Sunteți pe pagina 1din 8

Behavioral Neuroscience Copyright 2003 by the American Psychological Association, Inc.

2003, Vol. 117, No. 6, 1142–1149 0735-7044/03/$12.00 DOI: 10.1037/0735-7044.117.6.1142

Olfactory Sensitivity for Aliphatic Esters in Spider Monkeys


(Ateles geoffroyi)

Laura Teresa Hernandez Salazar Matthias Laska


Universidad Veracruzana University of Munich Medical School

Ernesto Rodriguez Luna


Universidad Veracruzana

Using a conditioning paradigm, the authors investigated the olfactory sensitivity of 3 spider monkeys
(Ateles geoffroyi) for a homologous series of aliphatic esters (ethyl acetate to n-octyl acetate) and
isomeric forms of some of these substances. With all odorants, the monkeys significantly discriminated
concentrations below 1 ppm from the odorless solvent, and in several cases, individual monkeys even
demonstrated thresholds below 1 ppb. The results showed spider monkeys to have a high olfactory
sensitivity for aliphatic esters, which for the majority of substances matches or even is better than that
of species such as the rat, the mouse, or the dog. These findings support the assumption that between-
species comparisons of neuroanatomical features are poor predictors of olfactory performance.

There is a long-standing tradition in olfactory research of as- to members of this order of mammals (Heymann, 1998; Kappeler,
signing animal species or even whole orders of animals as being 1998; Smith & Abbott, 1998).
“macrosmatic,” that is, having a keen sense of smell, or as being Laska and coworkers introduced a new behavioral test that, for
“microsmatic,” that is, having a poor sense of smell. This dichot- the first time, allowed the assessment of olfactory performance in
omy is mainly, if not exclusively, based on an interpretation of two species of nonhuman primates, the squirrel monkey (Saimiri
neuroanatomical features such as the relative size of olfactory sciureus) and the pigtail macaque (Macaca nemestrina), using
brain structures or the absolute size of olfactory epithelia (Farb- psychophysical methods (Hübener & Laska, 2001; Laska & Hud-
man, 1992; Stephan, Baron, & Frahm, 1988). However, physio- son, 1993a). Subsequent studies demonstrated that both species
logical evidence supporting a positive correlation between allo- possess highly developed olfactory discrimination abilities (Laska
metric measures of neuroanatomical structures and olfactory & Freyer, 1997; Laska & Hudson, 1993b, 1995; Laska, Liesen, &
performance is largely lacking (Brown, 2001; De Winter & Ox- Teubner, 1999; Laska & Teubner, 1998; Laska, Trolp, & Teubner,
nard, 2001; Schoenemann, 2001).
1999), as well as an excellent long-term memory for odors
Primates, for example, are invariably regarded as microsmatic
(Hübener & Laska, 1998; Laska, Alicke, & Hudson, 1996). Fur-
animals (e.g., King & Fobes, 1974; Walker & Jennings, 1991), at
ther, these studies showed that Saimiri sciureus and Macaca nem-
least partly as a result of the fact that the massive increase in their
estrina have a well-developed olfactory sensitivity for aliphatic
neocortex volume during evolution led to a decrease in the relative
carboxylic acids (Laska, Seibt, & Weber, 2000), esters (Laska &
size (although not necessarily the absolute size) of their olfactory
bulbs, that is, the first neuropil of the olfactory pathway (Barton, Seibt, 2002b), alcohols (Laska & Seibt, 2002b), and aldehydes
1996). In recent years, however, an increasing number of behav- (Laska, Hofmann, & Simon, 2003) that is not generally inferior to,
ioral observations call into question the still widely held belief that and in some cases is even better than, that of species traditionally
primates have generally only poor olfactory capabilities and that regarded as being macrosmatic.
this sensory modality is of only little, if any, behavioral relevance In order to get an indication of whether squirrel monkeys and
pigtail macaques are special in their olfactory performance or
whether a more general reevaluation of the efficiency of the sense
Laura Teresa Hernandez Salazar and Ernesto Rodriguez Luna, Instituto of smell in simian primates should be considered, we believe that
de Neuro-Etologia, Universidad Veracruzana, Xalapa, Mexico; Matthias it is worthwhile to conduct corresponding investigations of olfac-
Laska, Department of Medical Psychology, University of Munich Medical tory capacity using other primate species. This should also allow
School, Munich, Germany. us to further assess the suitability of allometric comparisons of
Financial support provided by the Volkswagen Foundation (I/77 391) is neuroanatomical features to predict olfactory performance, and
gratefully acknowledged. Laura Teresa Hernandez Salazar was awarded
thus the validity of the concept of microsmatic and macrosmatic
Grant 124786 by Conacyt Mexico.
Correspondence concerning this article should be addressed to Matthias
species. A comparison of olfactory capabilities between different
Laska, Department of Medical Psychology, University of Munich Medical primate species might also help to reveal evolutionary adaptations
School, Goethestrasse 31, D-80336 Munich, Germany. E-mail: of this sensory system among this order of mammals (Barton,
laska@imp.med.uni-muenchen.de Purvis, & Harvey, 1995).

