Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/228348097

Thermal time - Concepts and utility

Article  in  Annals of Applied Biology · January 2005


DOI: 10.1111/j.1744-7348.2005.04088.x

CITATIONS READS

397 1,527

4 authors:

David Trudgill Alois Honek


James Hutton Institute Crop Research Institute
129 PUBLICATIONS   3,648 CITATIONS    192 PUBLICATIONS   6,314 CITATIONS   

SEE PROFILE SEE PROFILE

Daiqin Li Nico M. van Straalen


National University of Singapore Vrije Universiteit Amsterdam
198 PUBLICATIONS   3,673 CITATIONS    395 PUBLICATIONS   16,251 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The Study of the Function, Functioning and Evolution of Bird Dropping Masquerading by One Kinds Crab Spiders View project

insecticide resistance View project

All content following this page was uploaded by David Trudgill on 04 October 2019.

The user has requested enhancement of the downloaded file.


Annals of Applied Biology (2005), 146:1–14
1

Thermal time – concepts and utility


D L TRUDGILL1*, A HONEK2, D LI3 and N M VAN STRAALEN4

Scottish Crop Research Institute, Invergowrie, Dundee, Scotland DD2 5DA, UK


1
2
Research Institute of Crop Production, Drnoska 507, CZ-16106 Praha, 6-Ruzyne, Czech Republic
3
Biological Sciences, University of Singapore, 10 Kent Ridge Cresent, Singapore 1192600
4
Vrije Universiteit, Institute of Ecological Science, De Boelelaan 1087, 1081 HV Amsterdam, The
Netherlands

Summary
This paper (i) reviews temperature/development rate relationships in plants and poikilothermic
invertebrates, (ii) argues that the relationship is often linear over much of the range up to the thermal
optimum (To) and provides a possible mechanism, (iii) provides evidence of a trade-off between the
base temperature (Tb) and the thermal constant (DD) that enables each species to adapt to its thermal
environment, and (iv) indicates some of the practical and ecological implications. Where a linear
relationship has been characterised it is possible to estimate the base temperature for development (Tb,
expressed in °C) and the thermal constant for development (DD, the reciprocal of the temperature
coefficient (a), expressed in degree [°C] days accumulated above Tb). A possible basis for the linear
relationship between rate and temperature is proposed based on the Arrhenius and Sharpe-Schoolfield
equations involving activation enthalpy and progressive inactivation of the reactant molecules at both
low and high temperatures. Knowledge of Tb and DD enables rates of development of organisms/
processes to be calculated and compared at any given temperature between Tb and To. An analysis of
published results for differentiation processes (differentiation = a change of state) in species of insects,
Collembola, spiders, nematodes and plants showed that Tb tended to vary with the temperature of the
niche to which the organism is adapted, and that there was a trade-off between Tb and DD. Tropical
species had higher values of Tb than temperate and DD decreased as Tb increased (and vice versa). This
conferred a competitive advantage on each species in the thermal environment to which it was adapted.
The decrease in DD tended to be relatively greater than the increase in Tb, further favouring a high Tb
in tropical species. A mechanism for the trade-off is suggested whereby DD and Tb were shown to be
correlated (P < 0.01) with the activation enthalpy (HA) of an assumed, rate-limiting enzyme. Thermal
time can also be applied to processes involving growth (= an increase in dry weight) when the DD
requirement for development to maturity is the sum of the requirements for differentiation and growth.
Rates of both differentiation and growth can vary greatly between species, depending upon the niche
they inhabit, and the implications of such differences for resource requirements are considered. In
insects and nematodes, but not in annual plants, development is usually coupled to growth. Consequently,
when resources are inadequate, mature size in these animals varies less than in plants. Thermal time is
shown to provide insight into the life strategies of species within their communities and to have practical
implications.

