, often considered as the undisputed super-
This is remarkable given that these nonhuman primates nose of the animal kingdom, with 5 of the 15 odorants
are highly frugivorous suggesting that a high olfactory tested with both species. The fact that these 5 odorants
sensitivity for fruit-associated odorants such as acetic comprise plant odor components such as -ionone and
esters should be adaptive for these species. However, n-pentyl acetate suggests that the behavioral relevance
both human and nonhuman primates generally outper- of odorants rather than neuroanatomical or genetic fea-
form granivorous species, such as rats, insectivorous tures may strongly affect a species olfactory sensitivity.

Part D | 32.2
species such as bats, and herbivorous species such as This idea is further supported by the fact that 7 of the
the rabbit with this class of odorants. 10 odorants for which the dog has been reported to be
Figure 32.9 summarizes all comparisons of olfac- more sensitive than humans comprise carboxylic acids
tory detection thresholds between human subjects and and thus typical components of the odor of prey species
other mammal species. Depicted are the number of of the dog.
odorants for which either human subjects or a given Thus, based on these comparisons, and contrary to
species of mammal are more sensitive. With the excep- traditional textbook wisdom, humans are not generally
tion of the dog (and the harbor seal, which has been inferior in their olfactory sensitivity compared to ani-
tested with only one odorant), human subjects have mals.

32.2 Olfactory Discrimination Ability


Olfactory discrimination can be defined as the abil- A statistical criterion can then be used to decide
ity to reliably respond differently to the successive whether a human subject or an animal is able to dis-
presentation of two nonidentical odorants. Thus, ol- criminate between the two odorants in question or
factory discrimination is usually assessed by deter- not.
mining the proportion of correct responses with re- Human studies on olfactory discrimination ca-
peated presentations of a given pair of odorants. pabilities usually either employ structurally related
682 Part D Odorant Sensing and Physiological Effects

Table 32.2 Between-species comparison of olfactory discrimination performance with aliphatic odorants sharing the
same functional group but differing in carbon chain length
Human Squirrel Asian Fur seals CD-1 mice Honey bees
subjects monkeys elephants
1-Alcohols +++++ + +++ +++++ ++++++ +++++
n-Aldehydes ++++ +++ +++ +++++ ++++++ ++++++
2-Ketones +++++ +++ +++ +++++ ++++++ ++++++
Acetic esters +++++ +++ +++ ++++ ++++++
n-Carboxylic acids ++++++ +++ +++ +++++ ++++++
Success rate 25/30 13/15 15/15 24/25 30/30 17/18
A + symbol indicates that the group of human subjects or animals succeeded in discriminating a given aliphatic odor pair, a 
symbol indicates failure to do so, and a symbol indicates that this odor pair was not tested. The six symbols in each table cell
refer to the discrimination of carbon chain lengths C4 versus C5 , C4 versus C6 , C4 versus C7 , C5 versus C6 , C5 versus C7 , and C6
versus C7 , respectively. Human data: [32.5860]; squirrel monkey data: [32.59, 61, 62]; Asian elephant data: [32.63, 64]; fur seal
data: [32.65]; mouse data: [32.66]; honeybee data: [32.67]

Table 32.3 Between-species comparison of olfactory discrimination performance with aliphatic odorants sharing the
same carbon length but differing in functional group
Human subjects Squirrel monkeys CD-1 mice Honey bees
1-Alcohols versus n-aldehydes +++ +++ +++ +++
1-Alcohols versus 2-ketones +++ +++ +++ +++
1-Alcohols versus n-carboxylic acids +++ +++ +++
n-Aldehydes versus 2-ketones +++ +++ +++ +++
n-Aldehydes versus n-carboxylic acids +++ +++ +++
2-Ketones versus n-carboxylic acids +++ +++ +++
Success rate 18/18 18/18 18/18 9/9
A + symbol indicates that the group of human subjects or animals succeeded in discriminating a given aliphatic odor pair, a 
symbol indicates failure to do so, and a symbol indicates that this odor pair was not tested. The three symbols in each table cell
refer to the discrimination of odorants that share chain lengths of either 4, or 6, or 8 carbon atoms, respectively. Human data: [32.68];
squirrel monkey data: [32.69]; mouse data: [32.66]; honeybee data: [32.67]

monomolecular odorants (to investigate correlations be- ants, the proportion of successfully discriminated odor
tween structure and perceived quality of odorants), or pairs is slightly lower in humans compared to squirrel
complex odor mixtures of commercial use such as fra- monkeys, fur seals, honey bees, Asian elephants, and
Part D | 32.2

grances (perfumes, body care products) or food odors mice. Nevertheless, humans succeeded with more than
(wines, coffees, artificial flavors). In contrast, the vast 80% of these odor pairs, which are both structurally and
majority of animal studies assessing olfactory discrimi- qualitatively similar to each other.
nation capabilities employ naturally occurring complex Table 32.3 compares the ability of humans and
odor mixtures that are behaviorally relevant for the several species of animals to discriminate between
species under study such as conspecific body odors, aliphatic odorants sharing the same carbon chain length
species-typical food odors, or predator odors. As a con- but differing in functional group. With this set of odor-
sequence, there is only little overlap in the stimuli used ants, not only squirrel monkeys, mice, and honey bees,
between human and animal studies on olfactory dis- but also human subjects successfully discriminated be-
crimination. However, at least a few animal species tween all odor pairs tested.
have also been studied for their ability to distinguish Table 32.4 compares the ability of humans and
between some of the structurally related monomolecu- several species of animals to discriminate between
lar odorants tested with humans. enantiomers. These are pairs of molecules with mirror-
Table 32.2 compares the ability of humans and sev- image structures that exhibit identical chemical and
eral species of animals to discriminate between mem- physical properties except for their optical activity, that
bers of homologous series of aliphatic odorants. These is, rotation of polarized light. They are particularly
are odorants sharing the same oxygen-containing func- useful for assessing how molecular structure is en-
tional group (e.g., an alcohol- or an aldehyde group) but coded by olfactory systems as perceptual differences
differing in carbon chain length. With this set of odor- between enantiomers cannot be caused by differing
Human and Animal Olfactory Capabilities Compared 32.2 Olfactory Discrimination Ability 683