1142
OLFACTORY SENSITIVITY IN SPIDER MONKEYS 1143

Recently, the authors of the present study successfully intro- increasing dilutions of an odorant used as S⫹, and the alternative paper
duced the behavioral test used with squirrel monkeys and pigtail strip scented with the odorless solvent alone used as S⫺. Starting with a
macaques to another primate species, the spider monkey (Ateles dilution of 1:100, each odorant was successively presented in 10-fold
geoffroyi; Laska, Hernandez Salazar, & Rodriguez Luna, 2003). dilution steps for three sessions each until a monkey failed to significantly
discriminate the odorant from the solvent. Subsequently, this descending
It was therefore the aim of the present study to assess the
staircase procedure was repeated for three more sessions per dilution step.
olfactory sensitivity of Ateles geoffroyi by determining olfactory
Finally, intermediate dilutions were tested in order to determine the thresh-
detection thresholds for an array of monomolecular odorants. We old value more exactly. If, for example, a monkey significantly discrimi-
chose aliphatic esters as odor stimuli because this class of sub- nated a 1:10,000 dilution from the solvent but failed to do so with a
stances comprises the quantitatively and qualitatively dominant 1:100,000 dilution, then the monkey was presented with a 1:30,000 dilu-
components in a wide variety of fruit odors (Nursten, 1970; tion. To prevent the more challenging conditions leading to extinction or to
Rouseff & Leahy, 1995) and thus is presumed to be behaviorally a decline in the monkey’s motivation, these were always followed by a
relevant for these primates, and because comparative data from return to an easy control task. This consisted of the discrimination between
humans and, at least for the majority of odorants, from other a 100-fold dilution of the S⫹ and the odorless solvent as S⫺.
mammals including two species of nonhuman primates are at hand.
Odorants
Method A set of 10 odorants was used: ethyl acetate, n-propyl acetate, n-butyl
acetate, n-pentyl acetate, n-hexyl acetate, n-heptyl acetate, n-octyl acetate,
Subjects isopropyl acetate, isobutyl acetate, and isopentyl acetate. The rationale for
Testing was performed with 3 adult female spider monkeys (Ateles choosing these substances was to assess the monkeys’ sensitivity for
geoffroyi) maintained as part of a breeding colony at the Parque de la Flora odorants representing members of a homologous series of aliphatic com-
y Fauna Silvestre Tropical, Catemaco, Veracruz, Mexico. All monkeys had pounds, that is, substances sharing the same functional group but differing
served as subjects in previous olfactory experiments (Laska, Hernandez in carbon chain length, as well as isomeric forms of some of these
Salazar, & Rodriguez Luna, 2003) and were completely familiar with the compounds, that is, substances sharing the same sum formula and func-
basic test procedure. The group was housed under seminatural conditions tional group but differing in the branching of their carbon chains, allowing
in an enclosure of 72 m3 with three adjacent single cages that could be us to assess the impact of both structural features on detectability. All
closed by sliding doors, thus allowing the temporary separation of monkeys substances were obtained from Merck (Darmstadt, Germany) and had a
for individual testing. The monkeys were trained to enter their test cage nominal purity of at least 99%. They were diluted with odorless diethyl
voluntarily and remained in visual and auditory contact with the rest of phthalate (Merck) as the solvent.
their social group during testing. Monkeys were fed fresh fruit and vege- The order of presentation of odorants did not follow the sequence
tables, with ad-lib access to water. mentioned above but was pseudorandomized between monkeys.
The experiments reported here comply with the Guide for the Care and
Use of Laboratory Animals (National Institutes of Health Publication No. Data Analysis
86 –23, revised 1985) and with current German and Mexican laws.
In the method described here, the monkey had two options: to correctly
open the container that contained the positive stimulus (hit), or to incor-
Behavioral Test rectly open the container that contained the negative stimulus (false alarm).
The spider monkeys were tested in a two-choice instrumental condition- For each individual spider monkey, the percentage of hits from the best
ing paradigm (Hübener & Laska, 2001; Laska, Hernandez Salazar, & three consecutive sessions per dilution step, comprising a total of 30
Rodriguez Luna, 2003). Two cube-shaped open PVC containers (side decisions, was calculated and taken as the measure of performance.
length, 5.5 cm) were attached to a metal bar (50 cm long and 6 cm wide) Significance levels were determined by calculating binomial z-scores
at a distance of 22 cm. Each container was equipped with a hinged metallic corrected for continuity (Siegel & Castellan, 1988) from the number of
lid, hanging 2 cm down the front of the container. From the center of the correct and false responses for each individual monkey and condition. All
front part of the lid, a pin of 3 cm length extended toward the monkey and tests were two-tailed, and the alpha level was set at 0.05.
served as a lever to open the lid. A clip on top of each lid held an absorbent Correlations between olfactory threshold values and carbon chain length
paper strip (70 ⫻ 10 mm) impregnated with 10 ␮l of an odorant, signaling of the substances tested were calculated by using both the Spearman
either that the container held a Kellogg’s Honey Loop food reward (S⫹) or rank-correlation test and second order polynomial regression analysis.
that it did not (S⫺). The odorized paper strips extended 5 cm into the test
cage when the apparatus was attached to the front of the cage. Results
When a monkey tried to open a container without prior sniffing, the
experimenter held a chain connected to the lid tight so that the monkey Figure 1 shows the performance of the spider monkeys in
could not move the lid. After the olfactory investigation had taken place, discriminating between various dilutions of a given odorant and
the chain was loosened and the monkey could open the container. the odorless solvent. All 3 monkeys significantly distinguished
In each test trial, each monkey sniffed at both options for as often as it dilutions as low as 1:1 million ethyl acetate, 1:100,000 n-propyl
liked and then decided to open one of the two boxes. After each decision, acetate, 1:30 million n-butyl acetate, 1:30 million n-pentyl acetate,
the experimenter removed the apparatus from the mesh and— out of sight 1:1 million n-hexyl acetate, 1:300,000 n-heptyl acetate, 1:100,000
of the test monkey—prepared it for the next trial by again baiting the
n-octyl acetate, 1:10 million isopropyl acetate, 1:30,000 isobutyl
container bearing the S⫹. A pseudorandomized sequence of presentations
of the S⫹ on the left or on the right side was adopted. Ten such trials were
acetate, and 1:30 million isopentyl acetate from the solvent (bino-
conducted per monkey and session, and, usually, three sessions were mial test, p ⬍ .05), with single individuals even scoring better.
conducted per day. The individual spider monkeys demonstrated very similar
Olfactory detection thresholds were determined by testing the monkey’s threshold values, and with 8 of the 10 odorants, they differed only
ability to discriminate between an absorbent paper strip scented with by a dilution factor of 3 or 10 between the highest- and the
OLFACTORY SENSITIVITY IN SPIDER MONKEYS 1145