Key words: Temperature, degree days, physiological time, threshold or base temperature, rates of
growth and development, resources, size, modelling, nematodes, insects, spiders, Collembola, plants

Introduction temperature (T e ) decreases, their rates of


development slow and, if the temperature falls low
Temperature is a basic factor that influences the enough, development will cease at their lower
course of life, specifically the rate of its change – developmental (base) temperature (T b ). As
physiological time. Thermal time provides a measure temperature increases, their rates of development
of physiological time as it relates to many increase up to a temperature optimum (To), above
poikilothermic species (those whose body which they again decrease and eventually cease at
temperature is effectively identical to that of their their temperature maximum (Tm). When data on
environment) of plants and invertebrate animals. durations of development at temperatures between
Most poikilothermic species are adapted to particular Tb and To are converted to their reciprocals, i.e. to
temperature ranges and temperature is often a major rates of development, the relationship between rate
environmental factor influencing their distribution. and temperature is often close to linear (e.g. Fig. 1)
Within those ranges, as the environmental for plants and many invertebrate animals

*Corresponding Author E-mail: jeantrudgill@yahoo.co.uk

© 2005 Association of Applied Biologists


2 D L TRUDGILL ET AL.

(ectothermic vertebrates appear to have more al., 1996). Pathologists, especially entomologists
complicated responses to temperature). From this (Honek (1996b) lists 335 references to the thermal
relationship two vital constants can be determined: time requirements of insects) have used thermal time
1) the thermal constant (the number of degree-days to determine and compare the requirements of the
– DD – needed to complete the developmental different stages and species of various pests and
process at any temperature within the range Tb to pathogens in relation to their biology (Dean, 1974;
To) and 2) the base temperature Tb. Griffith et al., 1996; Honek, 1996a,b; Tzortzakakis
The thermal constant (DD) requirement for a & Trudgill, 1996), epidemiology (Kaukoranta, 1996)
developmental process provides a measure of the and control (Forcella & Banken, 1996; Soto et
physiological time required for the completion of al.,1999), and as a basis for risk assessment
that process. It is expressed as the number of degree (Lahtinen et al., 1988). The ability to apply thermal
(in °C) days – hence DD and the term “thermal time”. time to the development of both pests and their plant
Together with Tb, it can be used to calculate the rate hosts makes it particularly useful when modelling
of the process at any value of Te (up to To). A crucial dynamic interactions (Ferris, 1976; Griffith et al.,
feature of thermal time is that the DD requirement 1996), including possible consequences of global
is based on a summation of the physiologically warming (Kaukoranta, 1996). Thermal time has also
effective temperature (i.e. Te minus Tb) over the been applied to other complex interactions, including
duration of the process, and expressing time in rates of soil organic matter decay and nitrogen
thermal units removes the time-dependency of utilisation (Smith et al., 1998; Dejoux et al., 2000).
biological processes where this is due to temperature Despite its potential utility, thermal time has not
change (Van Straalen, 1983). A linear relation been widely used by ecologists and there have been
between Te and the rate of the process under study is comparatively few thermal time studies applied to
essential for the estimation of both Tb (the intercept natural systems. These include a model of the effect
on the Te axis; Allee et al., 1949) and DD (the of climate change on tree species distributions and
reciprocal of the temperature coefficient (a), i.e. the dynamics in northern Europe (Sykes & Prentice,
reciprocal of the slope of the regression; Garcia- 1996), and another describing the phenology of nine
Huidobro et al., 1982). Once the values of Tb and wild plant species in relation to their use as climate
DD have been determined, the rate and duration of indicators (Spano et al., 1999). Others related to the
a process (e.g. moulting of arthropods, seed effects of thermal stratification in a marine
germination, etc.) at any temperature (between Tb environment on copepod abundance and diversity
and To) can be readily calculated (see Appendix 1, (Valdes & Moral, 1998), and to effects of incubation
Eqns 1 and 3). temperature on rates of embryonic development in
Thermal time is not the only approach to aquatic insects (Gillooly & Dodson, 2000). A
physiological time (e.g. Gillooly et al., 2001, particularly timely application was modelling the
proposed a unifying theory linking size, temperature effects of climate change on the distribution of
and metabolic rate across the animal kingdom), but poikilothermic animals close to their climatic limits
it has proved a useful concept. A search of biological (Thomas et al., 1999).
abstracting services (CAB, Web of Science, BIOSIS) This review essay provides references to, and
using the terms ‘thermal time’, ‘degree days’ and examples of, a linear relationship for various
‘thermal constants’ produced several hundred processes of development involving only
qualifying publications over the last 25 years. differentiation (= a change of state e.g. seed
However, the greater majority related to agriculture germination, egg hatch, metamorphosis) but also
and forestry. Thermal time has been used by applied shows that a linear relationship may sometimes be
biologists to analyse the effects of temperature on applied to processes involving growth (= increase
the germination of crop and weed seeds (Washitani in dry mass), such as development to maturity. We
& Takenaka, 1984a,b; Steinmaus et al., 2000), rates show that low T b values (< 8°C) are typically
of leaf and inflorescence appearance and associated with cool, and high values (> 12°C) with
development (Ellis et al., 1986, 1992; NeSmith, warm environments, and that the DD value reflects
1997; Granier & Tardieu, 1998; Yeh et al., 1999, both the amount and rate of differentiation and/or
2000), potential productivity (Black & Ong, 2000; growth. Based on a re-analysis of several data sets,
Clifton-Brown et al., 2000), and determination of we propose that the adaptation of poikilothermic
optimum planting/harvesting/spraying dates (Wilson animals and plants to their thermal environment
et al., 1999; Fisher & Lieth, 2000; Steinmaus et al., involves a trade-off between the values of Tb and
2000). Thermal time has been used to model the DD whereby one increases as the other decreases. It
effects of climate change, both on plants (Pearson is evident that the DD requirement for development
et al., 1997) and their pathogens (Kaukoranta, 1996), to maturity has implications for resource
and the dynamics of the interaction between requirements and we show that plants and animals
pathogen and host (Mullens et al., 1995; Griffith et have different strategies for responding to resource
Thermal time – concepts and utility 3