Table 32.4 Between-species comparison of olfactory discrimination performance with enantiomers


Human Squirrel Asian Fur CD-1 Honey Pigtail SD/LE
subjects monkeys elephants seals mice bees macaques rats
Carvone + + + + + + + +
Dihydrocarvone + + + + + +
Dihydrocarveol + + + + + +
Limonene + + +  + + + +
-Pinene + + + + +
Isopulegol   +  + +
Menthol   + + + +
-Citronellol   + + + +
Rose oxide   +  + 
Fenchone  + + + +  +
Limonene oxide   + + + 
Camphor   +  + 
Success rate 5/12 6/12 12/12 8/12 11/11 5/8 5/6 3/3
A + symbol indicates that the group of animals or subjects succeeded in discriminating a given enantiomeric odor pair, and a 
symbol indicates failure to do so. Human data: [32.70, 71]; squirrel monkey data: [32.72, 73]; Asian elephant data: [32.63]; fur seal
data: [32.74]; mouse data: [32.75]; honeybee data: [32.76]; pigtail macaque data: [32.73]; rat data: [32.77, 78]

diffusion rates in the mucus covering the olfactory ep- to be reliably perceived as stronger than the stan-
ithelium or differing airmucus partition coefficients, dard.
but must originate from chiral selectivity at the receptor The human JND has been found to be odorant-
level [32.79]. dependent and Weber fractions have been reported
Human subjects, as a group, are only able to dis- to range from 0:09 for n-pentyl butyrate, 0:16 for
criminate between 5 of the 12 enantiomeric odor pairs n-pentanol, 0:30 for pyridine, 0:35 for n-butanol, to
tested and thus perform similar to squirrel monkeys, 0:47 for phenylethanol [32.8082]. The only study that
which succeeded with 6 out of 12 pairs. Asian ele- directly compared JNDs between species found the We-
phants, mice, and rats, in contrast, succeeded with 12 ber fraction for n-pentyl acetate to be 0:041 in rats and
out of 12, 11 out of 11, and 3 out of 3 of these odor pairs. thus considerably lower than the 0:315 found with this
South African fur seals, pigtail macaques, and honey- odorant in humans [32.83]. However, a later study re-
bees are able to discriminate between the majority, but ported the rats Weber fraction for the same odorant to
not all enantiomeric odor pairs tested with these species be 0:28, and thus comparable to that of humans [32.84].
(Table 32.4). A possible explanation for the extreme paucity of

Part D | 32.2
Based on these comparisons, humans appear to animal data on olfactory JNDs is their difficulty in
have a slightly inferior olfactory discrimination ability learning the concept that two different concentrations
compared to some species such as mice and Asian ele- of the same odorant should have different reward val-
phants. However, their performance in discriminating ues. To understand this difficulty, one must know that
between structurally related monomolecular odorants in operant conditioning procedures an animal learns
appears to be comparable to that of squirrel monkeys, that an odorant A is rewarded and that an odorant B
fur seals, and honeybees. (or a blank stimulus) is not rewarded. Once an ani-
A special, and only rarely investigated, aspect of mal has successfully learned the association between
olfactory discrimination performance is the ability to odorant A and a reward, it will consider this odorant
distinguish between different concentrations of a given as a rewarded stimulus irrespective of its concentration.
odorant. The smallest concentration difference that Biologically, this makes perfect sense as odors that are
a nose can reliably detect is often referred to as just behaviorally relevant for an animal hardly ever change
noticeable difference (JND) and is usually expressed their significance as a function of concentration: a food
as a so-called Weber fraction (according to Webers odor will always be attractive, whether detected at high
law, which states that the JND between two stimuli is or low concentrations, and a predator odor will be al-
proportional to the magnitude of the stimuli: I=I D ways avoided, whether at high or low concentrations.
constant). A Weber fraction of 0:3, for example, in- Thus, it takes an animal extensive training to overcome
dicates that a stimulus has to be presented at a 30% this perseverance with regard to the learned reward
higher intensity, relative to a standard stimulus, in order value of odorants.
684 Part D Odorant Sensing and Physiological Effects