lowest-scoring monkey. The largest difference in sensitivity for a


given odorant between individual monkeys represented a dilution
factor of 33 and was found with n-hexyl acetate and n-heptyl
acetate.
Figure 2 shows the olfactory detection threshold values (ex-
pressed as vapor phase concentrations) of the spider monkeys as a
function of carbon chain length of the aliphatic esters tested. When
linear correlational statistics were applied to the data, no signifi-
cant correlation between perceptibility in terms of olfactory detec-
tion thresholds and carbon chain length of the n-acetic esters was
found (Spearman, rs ⫽ ⫺.37, p ⬎ .05). However, when the
threshold values for the two substances with the longest carbon
chains tested, that is, n-heptyl acetate and n-octyl acetate, were
removed from the calculations, a statistically significant negative
correlation was found (Spearman, rs ⫽ ⫺.63, p ⬍ .02). Similarly,
when the threshold values for the two substances with the shortest
carbon chains tested, that is, ethyl acetate and n-propyl acetate,
Figure 2. Olfactory detection threshold values (expressed as vapor phase
were removed from the calculations, a statistically significant concentrations) of the 3 spider monkeys (Nanny, circles; Suki, squares,
positive correlation was found (Spearman, rs ⫽ .67, p ⬍ .02). Piolina, triangles) as a function of carbon chain length of the aliphatic
Thus, the correlation between olfactory detection thresholds of the esters tested. ppm ⫽ parts per million.
spider monkeys and carbon chain length of the n-aliphatic esters
tested can best be described as a U-shaped function (second-order
polynomial regression, r ⫽ .75, p ⬍ .01). Although only 3 monkeys were tested, the results appear robust,
With the three isoforms of the acetic esters tested, linear corre- as interindividual variability was remarkably low and generally
lational analysis also failed to find a statistically significant cor- smaller than the range reported in studies on human olfactory
relation between threshold values and carbon chain length (Spear- sensitivity, that is, within three orders of magnitude (Stevens,
man, rs ⫽ ⫺.29, p ⬎ .05). However, when applying second-order Cain, & Burke, 1988). In fact, for the majority of substances tested,
polynomial regression analysis to the data, the correlation was there was a factor of only 3 or 10 between the threshold values of
found to follow an inverted U-shaped function (r ⫽ .97, p ⬍ .01). the highest and the lowest scoring monkey. Further, with all
A comparison of detection threshold values for n- and isoforms substances tested, the monkeys’ performance with the lowest con-
of acetic esters sharing the same number of carbons shows the centrations presented dropped to chance level, suggesting that the
following: Whereas the isoform of propyl acetate is detected at statistically significant discrimination between higher concentra-
markedly lower concentrations compared with the n-form by all 3 tions of an odorant and the odorless diluent was indeed based on
monkeys, the reverse is true for the two butyl acetates tested. The olfactory perception and not on other cues.
two isomers of pentyl acetate were detected at similar concentra- Figure 3 compares the olfactory detection threshold values
tions (cf. Figure 2). obtained with the spider monkeys for the substances tested to those
Table 1 summarizes the threshold dilutions for both the best and from other mammalian species. Although such across-species
the poorest performing spider monkeys, and shows various mea- comparisons should be considered with caution, as different meth-
sures of corresponding vapor phase concentrations (Weast, 1987) ods may lead to widely differing results—as can be seen with the
allowing readers to easily compare the data obtained in the present threshold values depicted for n-butyl acetate, n-pentyl acetate, and
study to those reported by other authors using one of these con- n-hexyl acetate in the rat—it seems admissible to state that Ateles
vertible measures. All threshold dilutions correspond to vapor geoffroyi is far from being considered a microsmat, that is, a
phase concentrations below 1 ppm, and in several cases, the species with a poorly developed sense of smell. With all n-acetic
monkeys were even able to detect concentrations below 1 ppb. esters, the spider monkeys demonstrated olfactory threshold values
that are at least two orders of magnitude lower than those of the rat
(Moulton, 1960) and the dog (Passe & Walker, 1985), both of
Discussion
which are traditionally regarded as macrosmatic animals, that is,
The results of this study demonstrate, for the first time, that species with a highly developed sense of smell. Similarly, the
spider monkeys have a high olfactory sensitivity for monomolec- sensitivity of the spider monkey for n-pentyl acetate was found to
ular odorants belonging to the class of aliphatic esters. Further, be markedly higher than that of the rabbit (Passe & Walker, 1985),
they show a nonlinear correlation between perceptibility in terms and for ethyl acetate to be as good as that of the mouse (Bodyak
of olfactory detection thresholds and carbon chain length of the & Slotnick, 1999), two other species usually considered to be
substances tested. macrosmats. It should be mentioned that all animal data shown in