limitations. The value of this information is that it external stimuli (e.g. to overcome dormancy or
provides a means of describing and comparing the diapause), and the external resources required for
rates and durations of processes in, and the thermal growth should be optimal or standardised. This essay
adaptations of, poikilothermic animals and plants. only relates to such active processes and stages, but
A linear relationship between the rate of a process the study of the additional effects of factors such as
and temperatures between Tb and To is fundamental varying external resources are not precluded,
to the thermal time method. However, until now there provided appropriate procedures and controls are
has been no explanation for such a linear relationship used.
(the relationship would be expected to be curvi-linear The data on development duration and temperature
based on Van’t Hoff’s law, whereby rates of are the basis for calculating the thermal constants
development approximately double for every for development (see Appendix 1). It is important
increase of 10°C). Crucially, therefore, we offer a to note that the smaller the DD, the greater the slope
physiological explanation for rates of development (a) of the linear part of the regression and that the
being linearly related to temperature. Overall, we rate and duration of development at any given value
propose that the thermal time approach has wide of Te (up to To) depends on the values of both Tb and
utility because it applies to both plants and many DD. Hence, in Fig. 2, a tropical species of root-knot
invertebrate animals, and because the values of Tb nematode (Meloidogyne javanica) has a smaller DD
and of DD for development to maturity provide a requirement and a higher Tb than a comparable
basis for characterising developmental and temperate species (M. hapla).
ecological strategies.
Development Rate vs Temperature – Why is it
Thermal Time – Methods and Limitations Linear?