32.3 Qualitative Comparisons of Olfactory Capabilities


Between Species
In addition to using quantifiable measures of olfac- surprisingly, anosmic subjects particularly those who
tory performance, such as sensitivity and discrimination lost their sense of smell instantaneously, for exam-
ability, one can also try to compare the sense of smell ple, due to head trauma often report a consider-
between species using qualitative measures. In the sim- able loss of quality of life with regard to enjoying
plest case, this means to assess whether a given species food [32.91].
possesses a certain ability or not. To this end, it might be
useful to take a look at different behavioral contexts in 32.3.3 Spatial Orientation
which the sense of smell is known to play a role for an-
imals and then to ask whether human subjects are able Quite a number of animal species rely on their sense
to use their noses for the same purpose. of smell for spatial orientation. This is particularly true
for nocturnal and subterranean species. There are two
32.3.1 Gathering Information basic mechanisms that allow animals to find their way
About the Chemical Environment through their habitat using their sense of smell: they
can either use existing landmarks that emit an odor,
The ability to gather information about the chemical en- or they can deposit scent marks themselves at certain
vironment is almost ubiquitous in the animal kingdom. points in their habitat. Both types of odor sources serve
In most species of animals, the detection of chemi- as navigational landmarks that can be used by animals
cal hazards, for example, leads to adaptive behavioral to recognize their position in space.
responses intended to minimize further exposure. Hu- Although humans with an intact sense of vision
mans are no exception to this rule and display adaptive hardly ever need to use their noses for finding their way,
behaviors such as head turning, eye closure, apnea (sus- they do possess the ability to follow scent trails laid
pension of breathing), sneezing, and coughing when by conspecifics and when blindfolded [32.88]. A re-
exposed to harmful volatile chemicals. Although some cent study suggests the existence of spatial information
of these reflex-like responses involve the nasal trigem- processing in the human olfactory system and thus of
inal system, the significance of the olfactory system in an implicit ability for directional smelling [32.92]. Fur-
this context becomes apparent when considering anos- ther, there is anecdotal evidence that blind persons use
mic subjects: humans without a functioning sense of olfactory landmarks as sensory cues for spatial orienta-
smell run a significantly higher risk of suffering from tion [32.93, 94].
gas poisoning [32.85] and of not detecting fire [32.86]
compared to healthy controls. 32.3.4 Social Communication
Part D | 32.3

32.3.2 Foraging and Food Selection Many species of animals have been shown to use body-
borne odors for social communication. Such odors may
The ability to find and select food using the sense convey information about species, social group, sex,
of smell is also widespread among animal species. age, reproductive status, health status, social rank, in-
Human babies, similar to the offspring of other mam- dividual identity, genetic relatedness, dietary habit, and
mals, already display adaptive behavioral responses emotional state of the odor donor.
such as positive head turning and gaping mouth move- Psychophysical studies have demonstrated that hu-
ments when presented with the odor of their mothers mans are able to correctly assign body odors to
breast or the odor of breast milk [32.87]. Although hu- sex [32.95] and to age classes [32.96]. Similarly, it has
mans nowadays hardly ever need to use their noses been shown that humans are able to correctly assign
to forage, that is, to search for food, they do possess body odors to reproductive status [32.97, 98] and to
the ability to follow a food odor scent trail [32.88]. health status [32.99, 100].
The involvement of the sense of smell in human food Further, humans are able to distinguish between in-
selection is more obvious: anosmic subjects run a sig- dividual body odors [32.101] and between body odors
nificantly higher risk of suffering from food poisoning of kin and nonkin [32.102, 103]. This includes mutual
compared to healthy controls with an intact sense of recognition of the body odors between mothers and
smell [32.89]. Similarly, the risk of suffering from their babies [32.104].
malnutrition is markedly higher in anosmic subjects Humans are also able to distinguish between the
compared to normosmic control subjects [32.90]. Not body odors of vegetarians and meat eaters [32.105] and
Human and Animal Olfactory Capabilities Compared 32.4 General Conclusions 685

thus they can sniff out the dietary habits of conspecifics. number of sexual relationships compared to norm-
Recent studies suggest that humans are also able to de- osmic men, and that congenitally anosmic women feel
tect emotions via body odors [32.106, 107]. less secure about their romantic partner compared to
Dogs, mice, and rats are not only able to discrimi- normosmic women [32.115]. Further, anosmia-related
nate between the odors of individual conspecifics, but depression has been shown to reduce sexual appetite in
also between the odors of individuals of other species, humans [32.116].
for example, of individual humans. However, it has been
demonstrated that human subjects are also able to cor- 32.3.6 Learning and Memory
rectly identify the odor of their own pet dog among other
dog odor samples [32.108], to correctly assign body Learning and memory are inevitably linked to sensory
odor samples to individual gorillas [32.109], and to cor- input. Many species of animals rely on olfactory cues
rectly discriminate between the body odors of mouse for learning about their environment or about situa-
strains that only differ from each other in their major his- tional contexts as well as for building and retrieving
tocompatibility complex (MHC) genes [32.110]. memories. Humans are no exception to this as they
are able to rapidly and robustly learn long-lasting as-
32.3.5 Reproduction sociations between food odors and positive or negative
physiological consequences of ingestion [32.117]. Sim-
The ability to correctly identify the reproductive status ilarly, humans are able to rapidly and robustly learn
of potential mates using olfactory cues is widespread appetitive and aversive associations between odors and
among mammals. Similarly, mate choice has been visual stimuli [32.118]. Several studies reported human
shown to be based on or at least influenced by ol- long-term memory for odors to be outstanding and su-
factory cues in a number of mammal species. Human perior to that for other sensory modalities [32.119]. The
males are able to distinguish between female body longest interval tested so far with humans for successful
odors from different phases of the estrous cycle and pre- odor recognition was 1 year [32.120], and for above-
fer the body odor of females produced around the time chance level retention of odorname associations even
of ovulation [32.97, 98, 111]. Several lines of evidence 9 years [32.121].
suggest that human mate choice may be influenced From these qualitative comparisons of olfactory ca-
by body odors and that, as is the case in mice, MHC pabilities, one must conclude that humans have at least
genes are involved in the formation of individual odor the basic ability of using their sense of smell in all be-
signatures [32.112114]. A recent study found that havioral contexts in which animals are known to use
congenitally anosmic men exhibit a strongly reduced their noses.