Figure 1 (opposite). Performance of 3 spider monkeys in discriminating between various dilutions of a given
odorant and the odorless solvent. Each data point represents the percentage of correct choices from 30 decisions.
Filled symbols indicate dilutions that were not discriminated above chance level (binomial test, p ⬎ .05).
1146 HERNANDEZ SALAZAR, LASKA, AND RODRIGUEZ LUNA

Table 1
Olfactory Detection Threshold Values in Ateles geoffroyi Expressed in Various Measures of
Vapor Phase Concentrations

Odorant Dilution Molec./cm3 ppm log ppm Mol/L log mol/L


⫺9
ethyl acetate 1:1 million 3.0 ⫻ 10 12
0.11 ⫺0.96 5.0 ⫻ 10 ⫺8.30
1:3 million 1.0 ⫻ 1012 0.036 ⫺1.43 1.7 ⫻ 10⫺9 ⫺8.78
n-propyl acetate 1:100,000 1.2 ⫻ 1013 0.45 ⫺0.34 2.0 ⫻ 10⫺8 ⫺7.69
1:1 million 1.2 ⫻ 1012 0.045 ⫺1.35 2.0 ⫻ 10⫺9 ⫺8.69
n-butyl acetate 1:30 million 1.6 ⫻ 1010 0.0006 ⫺3.22 2.7 ⫻ 10⫺11 ⫺10.57
1:300 million 1.6 ⫻ 109 0.00006 ⫺4.22 2.7 ⫻ 10⫺12 ⫺11.57
n-pentyl acetate 1:30 million 7.3 ⫻ 109 0.00027 ⫺3.57 1.2 ⫻ 10⫺11 ⫺10.92
1:300 million 7.3 ⫻ 108 0.000027 ⫺4.57 1.2 ⫻ 10⫺12 ⫺11.92
n-hexyl acetate 1:1 million 1.1 ⫻ 1011 0.0039 ⫺2.40 1.8 ⫻ 10⫺10 ⫺9.75
1:30 million 3.5 ⫻ 109 0.00013 ⫺3.88 5.9 ⫻ 10⫺12 ⫺11.23
n-heptyl acetate 1:300,000 1.6 ⫻ 1011 0.0059 ⫺2.23 2.7 ⫻ 10⫺10 ⫺9.58
1:10 million 4.8 ⫻ 109 0.00018 ⫺3.75 7.9 ⫻ 10⫺12 ⫺11.10
n-octyl acetate 1:100,000 2.7 ⫻ 1011 0.0099 ⫺2.01 4.4 ⫻ 10⫺10 ⫺9.35
1:300,000 8.9 ⫻ 1010 0.0033 ⫺2.48 1.5 ⫻ 10⫺10 ⫺9.83
iso-propyl acetate 1:10 million 1.9 ⫻ 1011 0.0072 ⫺2.15 3.2 ⫻ 10⫺10 ⫺9.49
1:30 million 6.4 ⫻ 1010 0.0024 ⫺2.63 1.1 ⫻ 10⫺10 ⫺9.97
iso-butyl acetate 1:30,000 2.3 ⫻ 1013 0.85 ⫺0.07 3.9 ⫻ 10⫺8 ⫺7.41
1:100,000 7.0 ⫻ 1012 0.26 ⫺0.58 1.2 ⫻ 10⫺8 ⫺7.93
iso-pentyl acetate 1:30 million 9.8 ⫻ 109 0.00036 ⫺3.44 1.6 ⫻ 10⫺11 ⫺10.79
1:100 million 2.9 ⫻ 109 0.00012 ⫺3.96 4.9 ⫻ 10⫺12 ⫺11.31

Note. With each stimulus, the upper line represents the lowest concentration that all 3 monkeys were able to
detect, and the lower line represents the lowest concentration that the best performing monkey was able to detect.
molec. ⫽ molecules; ppm ⫽ parts per million.