Laboratory-based methods for assessing the Rates of development of poikilothermic species


thermal time requirements of poikilothermic are temperature-dependent because temperature
organisms have been described in numerous affects numerous physiological processes, including
publications (e.g. Covell et al., 1986; Leach et al., rates of diffusion and the degree of desaturation of
1986; Gillooly & Dodson, 2000; Fig. 1). All such membrane lipids. However, it seems probable that
studies are based on the Celsius rather than the
Kelvin (absolute) scale because the former is more
biologically relevant. For this reason, the Celsius
scale is also used here. In thermal time studies the
duration of the process being studied is accurately
measured at a series of precise constant temperatures,
preferably including at least one or two values close
to or beyond the extremes of the temperature range
the organism might naturally experience. Precise
start and end points are essential – in seed
germination studies the chosen end point is often
Rate (1/days)

50% of the seed producing a visible radical (1 mm)


(e.g. Ramin, 1997).
Development to maturity can be thought of as
having two components – differentiation and growth
(Van der Have & de Jong, 1996). Differentiation,
which does not require external resources, has been
most studied, but thermal time information can also
be applied to processes involving growth, which do
require external resources. To determine accurately
when the end point has been reached (i.e. the seed
has germinated, the egg has hatched, the insect has
matured etc.) the time interval between observations
should be small and, ideally, be varied in relation to
temperature and duration. The aim is to obtain
precise data on which to base a linear regression as
the estimates of Tb and DD are both derived from Temperature (°C)
the slope of that regression (a) and, therefore, are Fig. 1. Relationship between temperature and
inversely correlated. The process under study must development rate for one generation of the rhabditid
be continuous, it must be independent of any other nematode Goodeyus ulmi (Leach et al., 1986).
4 D L TRUDGILL ET AL.

enzymes are of central importance. The activity became non-linear towards the temperature
profiles of the enzymes of different species reflect extremes. Alternative models with a mechanistic
adaptations to different environments, including basis are those of Arrhenius and of Sharpe &
temperature (Somero 1986; Milanovic et al., 1989). Schoolfield (see Gutierrez, 1996). Arrhenius’s model
For example, in Colias butterflies the occurrence of extends Van’t Hoff’s law by incorporating the
alleles of phosphoglucose isomerases with different concept of activation energy whereby a temperature
kinetics was correlated with temperature-related increase acts not only to increase the kinetic energy
behavioural differences (Watt et al., 1983). However, of the reactants, but is also required to overcome
Van’t Hoff’s law assumes an exponential relationship the energy threshold (HA, ‘activation enthalpy’)
between enzyme reaction rate and temperature needed to form the activated molecular complex that
whereby a 10°C increase in temperature produces a precedes a chemical reaction. The Sharpe-
standard increase in rate – usually designated Q10. Schoolfield model further extends the Arrhenius by
By this “law”, which has a solid foundation in assuming three states for molecules in a chemical
physical chemistry, the response to Te (in °C) is reaction: i) inactive due to low temperature, ii) active,
exponential rather than linear. Consequently, the lack and iii) inactive due to high temperature (Sharpe &
of a theoretical basis in enzyme kinetics for a linear De Michele, 1977; Schoolfield et al., 1981).
relationship between temperature and development To reconcile the linear model with the Sharpe/
rate is a major factor deterring many from using the Schoolfield mechanisms, we propose that the linear
thermal time model. relation between Te and rate of development is the
Schoolfield et al. (1981) went some way to result of progressive inactivation of enzymes (i and
overcoming this objection by developing a iii above) as Te approaches To or Tb. This will have
biophysical model that described a linear response the effect of turning down both ends of the reaction
at intermediate temperatures, but the response still curve, diminishing the increasing acceleration effect
implied by Van’t Hoff’s law and the Arrhenius
model. This process is clearly evident as T e
approaches and then exceeds To when a progressively
increasing proportion of enzyme molecules become
Rate of development (1/number of days)

inactive, resulting in a progressive diminution in the


rate (Fig. 1), eventually decreasing to zero at Tm (the
value of Te when all the molecules are inactivated).

Relation to Thermal Environment and Trade-


off Between Tb and DD

Thermal constants provide practically and


ecologically useful information. Furthermore, as
they differ, even between related species (because
of adaptation of each to its environment), their study
reveals various aspects of temperature adaptation.