32.4 General Conclusions

Part D | 32.4
The amount of published data on quantifiable measures mance in discriminating between structurally related
of olfactory performance, which allow for direct com- monomolecular odorants appears to be comparable to
parisons between humans and animals, is rather limited. that of species such as squirrel monkeys, fur seals, and
However, based on this limited set of data the following honeybees.
conclusions can be drawn: Qualitative comparisons of olfactory capabilities
Human subjects have lower olfactory detection suggest that human subjects have at least the basic
thresholds, that is, a higher sensitivity with the major- ability of using their sense of smell in all behavioral
ity of odorants tested so far compared to most of the contexts in which animals are known to use their noses.
mammal species tested so far. This includes species tra- This includes gathering of information about the chem-
ditionally considered to have a highly developed sense ical environment, food selection, spatial orientation,
of smell such as mice, rats, hedgehogs, shrews, pigs, social communication, reproduction, as well as learn-
and rabbits. ing and memory.
Human subjects appear to have a slightly inferior ol- Taken together, these findings suggest that the human
factory discrimination ability compared to species such sense of smell is not generally inferior compared to that
as mice and Asian elephants. However, their perfor- of animals and much better than traditionally thought.
686 Part D Odorant Sensing and Physiological Effects

References

32.1 L.J. van Gemert: Odour Thresholds. Compilations (Desmodus rotundus, Artibeus lituratus, Phyl-
of Odour Threshold Values in Air, Water and Other lostomus discolor), Z. Sugetierkd. 40, 269298
Media, 2nd edn. (OPP, Utrecht 2011) (1975)
32.2 R. Schmidt, W.S. Cain: Making scents: Dynamic 32.19 J. Hammock: Structure, Function and Context: The
olfactometry for threshold measurement, Chem. Impact of Morphometry and Ecology on Olfactory
Senses 35, 109120 (2010) Sensitivity, Ph.D. Thesis (MIT, Cambridge 2005)
32.3 J.C. Stevens, W.S. Cain, R.J. Burke: Variability of 32.20 H. Bretting: Die Bestimmung der Riechschwellen
olfactory thresholds, Chem. Senses 13, 643653 bei Igeln fr Einige Fettsuren, Z. Sugetierkd. 37,
(1988) 286311 (1972)
32.4 M. Laska, A. Ringh: How big is the gap between 32.21 L. Sigmund, F. Sedlacek: Morphometry of the ol-
olfactory detectin and recognition of aliphatic factory organ and olfactory thresholds of some
aldehydes?, Attent. Percept. Psychophys. 72, fatty acids in Sorex araneus, Acta Zool. Fennica
806812 (2010) 173, 249251 (1985)
32.5 J.E. Cometto-Muiz, M.H. Abraham: Structure- 32.22 M. Laska, R.M. Rivas Bautista, L.T. Hernandez
activity relationships on the odor detectability Salazar: Olfactory sensitivity for aliphatic alcohols
of homologous carboxylic acids by humans, Exp. and aldehydes in spider monkeys, Ateles geof-
Brain Res. 207, 7584 (2010) froyi, Am. J. Phys. Anthropol. 129, 112120 (2006)
32.6 J.M. Pearce: Animal Learning and Cognition (Psy- 32.23 J. Larson, J.S. Hoffman, A. Guidotti, E. Costa: Ol-
chology, New York 2008) factory discrimination learning in heterozygous
32.7 B.W. Ache, J.M. Young: Olfaction: Diverse species, reeler mice, Brain Res. 971, 4046 (2003)
conserved principles, Neuron 48, 417430 (2005) 32.24 D.W. Smith, S. Thach, E.L. Marshall, M.G. Men-
32.8 R.A. Barton: Olfactory evolution and behavioral doza, S.J. Kleene: Mice lacking NKCC1 have normal
ecology in primates, Am. J. Primatol. 68, 545558 olfactory sensitivity, Physiol. Behav. 93, 4449
(2006) (2008)
32.9 R.L. Doty, I. Petersen, N. Mensah, K. Christensen: 32.25 P.K. Ltvedt, S.K. Murali, L.T. Hernandez Salazar,
Genetic and environmental influences on odor M. Laska: Olfactory sensitivity for green odors
identification ability in the very old, Psychol. Ag- (aliphatic C6 alcohols and C6 aldehydes) A com-
ing 26, 864871 (2011) parative study in male CD-1 mice (Mus musculus)
32.10 M. Barresi, R. Ciurleo, S. Giacoppo, V.F. Cuzzola, and female spider monkeys (Ateles geoffroyi),
D. Celi, P. Bramanti, S. Marino: Evaluation of Pharmacol. Biochem. Behav. 101, 450457 (2012)
olfactory dysfunction in neurodegenerative dis- 32.26 M. Laska, A. Seibt: Olfactory sensitivity for
eases, J. Neurol. Sci. 323, 1624 (2012) aliphatic alcohols in squirrel monkeys and pig-
32.11 M. Laska, A. Wieser, R.M. Rivas Bautista, L.T. Her- tail macaques, J. Exp. Biol. 205, 16331643 (2002)
nandez Salazar: Olfactory sensitivity for carboxylic 32.27 D.G. Moulton, J.T. Eayrs: Studies in olfactory acu-
acids in spider monkeys and pigtail macaques, ity. 2. Relative detectability of n-aliphatic alco-
Chem. Senses 29, 101109 (2004) hols by the rat, Q. J. Exp. Psychol. 12, 99109
32.12 S.C. Gven, M. Laska: Olfactory sensitivity and (1960)
Part D | 32