Figure 3 are taken from studies that used instrumental conditioning tory performance should be discussed with regard to possible
paradigms (and not spontaneous preference or habituation/disha- explanations for differences in sensitivity between substances.
bituation paradigms), and it is commonly agreed that such methods Our finding of a high olfactory sensitivity for aliphatic esters in
provide the best approximation of an animal’s sensory capabilities spider monkeys is yet another example showing that allometric
(Hastings, 2003). comparisons of olfactory brain structure volumes or of the absolute
A comparison of the spider monkeys’ performance with that of size of olfactory epithelia are poor predictors of chemosensory
two other species of nonhuman primates tested in an earlier study performance. There is no doubt that the relative size of the rat’s or
(Laska & Seibt, 2002b) using exactly the same (pigtail macaques) the dog’s brain structures devoted to processing olfactory infor-
or a very similar (squirrel monkeys) method as that of the present mation and the total area of the dog’s olfactory epithelium (and
study reveals that, with the exception of n-heptyl acetate, spider concomitantly its total number of olfactory receptors) are both
monkeys and squirrel monkeys display very similar or even iden- considerably larger than those of the spider monkey or of other
tical olfactory detection thresholds for acetic esters, and that, with human or nonhuman primates (Farbman, 1992; Stephan et al.,
the exception of isobutyl acetate, both New World primate species 1988). The present data as well as those from other studies that
perform better than the pigtail macaque, an Old World primate reported olfactory detection threshold values for aliphatic alcohols,
species. All three species of nonhuman primates generally show aldehydes, and carboxylic acids in squirrel monkeys and pigtail
lower threshold values compared with human subjects, which macaques (Laska et al., 2000; Laska, Hofmann, & Simon, 2003;
nevertheless appear to be more sensitive for most of the n-acetic Laska & Seibt, 2002a), however, clearly show that such compar-
esters tested than the rat (cf. Figure 5). isons of neuroanatomical structures do not allow generalizable
It should be mentioned that the threshold values of the human conclusions as to olfactory sensitivity of any two species.
subjects as depicted in Figure 3 are taken from the study by Considering that even for the most intensively studied species of
Cometto-Muniz and Cain (1991). Although some other studies nonhuman mammals, measurements of olfactory sensitivity or
reported slightly lower values for some of the substances (e.g., for discrimination abilities have so far usually been restricted to little
ethyl acetate [Hellman & Small, 1974] and for n-butyl acetate more than a handful of substances (Walker & Jennings, 1991), it is
[Dravnieks & Laffort, 1972]), these other studies had tested only obvious that the assignment of general labels such as microsmat or
one or a few members of the homologous series of esters, and none macrosmat to any species is at least premature and does not take
of them had used signal detection methods and a comparably into account the vast complexity of our natural odor world (Laing,
sophisticated mode of stimulus presentation as had Cometto- Cain, McBride, & Ache, 1989) and the diversity of contexts in
Muniz and Cain. which the sense of smell may be crucial for an animal (Doty,
Across-species comparisons of olfactory performance raise the 1986). Therefore, we argue that these terms should no longer be
question as to possible reasons for the observed similarities and— used.
sometimes marked— differences in olfactory sensitivity for a In order to explain similarities or differences in olfactory per-
given substance. Likewise, within-species comparisons of olfac- formance between or within species, it might be more promising to
OLFACTORY SENSITIVITY IN SPIDER MONKEYS 1147

Figure 3. Comparison of the olfactory detection threshold values (expressed as vapor phase concentrations) of
the spider monkeys for aliphatic esters with those of other mammalian species (human data: Cometto-Muniz &
Cain, 1991; animal data: Bodyak & Slotnick, 1999; Laska, 1990; Laska & Seibt, 2002b; Moulton, 1960; Passe
& Walker, 1985). ppm ⫽ parts per million.