Differences in Tb
The temperature ranges where growth and
development of poikilothermic species may
potentially occur are reflected in their values of Tb,
To, and Tm (Scholander et al., 1953). Pritchard et al.
(1996) gave a range from –5 to > 15°C as the
Temperature (°C) estimated Tb values for egg development of 40
populations of aquatic insects. Tb values below 0°C
Fig. 2. Relationship between temperature and duration were assigned to the eggs and larvae of several
of one generation of a temperate (Meloidogyne hapla, species of stonefly (Plecoptera) (Brittain & Mutch,
Tb = 8.25°C and DD = 554 °C days) and tropical (M.
javanica, Tb = 12.9°C and DD = 350 °C days) root-
1984; Brittain et al., 1984; Mutch & Pritchard, 1986)
knot nematode. Their rates of development are and the Antarctic beetle Hydromedion sparsutum
identical at 21°C (where the regression lines intersect). (Meyer-Arndt, 1984). Even though little
Below this temperature the temperate species has the development could occur below 0°C, the occurrence
greater rate of development, and above this temperature of such low estimated Tb values indicates the extent
the tropical species has the greater rate (data from of low-temperature adaptation. There is also
Trudgill & Perry, 1994). abundant evidence that tropical species have higher
Thermal time – concepts and utility 5

Tb values than equivalent temperate species (e.g. compared five temperate and six tropical legumes
Angus et al., 1981; Honek, 1996b). The widespread and demonstrated that the Tb for germination was
construction of glasshouses in temperate regions is consistently less for the temperate (mean 1°C) than
a practical demonstration of the higher temperature the tropical (mean 9°C). All the temperate legumes
requirements of plants from warmer latitudes; and had high DD requirements (mean 152 °C days) but,
pests from similar latitudes also flourish in such whilst the DD requirements of the tropical legumes
glasshouses. From published data (n = 335), Honek were generally less (mean 87 °C days), they were
(1996b) estimated that the mean Tb values for the more variable. Craufurd et al. (1996) conducted two
differentiation of insect eggs from temperate (40– experiments studying the germination of 12 cowpea
60 oN or S) and tropical (0–23 oN or S) regions were genotypes from West Africa and the estimated values
8.7°C and 12.9°C respectively (significantly of Tb were highest, and those of DD generally lowest,
different P < 0.01). in the genotypes originating from close to the
equator. Based on a linear regression (Fig. 3; r2 =
Trade-off between Tb and DD and the competitive 0.63, P < 0.001), DD decreased from 50 to 23 °C
value it confers days as Tb correspondingly increased from 6.5°C to
T b defines the lower temperature limit for 10.3°C. However, this inverse relationship was not
development and the reason for temperate species observed in a third experiment with a wider
having low values is self-evident. However, the geographical range of genotypes.
reason for T b being higher in tropical than in The cowpea data of Craufurd et al. (1996; Fig. 3)
temperate species is less obvious and presents a was almost as effectively described by a power
conundrum, as the effective temperature available function as by a linear regression (r2 = 0.61 and 0.63
for development (Te minus Tb, Appendix 1, Eqns 2 respectively). A linear regression indicated there was
and 3) decreases as Tb increases (except where Te > > 2-fold range in the DD requirement (50 to 22 °C
To; Honek & Kocourek, 1990; Trudgill, 1995a). In days) compared with a 1.5-fold range for Tb (6.9°C
this section we provide further evidence to support to 10.4°C). With these values, the most warm-
the suggestion of Honek & Kocourek (1990) and adapted genotype will germinate more rapidly than
Trudgill (1995a) that the answer to this conundrum the most cold-adapted at all values of Te > 13.1°C.
lies in a trade-off between Tb and DD. We further Based on data sets for many insect species Honek
suggest that the value of Tb and the relative rate of & Kocourek (1988, 1990) and Honek (1996a)
the trade-off between Tb and DD modify fitness and suggested an inverse relationship between the values
thus influence the outcome of competition between of Tb and DD for the differentiation of eggs and of
related species and their distributions. We examine pupae. Honek (1996b) recalculated the thermal
several data sets, five of them in considerable detail, requirements of 335 species in relation to their origin
to reveal examples of adaptive trade-offs between and demonstrated that, despite considerable
Tb and DD and suggest that the rate of trade-off variation, Tb tended to decrease with increasing
between Tb and DD is such that, as Te increases, it latitude of origin whereas the value of DD tended to
tends to favour tropical species. The evidence we increase. The data sets of Honek (1996a) are
provide is based on data sets for processes only relatively variable because of the large number of
involving differentiation. We have avoided analysing different experiments and unrelated species
data for processes involving growth because the DD involved, but an analysis of the data for the
requirement may be further influenced by other differentiation of insect pupae shows that the
factors (e.g. differences in adult size, resource
availability etc.).
A complication when comparing the Tb and DD
requirements of species from different latitudes is
that the data for individual species often come from
Tb (°C)