odor structure-activity relationships for aliphatic 32.28 D.G. Laing: A comparative study of the olfactory
carboxylic acids in CD-1 mice, PLoS ONE 7, e34301 sensitivity of humans and rats, Chem. Senses Fla-
(2012) vor 1, 257269 (1975)
32.13 M. Laska, A. Seibt, A. Weber: Microsmatic primates 32.29 U. Schmidt, C. Schmidt: Olfactory thresholds in
revisited Olfactory sensitivity in the squirrel four microchiropteran bat species, Proc. 4th Int.
monkey, Chem. Senses 25, 4753 (2000) Bat Res. Conf., Nairobi (1978) pp. 713
32.14 B.M. Slotnick, G.A. Bell, H. Panhuber, D.G. Laing: 32.30 J.B. Jones, C.M. Wathes, K.C. Persaud, R.P. White,
Detection and discrimination of propionic acid R.B. Jones: Acute and chronic exposure to ammo-
after removal of its 2-DG identified major focus nia and olfactory acuity for n-butanol in the pig,
in the olfactory bulb: A psychophysical analysis, Appl. Anim. Behav. Sci. 71, 1328 (2001)
Brain Res. 762, 8996 (1997) 32.31 L.M. Sndergaard, I.E. Holm, M.S. Herskin, F. Dag-
32.15 M. Laska: Olfactory sensitivity to food odor com- naes-Hansen, M.G. Johansen, A.L. Jrgensen,
ponents in the short-tailed fruit bat, Carollia per- J. Ladewig: Determination of odor detection
spicillata (Phyllostomatidae, Chiroptera), J. Comp. threshold in the Gttingen minipig, Chem. Senses
Physiol. A 166, 395399 (1990) 35, 727734 (2010)
32.16 W. Neuhaus: ber die Riechschrfe des Hundes 32.32 L.T. Hernandez Salazar, M. Laska, E. Rodriguez
fr Fettsuren, Z. Vergl. Physiol. 35, 527552 (1953) Luna: Olfactory sensitivity for aliphatic esters in
32.17 D.G. Moulton, D.H. Ashton, J.T. Eayrs: Studies in spider monkeys, Ateles geoffroyi, Behav. Neu-
olfactory acuity. 4. Relative detectability of n- rosci. 117, 11421149 (2003)
aliphatic acids by the dog, Anim. Behav. 8, 117 32.33 V. Vedin, B. Slotnick, A. Berghard: Zonal ablation
128 (1960) of the olfactory sensory neuroepithelium of the
32.18 U. Schmidt: Vergleichende Riechschwellenbes- mouse: Effects on odorant detection, Eur. J. Neu-
timmungen bei neotropischen Chiropteren rosci. 20, 18581864 (2004)
Human and Animal Olfactory Capabilities Compared References 687

32.34 J.C. Walker, R.J. OConnell: Computerized odor indols in three species of nonhuman primates,
psychophysical testing in mice, Chem. Senses 11, J. Exp. Biol. 210, 41694178 (2007)
439453 (1986) 32.49 L. Kjeldmand, L.T. Hernandez Salazar, M. Laska:
32.35 A.C. Clevenger, D. Restrepo: Evaluation of the va- Olfactory sensitivity for sperm-attractant aro-
lidity of a maximum likelihood adaptive staircase matic aldehydes A comparative study in human
procedure for measurement of olfactory detection subjects and spider monkeys, J. Comp. Physiol. A
threshold in mice, Chem. Senses 31, 926 (2006) 197, 1523 (2011)
32.36 M. Laska, A. Seibt: Olfactory sensitivity for 32.50 M. Laska, D. Joshi, G.M. Shepherd: Olfactory sen-
aliphatic esters in squirrel monkeys and pigtail sitivity for aliphatic aldehydes in CD-1 mice, Be-
macaques, Behav. Brain Res. 134, 165174 (2002) hav. Brain Res. 167, 349354 (2006)
32.37 S. Krmer, R. Apfelbach: Olfactory sensitivity, 32.51 L. Larsson, M. Laska: Ultra-high olfactory sensi-
learning and cognition in young adult and aged tivity for the human sperm-attractant aromatic
male Wistar rats, Physiol. Behav. 81, 435442 aldehyde bourgeonal in CD-1 mice, Neurosci. Res.
(2004) 71, 355360 (2011)
32.38 S. Pierson: Conditioned suppression to odorous 32.52 W. Neuhaus: Die Riechschwellen des Hundes fr
stimuli in the rat, J. Comp. Physiol. Psychol. 86, Jonon und thylmercaptan und ihr Verhltnis zu
708717 (1974) anderen Riechschwellen bei Hund und Mensch,
32.39 R.G. Davis: Olfactory psychophysical parameters Z. Naturforsch. 9, 560567 (1954)
in man, rat, dog and pigeon, J. Comp. Physiol. 32.53 P.I. Ezeh, L.J. Myers, L.A. Hanrahan, R.J. Kemp-
Psychol. 85, 221232 (1973) painen, K.A. Cummins: Effects of steroids on the
32.40 D.B. Walker, J.C. Walker, P.J. Cavnar, J.L. Tay- olfactory function of the dog, Physiol. Behav. 51,
lor, D.H. Pickel, S.B. Hall, J.C. Suarez: Naturalis- 11831187 (1992)
tic quantification of canine olfactory sensitivity, 32.54 L.J. Myers, R. Pugh: Threshold of the dog for
Appl. Anim. Behav. Sci. 97, 241254 (2006) detection of inhaled eugenol and benzaldehyde
32.41 D.G. Moulton, A. Celebi, R.P. Fink: Olfaction in determined by electroencephalographic and be-
mammals two aspects: Proliferation of cells havioral olfactometry, Am. J. Vet. Res. 46, 2409
in the olfactory epithelium and sensitivity to 2412 (1985)
odours. In: Ciba Foundation Symposium on Taste 32.55 K.M. Dorries, E. Adkins-Regan, B.P. Halpern:
and Smell in Vertebrates, ed. by G.E.W. Wolsten- Olfactory sensitivity to the pheromone an-
holme, J. Knight (Churchill, London 1970) drostenone in sexually dimorphic in the pig,
32.42 D. Joshi, M. Vlkl, G.M. Shepherd, M. Laska: Olfac- Physiol. Behav. 57, 255259 (1995)
tory sensitivity for enantiomers and their racemic 32.56 S. Kowalewsky, M. Dambach, B. Mauck, G. Dehn-
mixtures A comparative study in CD-1 mice hardt: High olfactory sensitivity for dimethyl sul-
and spider monkeys, Chem. Senses 31, 655664 phide in harbour seal, Biol. Lett. 2, 106109
(2006) (2006)
32.43 M. Laska, O. Persson, L.T. Hernandez Salazar: Ol- 32.57 D.A. Marshall, L. Blumer, D.G. Moulton: Odor de-
factory sensitivity for alkylpyrazines A compar- tection curves for n-pentanoic acid in dogs and
ative study in CD-1 mice and spider monkeys, humans, Chem. Senses 6, 445453 (1981)
J. Exp. Zool. A 311, 278288 (2009) 32.58 M. Laska, P. Teubner: Odor structure-activity rela-
32.44 H. Walln, I. Engstrm, L.T. Hernandez Salazar, tionships of carboxylic acids correspond between