consider whether given odorants or classes of odorants might believed to be highly relevant for species feeding on animal prey
differ in their degree of behavioral relevance for a species. but presumably less important for mainly frugivorous primates.
Spider monkeys, as well as squirrel monkeys and pigtail ma- Despite the obvious role that the sense of smell plays in finding
caques, have been reported to include a considerable proportion of and selecting food in many species, it should be emphasized that
fruits in their diets, with the two first-mentioned species usually dietary specialization is only one of presumably numerous factors
scoring higher values regarding this dietary specialization com- that make up the ecological niche of a species and that are likely
pared with the last-mentioned species (Clutton-Brock & Harvey, to affect its pattern of olfactory sensitivity and discrimination
1977; Ross, 1992). Our finding that both spider monkeys and ability (Barton et al., 1995). Identifying such factors and their
squirrel monkeys are generally—and in some cases even mark- impact on measures of olfactory performance warrants further
edly—more sensitive to aliphatic esters than rats or dogs appears studies that include as many animal species and odor stimuli as
to make sense in terms of an evolutionary adaptation to optimal possible.
foraging (Stephens & Krebs, 1986), as these substances are known
A final aspect of the present study is our finding of a nonlinear
to be major components of a wide variety of fruit odors (Nursten,
(i.e., U-shaped) correlation between olfactory detection threshold
1970; Rouseff & Leahy, 1995) and thus are likely to be more
values obtained in the spider monkeys and carbon chain length of
relevant for species feeding on fruit than for a carnivorous species
the n-aliphatic esters tested. Although squirrel monkeys and pigtail
such as the dog or a granivorous species such as the rat (Chivers
macaques have been reported to show a significant linear negative
& Hladik, 1980). This idea is also supported by the finding that the
short-tailed fruit bat, Carollia perspicillata, a frugivorous species correlation between these two measures (Laska & Seibt, 2002b),
also belonging to an order of mammals traditionally considered as second order polynomial regression analysis, which was not per-
microsmatic, has been reported to show olfactory detection thresh- formed in the original study, reveals that in these two nonhuman
olds for aliphatic esters that are markedly lower than those of rats primate species, too, the distribution of sensitivity across the
and at least as low as those of dogs (Laska, 1990; cf. Figure 3). homologous series of esters can best be described by a U-shaped
Carnivorous, insectivorous, or sanguivorous species such as the function. This finding lends further support to the notion that
dog, the hedgehog, or the vampire bat, in turn, have been found to olfactory sensitivity for structurally related substances is not a
be more sensitive to short-chained carboxylic acids compared with simple function of vapor pressure, as the latter decreases mono-
squirrel monkeys or pigtail macaques (Hübener & Laska, 2001; tonically with increasing carbon chain length of the esters tested.
Laska et al., 2000). This class of odorants comprises the main Interestingly, with the three isoforms of the acetic esters tested,
components of body-borne prey odors (Flood, 1985) and thus is the squirrel monkeys, but not the pigtail macaques, showed the
1148 HERNANDEZ SALAZAR, LASKA, AND RODRIGUEZ LUNA