different studies using different methods and over


different temperature ranges. However, there are
exceptions. Angus et al. (1981) compared the
germination rates of seed from tropical and
temperate crops in field trials at different sites in
Australia. They found that the seed of the temperate
crops germinated at lower temperatures, but had a
greater DD requirement for germination than those DD (°C days)
of the tropical crops (e.g. Tb and DD for germination Fig. 3. Relationship between Tb and DD for 12 cowpea
of barley seed were c. 2.6°C and 80 °C days genotypes from West Africa. Data from Craufurd et
compared with c.16°C and 20 °C days for sesame). al. (1996), Expts 1 and 2. A linear regression was fitted
Similarly, in laboratory studies, Qi et al. (1999) (y = –0.14x + 13.5; r2 = 0.63).
6 D L TRUDGILL ET AL.

relationship between Tb and DD is best described suggesting that they are distinct and different. Van
by a linear regression (r2 = 0.24). Despite the Straalen (1994) concluded that differences in their
variability, the relationship is highly significant (P rates were associated more strongly with their habitat
< 0.001) and indicates a range in the DD requirement than with latitude. In warm environments (> 15°C),
from 400 to 20 °C days and for Tb of 3.1°C to 13.5°C. the eggs of above-ground species developed more
From this it can be calculated that at Te = 14°C and rapidly than those that lived below-ground, or in
22°C the most cold-adapted species will require 34 caves. The data sets of Van Straalen (1994) are reused
and 21 days for egg hatch compared with 40 and 2.4 here to estimate the values of Tb and DD for each
days respectively for the warm-adapted. species (typically r2 > 0.9) and a strong negative
Li (1998) also reported a strong, inverse correlation was demonstrated between Tb and DD.
relationship between T b and DD for egg A logistic function gave a better fit to the data (r2 =
differentiation in published data for 17 species of 0.70; Fig. 5) than a power function (r2 = 0.56) and
spiders from different environments. A curve fitted variation in Tb (–3.8°C to 11.4°C) is large, as is the
to this data using a power function (Fig. 4; r2 = 0.74, corresponding range for DD (650 to 50 °C days).
P < 0.001) showed that the value of DD ranged from Consequently, it can be calculated that the rate of
64 to 217 °C days (× 3.4 range) whereas the hatch of the eggs of warm-adapted Collembola
corresponding values of Tb ranged from 15.7 to 5.6°C species will be greater than that of the most cold-
(× 2.8 range). The competitive advantage as Te adapted for all values of Te > 12.7°C (e.g. at Te =
increases of a greater rate of change of DD as 18°C the cold-adapted and the warm-adapted species
compared to Tb is demonstrated below by using the will require 29.8 and 7.6 days respectively). A
extremes of the ranges as examples of a “temperate” subsequent study (Birkemoe & Leinaas, 2000) of
and a “tropical” species. With Te = 18°C, 21°C and the differentiation of the eggs of an arctic species of
24°C, egg hatch will take 17.5, 14.1 and 11.8 days Collembola (Hypogastrura tullbergi) suggested
respectively for the “temperate” species compared values of Tb and DD (–1.32°C and 295 °C days
with 27.8, 12.1 and 7.7 days for the “tropical” respectively) consistent with the regression in Fig.
species. The “tropical” species will have a more rapid 5.
rate of development than the “temperate” at all Differences in T b and DD may also reflect
values of Te > 19.9°C, whereas the reverse will apply differences in the time of year when species are most
below 19.9°C. active. Anderson & Coleman (1982) speculated that
Variations in the Tb vs. latitude relationship were four species of bacterial feeding nematodes were
observed by Honek (1996b) and by Li (1998) and able to cohabit because they had different
this is partly because geographic origin is only temperature ‘niche breadths’. Support for their
loosely related to the microclimate of the habitat and hypothesis came from Ferris et al. (1996) who
the season in which organisms develop. This is observed Tb and DD differences in six species of
demonstrated by Van Straalen (1994) who analysed bacterial feeding nematodes that all came from one
published data on the relationship between site. Their Tb values ranged from 0°C to 14.8°C and
temperature and the duration of egg differentiation the corresponding values of DD for egg
of 40 species of Collembola. A re-analysis of his
data showed that the 95% confidence limits for the
data points for each species typically overlapped
over only a small part of the total range, strongly
Tb (°C)
Tb (°C)