Part D | 32
M. Laska: Olfactory sensitivity for six amino acids: squirrel monkeys and humans, Am. J. Physiol.
A comparative study in CD-1 mice and spider 274, R1639R1645 (1998)
monkeys, Amino Acids 42, 14751485 (2012) 32.59 M. Laska, P. Teubner: Olfactory discrimination
32.45 M. Laska, M. Fendt, A. Wieser, T. Endres, L.T. Her- ability for homologous series of aliphatic alcohols
nandez Salazar, R. Apfelbach: Detecting danger and aldehydes, Chem. Senses 24, 263270 (1999)
or just another odorant? Olfactory sensitivity for 32.60 M. Laska, F. Hbener: Olfactory discrimination
the fox odor component 2,4,5-trimethyl thiazo- ability for homologous series of aliphatic ketones
line in four species of mammals, Physiol. Behav. and acetic esters, Behav. Brain Res. 119, 193201
84, 211215 (2005) (2001)
32.46 M. Laska, A. Wieser, L.T. Hernandez Salazar: Ol- 32.61 M. Laska, S. Trolp, P. Teubner: Odor structure-ac-
factory responsiveness to two odorous steroids tivity relationships compared in human and non-
in three species of nonhuman primates, Chem. human primates, Behav. Neurosci. 113, 981007
Senses 30, 505511 (2005) (1999)
32.47 M. Laska, D. Hfelmann, D. Huber, M. Schu- 32.62 M. Laska, D. Freyer: Olfactory discrimination abil-
macher: Does the frequency of occurrence of ity for aliphatic esters in squirrel monkeys and
odorants in the chemical environment deter- humans, Chem. Senses 22, 457465 (1997)
mine olfactory sensitivity? A study with acyclic 32.63 A. Rizvanovic, M. Amundin, M. Laska: Olfactory
monoterpene alcohols in three species of nonhu- discrimination ability of Asian elephants (Elephas
man primates, J. Chem. Ecol. 32, 13171331 (2006) maximus) for structurally related odorants, Chem.
32.48 M. Laska, R.M. Rivas Bautista, D. Hfelmann, Senses 38, 107118 (2013)
V. Sterlemann, L.T. Hernandez Salazar: Olfactory 32.64 J. Arvidsson, M. Amundin, M. Laska: Successful
sensitivity for putrefaction-associated thiols and acquisition of an olfactory discrimination test by
688 Part D Odorant Sensing and Physiological Effects