same inverted U-shaped correlation between sensitivity and mo- Bodyak, N., & Slotnick, B. M. (1999). Performance of mice in an auto-
lecular length as the spider monkeys (Laska & Seibt, 2002b). mated olfactometer: Odor detection, discrimination and odor memory.
Both linear and U-shaped correlations between sensitivity and Chemical Senses, 24, 637– 645.
carbon chain length of acetic esters have also been found in Brown, W. M. (2001). Natural selection of mammalian brain components.
humans (Cometto-Muniz & Cain, 1991) and in rats (Moulton, Trends in Ecology and Evolution, 16, 471– 473.
Chivers, D. J., & Hladik, C. M. (1980). Morphology of the gastrointestinal
1960). For other homologous series of aliphatic substances such as
tract in primates: Comparisons with other mammals in relation to diet.
alcohols, aldehydes, or carboxylic acids, corresponding correla- Journal of Morphology, 166, 337–386.
tions have been reported both in squirrel monkeys and in humans Clutton-Brock, T. H., & Harvey, P. H. (1977). Species differences in
(Cometto-Muniz, Cain, & Abraham, 1998; Laska et al., 2000; feeding and ranging behaviour in primates. In T. H. Clutton-Brock (Ed.),
Laska & Seibt, 2002a; Laska, Hofmann, & Simon, 2003). This Primate ecology: Studies of feeding and ranging behaviour in lemurs,
suggests that regular connections between perceptibility and this monkeys and apes (pp. 557–584). New York: Academic Press.
molecular structural feature might not be restricted to the class of Cometto-Muniz, J. E., & Cain, W. S. (1991). Nasal pungency, odor, and
odorants tested here but instead may represent a more general eye irritation thresholds for homologous acetates. Pharmacology Bio-
phenomenon. chemistry and Behavior, 39, 983–989.
This may not be surprising, considering that carbon chain length Cometto-Muniz, J. E., Cain, W. S., & Abraham, M. H. (1998). Nasal
of odorant molecules has been shown to be an important determi- pungency and odor of homologous aldehydes and carboxylic acids.
Experimental Brain Research, 118, 180 –188.
nant of the specificity of interaction between stimulus and receptor
De Winter, W., & Oxnard, C. E. (2001, February 8). Evolutionary radia-
(Kaluza & Breer, 2000) as well as of the chemotopic organization
tions and convergences in the structural organization of mammalian
and thus of odor quality coding within the olfactory bulb (Johnson brains. Nature, 409, 710 –714.
& Leon, 2000). However, with regard to differences in sensitivity Doty, R. L. (1986). Odor-guided behavior in mammals. Experientia, 42,
for members of a homologous series of substances at the organis- 257–271.
mal level, it should be considered that the quantitative distribution Dravnieks, A., & Laffort, P. (1972). Physico-chemical basis of quantitative
of individual receptor types, each responding selectively to a and qualitative odor discrimination in humans. In D. Schneider (Ed.),
limited range of carbon chain lengths and functional groups, may Olfaction and taste IV (pp. 142–148). Stuttgart, Germany: Wiss. Ver-
differ between species. This may explain why one species may lagsgesellschaft.
show a regular connection between sensitivity and carbon chain Farbman, A. I. (1992). Cell biology of olfaction. Cambridge, UK: Cam-
length of a given class of substances, whereas another species does bridge University Press.
not or displays a different type of correlation. Flood, P. (1985). Sources of significant smells: The skin and other organs.
In R. E. Brown, & D. W. MacDonald (Eds.), Social odours in mammals
A heightened expression of a given receptor type with a specific
(pp. 19 –36). Oxford, UK: Clarendon Press.
molecular receptive range may also explain phenomena such as the
Hastings, L. (2003). Psychophysical evaluation of olfaction in nonhuman
markedly higher sensitivity of the spider monkey for n-butyl mammals. In R. L. Doty (Ed.), Handbook of olfaction and gustation (2nd
acetate compared with its neighbor in the homologous series, ed., pp. 385– 401). New York: Marcel Dekker.
n-propyl acetate (cf. Figure 2). A possible reason underlying such Hellman, T. M., & Small, F. H. (1974). Characterisation of the odor
phenomena is that repeated exposure to a given odorant—perhaps properties of 101 petrochemicals using sensory methods. Journal of the
due to its high abundance in the odorous environment of or its Air Pollution Control Association, 24, 979 –982.
biological significance for an animal—may induce an increased Heymann, E. W. (1998). Sex differences in olfactory communication in a
expression of a corresponding receptor type (Yee & Wysocki, primate, the moustached tamarin, Saguinus mystax (Callitrichinae). Be-
2001). havioral Ecology and Sociobiology, 43, 37– 45.
In conclusion, the results of the present study provide first Hübener, F., & Laska, M. (1998). Assessing olfactory performance in an
evidence of a well-developed olfactory sensitivity in the spider Old World primate, Macaca nemestrina. Physiology & Behavior, 64,
521–527.
monkey. This finding supports the assumption that olfaction may
Hübener, F., & Laska, M. (2001). A two-choice discrimination method to
play an important role in the regulation of behavior in this species.
assess olfactory performance in pigtailed macaques, Macaca nemest-
Further, the results lend additional support to the suggestions that rina. Physiology & Behavior, 72, 511–519.
between-species comparisons of neuroanatomical features are a Johnson, B. A., & Leon, M. (2000). Odorant molecular length: One aspect
poor predictor of olfactory performance and that general labels of the olfactory code. Journal of Comparative Neurology, 426, 330 –338.
such as microsmat or macrosmat—which usually are based on Kaluza, J. F., & Breer, H. (2000). Responsiveness of olfactory neurons to
allometric comparisons of olfactory brain structures—are inade- distinct aliphatic aldehydes. Journal of Experimental Biology, 203, 927–
quate to describe a species’ olfactory capabilities. 933.
An ecological view of such capabilities that attempts to correlate Kappeler, P. (1998). To whom it may concern: The transmission and
sensory performance with behavioral relevance of odor stimuli function of chemical signals in Lemur catta. Behavioral Ecology and
might offer a promising approach in appraising the significance of Sociobiology, 42, 411– 421.
King, J. E., & Fobes, J. L. (1974). Evolutionary changes in primate sensory
the sense of smell for a particular species.
capacities. Journal of Human Evolution, 3, 435– 443.
Laing, D. G., Cain, W. S., McBride, R. L., & Ache, B. W. (1989).
References Perception of complex smells and tastes. Sydney, Australia: Academic
Barton, R. A. (1996). Neocortex size and behavioural ecology in primates. Press.
Proceedings: Biological Sciences, 263, 173–177. Laska, M. (1990). Olfactory sensitivity to food odor components in the
Barton, R. A., Purvis, A., & Harvey, P. H. (1995). Evolutionary radiation short-tailed fruit bat, Carollia perspicillata. Journal of Comparative
of visual and olfactory brain systems in primates, bats and insectivores. Physiology, A: Sensory, Neural, and Behavioral Physiology, 166, 395–
Philosophical Transactions: Biological Sciences, 348, 381–392. 399.
OLFACTORY SENSITIVITY IN SPIDER MONKEYS 1149