DD (°C days)
Fig. 5. Relationship between T b and DD for
embryogenesis in 38 species of Collembola. The data
is the same as that analysed by Van Straalen (1994).
DD (°C days) The data were pooled where more than one study was
Fig. 4. Relationship between Tb and DD for 17 species done on a species. A log regression line was fitted to
of spider (Arachnida). Data from Li (1998). Line fitted describe the relationship between T b and DD (y
using a ‘power’ function (y = 533.48x–0.847; r2 = 0.74). = –5.922logx + 34.7; r2 = 0.7).
Thermal time – concepts and utility 7

differentiation ranged from 36 to 10 °C days and for conformational changes in biological systems.
development to maturity from 173 to 35 °C days Support for this hypothesis comes from Van Straalen
respectively. Fitting a ‘power’ function showed that (1994), who used a modified Sharpe-Schoolfield
DD was inversely correlated with Tb (r2 = 0.91 and equation to estimate the kinetics of a hypothetical,
0.63, and P < 0.001, < 0.05 for development of eggs rate-controlling enzyme (see also Van der Have &
and development to maturity respectively). de Jong, 1996) and demonstrated a positive
Consequently, at Te = 16°C the most cold-adapted correlation between HA (activation enthalpy – the
species will hatch in less than half the time (2.25 energy required to change the state of the reactants
days compared with 5 days) than the most warm- from inactive to active) and rate of egg development
adapted, whereas at Te = 22°C the warm-adapted at 15°C of 38 species of Collembola. He concluded
species will hatch in 1.25 days compared with 1.62 that increasing HA was associated with more efficient
days for the cold-adapted. Such wide ranges in Tb use of each unit of temperature increase (i.e. with
and DD will potentially reduce competition between increasing a in Eqn 1 (Appendix 1), and therefore
nematodes adapted to the different seasons and soil decreasing DD). Regressing the estimated values of
horizons, but interpretation of the relationship for Tb and DD (Fig. 5) against the comparable estimates
development to maturity is complicated by of HA showed that there was a significant negative
differences in adult size and behaviour. correlation between HA and DD (r2 = 0.47; P < 0.01)
A slightly unusual, independent confirmation of a and a positive correlation with Tb (r2 = 0.43; P <
trade-off between Tb and DD in the development of 0.01) (Fig. 6a and b). These results support the view
the eggs of aquatic insects comes from Pritchard et that the apparent trade-off between Tb and DD has a
al. (1996). Instead of calculating the Tb for each basis in the thermal adaptation of enzymes.
species, they used 0°C as a universal Tb value when
calculating the DD requirements of 95 species over Other Factors Influencing DD Requirement
a range of values of Te. They then regressed the DD
requirements against Te and used the slope of the We have shown how and why the DD requirements
(linear) reaction norm as an index of adaptation to may be inversely related to Tb. However, DD also
low or high temperatures. Using this approach it can differs between species with a similar Tb. This
be readily calculated that, if there is a trade-off, as variation is due to innate differences in rates of
Te increases the DD requirement will gently increase differentiation and development.
for species with a true Tb < 0°C whereas it will
decrease for those with a true Tb > 0°C. The results Differences in rates of differentiation and
confirmed a trade-off as the species adapted to cold development
conditions had a positive (linear) regression, whereas The DD requirement for insect egg hatch range