Asian elephants, Elephas maximus, Physiol. Be- 32.81 W.S. Cain: Differential sensitivity for smell: Noise
hav. 105, 809814 (2012) at the nose, Science 195, 796798 (1977)
32.65 M. Laska, E. Lord, S. Selin, M. Amundin: Olfac- 32.82 L. Jacquot, J. Hidalgo, G. Brand: Just notice-
tory discrimination of aliphatic odorants in South able difference in olfaction is related to trigemi-
African fur seals (Arctocephalus pusillus), J. Comp. nal component of odorants, Rhinol. 48, 281284
Psychol. 124, 187193 (2010) (2010)
32.66 M. Laska, . Rosandher, S. Hommen: Olfactory 32.83 B.M. Slotnick, J.E. Ptak: Olfactory intensity-dif-
discrimination of aliphatic odorants at 1 ppm: Too ference thresholds in rats and humans, Physiol.
easy for CD-1 mice to show odor structure-activity Behav. 19, 795802 (1977)
relationships?, J. Comp. Physiol. A 194, 971980 32.84 B.M. Slotnick: Olfactory discrimination in rats
(2008) with anterior amygdala leasions, Behav. Neurosci.
32.67 M. Laska, C.G. Galizia, M. Giurfa, R. Menzel: Ol- 99, 956963 (1985)
factory discrimination ability and odor structure- 32.85 P. Bonfils, P. Faulcon, L. Tavernier, N.A. Bonfils,
activity relationships in honeybees, Chem. Senses D. Malinvaud: Home accidents associated with
24, 429438 (1999) anosmia, Presse Med. 37, 742745 (2008)
32.68 M. Laska, S. Ayabe-Kanamura, F. Hbener, 32.86 D.V. Santos, E.R. Reiter, L.J. DiNardo,
S. Saito: Olfactory discrimination ability for R.M. Costanzo: Hazardous events associated
aliphatic odorants as a function of oxygen moiety, with impaired olfactory function, Arch. Otolaryn-
Chem. Senses 25, 189197 (2000) gol. Head Neck Surg. 130, 317319 (2004)
32.69 M. Pontz: Untersuchung des geruchlichen Un- 32.87 S. Doucet, R. Soussignan, P. Sagot, B. Schaal: The
terscheidungsvermgens von Totenkopfaffen smellscape of mothers breast: Effects of odor
(Saimiri sciureus) fr aliphatische Substanzen masking and selective unmasking on neonatal
mit unterschiedlichen funktionellen Gruppen, arousal, oral, and visual responses, Dev. Psy-
Ph.D. Thesis (University of Munich, Munich 2000), chobiol. 49, 129138 (2007)
German 32.88 J. Porter, B. Craven, R.M. Khan, S.J. Chang, I. Kang,
32.70 M. Laska: Olfactory discrimination ability of hu- B. Judkewicz, J. Volpe, G. Settles, N. Sobel: Mech-
man subjects for enantiomers with an iso- anisms of scent-tracking in humans, Nature Neu-
propenyl group at the chiral center, Chem. Senses rosci. 10, 2729 (2007)
29, 143152 (2004) 32.89 I. Croy, S. Negoias, L. Novakova, B.N. Landis,
32.71 M. Laska, P. Teubner: Olfactory discrimination T. Hummel: Learning about the functions of the
ability of human subjects for ten pairs of enan- olfactory system from people without a sense of
tiomers, Chem. Senses 24, 161170 (1999) smell, PLoS ONE 7, e33365 (2012)
32.72 M. Laska, A. Liesen, P. Teubner: Enantioselectivity 32.90 K. Aschenbrenner, C. Hummel, K. Teszmer,
of odor perception in squirrel monkeys and hu- F. Krone, T. Ishimaru, H.S. Seo, T. Hummel: The
mans, Am. J. Physiol. 277, R1098R1103 (1999) influence of olfactory loss on dietary behaviors,
32.73 M. Laska, D. Genzel, A. Wieser: The number of Laryngoscope 118, 135144 (2008)
olfactory receptor genes and the relative size of 32.91 T. Hummel, S. Nordin: Olfactory disorders and
olfactory brain structures are poor predictors of their consequences for quality of life, Acta. Oto-
olfactory discrimination performance with enan- Laryngol. 125, 116121 (2005)
tiomers, Chem. Senses 30, 171175 (2005) 32.92 C. Moessnang, A. Finkelmeyer, A. Vossen,
Part D | 32

32.74 S. Kim, M. Amundin, M. Laska: Olfactory discrim- F. Schneider, U. Habel: Assessing implicit odor
ination ability of South African fur seals (Arc- localization in humans using a cross-modal
tocephalus pusillus) for enantiomers, J. Comp. spatial cueing paradigm, PLoS ONE 6, e29614
Physiol. A 199, 535544 (2013) (2012)
32.75 M. Laska, G.M. Shepherd: Olfactory discrimination 32.93 J.D. Porteous: Smellscape, Progr. Phys. Geogr. 9,
ability of CD-1 mice for a large array of enan- 356378 (1985)
tiomers, Neurosci. 144, 295301 (2007) 32.94 M. Beaulieu-Lefebvre, F.C. Schneider, R. Kupers,
32.76 M. Laska, C.G. Galizia: Enantioselectivity of odor M. Ptito: Odor perception and odor awareness in
perception in honeybees, Behav. Neurosci. 115, congenital blindness, Brain Res. Bull. 84, 206
632639 (2001) 209 (2011)
32.77 T. Clarin, S. Sandhu, R. Apfelbach: Odor detection 32.95 R.L. Doty, M.M. Orndorff, J. Leyden, A. Kligman:
and odor discrimination in subadult and adult Communication of gender from human axillary
rats for two enantiomeric odorants supported by odors: Relationship to perceived intensity and
c-fos data, Behav. Brain Res. 206, 229235 (2010) hedonicity, Behav. Biol. 23, 373380 (1978)
32.78 B.D. Rubin, L.C. Katz: Spatial coding of enan- 32.96 S. Mitro, A.R. Gordon, M.J. Olsson, J.N. Lundstrm:
tiomers in the rat olfactory bulb, Nat. Neurosci. The smell of age: Perception and discrimination of
4, 355356 (2001) body odors of different ages, PLoS ONE 7, e38110
32.79 K.J. Rossiter: Structure-odor relationships, Chem. (2012)
Rev. 96, 32013240 (1996) 32.97 D. Singh, P.M. Bronstad: Female body odour is
32.80 W.S. Cain: Odor magnitude: Coarse versus fine a potential cue to ovulation, Proc. Roy. Soc. B 268,
grain, Percept. Psychophys. 22, 545549 (1977) 797801 (2001)
Human and Animal Olfactory Capabilities Compared References 689