Laska, M., Alicke, T., & Hudson, R. (1996). A study of long-term odor Moulton, D. G. (1960). Studies in olfactory acuity: III. Relative detect-
memory in squirrel monkeys, Saimiri sciureus. Journal of Comparative ability of n-aliphatic acetates by the rat. Quarterly Journal of Experi-
Psychology, 110, 125–130. mental Psychology, 12, 203–213.
Laska, M., & Freyer, D. (1997). Olfactory discrimination ability for ali- Nursten, H. E. (1970). Volatile compounds: The aroma of fruits. In A. C.
phatic esters in squirrel monkeys and humans. Chemical Senses, 22, Hulme (Ed.), The biochemistry of fruits and their products (Vol. 1, pp.
457– 465. 239 –268). London: Academic Press.
Laska, M., Hernandez Salazar, L. T., & Rodriguez Luna, E. (2003). Passe, D. H., & Walker, J. C. (1985). Odor psychophysics in vertebrates.
Successful acquisition of an olfactory discrimination paradigm by spider Neuroscience and Biobehavioral Reviews, 9, 431– 467.
monkeys (Ateles geoffroyi). Physiology & Behavior, 78, 321–329. Ross, C. (1992). Basal metabolic rate, body weight and diet in primates: An
Laska, M., Hofmann, M., & Simon, Y. (2003). Olfactory sensitivity for evaluation of the evidence. Folia Primatologica, 58, 7–23.
aliphatic aldehydes in squirrel monkeys and pigtail macaques. Journal of Rouseff, R. L., & Leahy, M. L. (1995). Fruit flavors: Biogenesis, charac-
Comparative Physiology, A: Sensory, Neural, and Behavioral Physiol- terization, and authentication. Washington, DC: American Chemical
ogy, 189, 263–271. Society.
Laska, M., & Hudson, R. (1993a). Assessing olfactory performance in a Schoenemann, P. T. (2001). Brain scaling, behavioral ability and human
New World primate, Saimiri sciureus. Physiology & Behavior, 53, evolution. Behavioral and Brain Sciences, 24, 293–295.
89 –95. Siegel, S., & Castellan, N. J. (1988). Nonparametric statistics for the
Laska, M., & Hudson, R. (1993b). Discriminating parts from the whole: behavioral sciences. New York: McGraw Hill.
Determinants of odor mixture perception in squirrel monkeys, Saimiri Smith, T. E., & Abbott, D. H. (1998). Behavioral discrimination between
sciureus. Journal of Comparative Physiology, A: Sensory, Neural, and circumgenital odor from peri-ovulatory dominant and anovolatory fe-
Behavioral Physiology, 173, 249 –256. male common marmosets (Callithrix jacchus). American Journal of
Laska, M., & Hudson, R. (1995). Ability of female squirrel monkeys Primatology, 46, 265–284.
(Saimiri sciureus) to discriminate between conspecific urine odours. Stephan, H., Baron, G., & Frahm, H. D. (1988). Comparative size of brains
Ethology, 99, 39 –52. and brain structures. In H. Steklis & J. Erwin (Eds.), Comparative
Laska, M., Liesen, A., & Teubner, P. (1999). Enantioselectivity of odor primate biology (Vol. 4, pp. 1–38). New York: Alan R. Liss.
perception in squirrel monkeys and humans. American Journal of Phys- Stephens, D. W., & Krebs, J. R. (1986). Foraging theory. Princeton, NJ:
iology, 277, R1098 –R1103. Princeton University Press.
Laska, M., & Seibt, A. (2002a). Olfactory sensitivity for aliphatic alcohols Stevens, J. C., Cain, W. S., & Burke, R. J. (1988). Variability of olfactory
in squirrel monkeys and pigtail macaques. Journal of Experimental thresholds. Chemical Senses, 13, 643– 653.
Biology, 205, 1633–1643. Walker, J. C., & Jennings, R. A. (1991) Comparison of odor perception in
Laska, M., & Seibt, A. (2002b). Olfactory sensitivity for aliphatic esters in humans and animals. In D. G. Laing, R. L. Doty, & W. Breipohl (Eds.),
squirrel monkeys and pigtail macaques. Behavioural Brain Research, The human sense of smell (pp. 261–280). Berlin: Springer.
134, 165–174. Weast, R. C. (1987). Handbook of chemistry and physics (68th ed.). Boca
Laska, M., Seibt, A., & Weber, A. (2000). “Microsmatic” primates revis- Raton, FL: CRC Press.
ited—Olfactory sensitivity in the squirrel monkey. Chemical Senses, 25, Yee, K. K., & Wysocki, C. J. (2001). Odorant exposure increase olfactory
47–53. sensitivity: Olfactory epithelium is implicated. Physiology & Behavior,
Laska, M., & Teubner, P. (1998). Odor structure-activity relationships of 72, 705–711.
carboxylic acids correspond between squirrel monkeys and humans.
American Journal of Physiology, 274, R1639 –R1645.
Laska, M., Trolp, S., & Teubner, P. (1999). Odor-structure activity rela- Received May 5, 2003
tionships correspond between human and nonhuman primates. Behav- Revision received May 21, 2003
ioral Neuroscience, 113, 998 –1007. Accepted June 5, 2003 䡲

S-ar putea să vă placă și