32.98 S. Kuukasjrvi, C.J.P. Eriksson, E. Koskela, strains (Mus musculus) and major histocompat-
T. Mappes, K. Nissinen, M.J. Rantala: Attractive- ibility complex types by humans (Homo sapiens),
ness of womens body odors over the menstrual J. Comp. Psychol. 100, 262265 (1986)
cycle: The role of oral contraceptives and receiver 32.111 J. Havliek, R. Dvoakova, L. Bartos, J. Flegr:
sex, Behav. Ecol. 15, 579584 (2004) Non-advertized does not mean concealed: Body
32.99 C.L. Whittle, S. Fakharzadeh, J. Eades, G. Preti: odour changes across the human menstrual cycle,
Human breath odors and their use in diagnosis, Ethology 112, 8190 (2006)
Ann. N.Y. Acad. Sci. 1098, 252266 (2007) 32.112 C. Ober: Studies of HLA, fertility and mate choice
32.100 F. Prugnolle, T. Lefevre, F. Renaud, A.P. Mller, in a human isolate, Hum. Reprod. Update 5, 103
D. Miss, F. Thomas: Infection and body odours: 107 (1999)
evolutionary and medical perspectives, Infect. 32.113 R. Chaix, C. Cao, P. Donnelly: Is mate choice in hu-
Genet. Evol. 9, 10061009 (2009) mans MHC-dependent?, PLoS Genet. 4, e1000184
32.101 M. Schleidt: Personal odor and nonverbal com- (2008)
munication, Ethol. Sociobiol. 1, 225231 (1980) 32.114 J. Havliek, S.C. Roberts: MHC-correlated mate
32.102 R.H. Porter: Olfaction and human kin recognition, choice in humans: A review, Psychoneuroen-
Genetica 104, 259263 (1999) docrinology 34, 497512 (2009)
32.103 G.E. Weisfeld, T. Czilli, K.A. Phillips, J.A. Gall, 32.115 I. Croy, V. Bojanowski, T. Hummel: Men with-
C.M. Lichtman: Possible olfaction-ased mecha- out a sense of smell exhibit a strongly reduced
nisms in human kin recognition and inbreeding number of sexual relationships, women exhibit
avoidance, J. Exp. Child Psychol. 85, 279295 reduced partnership security A reanalysis of
(2003) previously published data, Biol. Psychol. 92, 292
32.104 R.H. Porter: Mutual mother-infant recognition in 294 (2013)
humans. In: Kin Recognition, ed. by P.G. Hepper 32.116 V. Gudziol, S. Wolff-Stephan, K. Aschenbrenner,
(Cambridge Univ. Press, Cambridge 1991) pp. 413 P. Joraschky, T. Hummel: Depression resulting
432 from olfactory dysfunction is associated with re-
32.105 J. Havliek, P. Lenochova: The effect of meat con- duced sexual appetite A cross-sectional cohort
sumption on body odor attractiveness, Chem. study, J. Sexual Med. 6, 19241929 (2009)
Senses 31, 747752 (2006) 32.117 J.M. Brunstrom: Dietary learning in humans: Di-
32.106 J. Albrecht, M. Demmel, V. Schopf, A.M. Klee- rections for future research, Physiol. Behav. 85,
man, R. Kopietz, J. May, T. Schreder, R. Zer- 5765 (2005)
necke, H. Bruckmann, M. Wiesmann: Smelling 32.118 J.A. Gottfried, J. ODoherty, R.J. Dolan: Appeti-
chemosensory signals of males in anxious versus tive and aversive olfactory learning in humans
nonanxious condition increases state anxiety of studied using event-related functional magnetic
female subjects, Chem. Senses 36, 1927 (2011) resonance imaging, J. Neurosci. 22, 1082910837
32.107 W. Zhou, D. Chen: Entangled chemosensory emo- (2002)
tion and identity: Familiarity enhances detection 32.119 D.A. Wilson, R.J. Stevenson: Learning to Smell. Ol-
of chemosensorily encoded emotion, Social Neu- factory Perception from Neurobiology to Behavior
rosci. 6, 270276 (2011) (Johns Hopkins Univ. Press, Baltimore 2006)
32.108 D.L. Wells, P.G. Hepper: The discrimination of dog 32.120 T. Engen, B.M. Ross: Long-term memory of odors
odours by humans, Percept. 29, 111115 (2000) with and without verbal descriptions, J. Exp. Psy-

Part D | 32
32.109 P.G. Hepper, D.L. Wells: Individually identifiable chol. 100, 221227 (1973)
body odors are produced by the gorilla and dis- 32.121 W.P. Goldman, J.G. Seamon: Very long-term
criminated by humans, Chem. Senses 35, 263268 memory for odors: Retention of odor-name as-
(2010) sociations, Am. J. Psychol. 105, 549563 (1992)
32.110 A.N. Gilbert, K. Yamazaki, G.K. Beauchamp,
L. Thomas: Olfactory discrimination of mouse

You might also like