for the more warm-adapted species there was a from < 20 to > 500 °C days (Honek, 1996a, 1999).
negative association that became more pronounced For insect eggs with a Tb of c. 9°C to 10°C, Honek
in proportion to the degree of adaptation. Overall, (1996a) reported differences in DD requirements for
the cold-adapted species had greater DD hatch ranging from < 10 °C days (e.g. Anaphes
requirements and much lower temperature ranges flavipes) to > 200 °C days (e.g. Typhlocyba
than the warm-adapted species. froggatti). This variation was only weakly correlated
These independent examples, derived from with egg mass. It was apparently due to species-
different plant and invertebrate taxa indicate a specific differences in rates of egg development
universal occurrence of a relationship between Tb (Honek, 1996a). Similarly, the DD requirements for
and DD. Explanations for differences in Tb and DD embryogenesis in nematodes range from < 10 to >
have mostly focused on the phenological and 260 °C days (Trudgill, 1995b). Eggs of the free-
evolutionary consequences of the temperature living nematode Caenorhabditis elegans (T b =
responses, e.g. the necessity to synchronise the 5.0°C) require 8.2 °C days to hatch (Ferris et al.,
various life-stages with the seasons (Gilbert & 1996) whereas those of the plant parasite Xiphinema
Raworth, 1996), and an understanding of the diversicaudatum (Tb = 7.6°C, Trudgill, 1995b)
underlying physiological mechanisms has been require 270 °C days to hatch. The much greater rate
lacking. of development of C. elegans compared with X.
diversicaudatum must reflect large differences in
Mechanisms their metabolic rates. The differences are so great
The proximate reasons for the variations in Tb, and that C. elegans can pass through several generations
for the trade-off between Tb and DD have been given (DD requirement for one generation = 43 °C days;
little attention. We attempt an explanation for the Trudgill, 1995b) in the time required for the eggs of
trade-off and suggest that the answer must be sought X. diversicaudatum to hatch.
in the fact that temperature affects not only the rate
of chemical reactions, but also induces
8 D L TRUDGILL ET AL.

a) b)
HA (kJ mol–1)

HA (kJ mol–1)
DD (°C days) Tb (°C)
Fig. 6. Regression of the activation enthalpy (HA) of a hypothetical rate-controlling enzyme for the embryogenesis of
38 species of Collembola (data from Van Straalen, 1994) and a) the value of DD and b) Tb calculated from the same
data. A log regression line was used for DD (y = –21.5logx + 183.5; r2 = 0.47), and an exponential line was fitted to
the data for Tb (y = 60.05 e0.039x; r2 = 0.43).

Development to maturity
As is demonstrated in Figs 1 and 7, linear
Rate (1/days)

relationships between Te and rate can also apply for


multi-step processes that involve both differentiation
and periods of growth (e.g. development to
maturity). However, this depends on there being a
linear relationship for each of the component stages
and processes (Fig. 7), and the greater the number
of developmental steps and the amount of growth,
the greater the DD (Honek, 1999). For arthropods,
the summation of the DD requirement for Temperature (°C)
development to maturity is simplified by the Fig. 7. Relationship with temperature for the
conclusion, derived from an analysis of data for development of eggs, nymphs and adults of Nysius
seven species of mites and 342 species of insects, vinitor. Redrawn from Kehat & Wyndham (1972).
that T b is usually the same for each of the
developmental stages within each species population influenced by i) the total increase in mass (P <
(Jarosik et al., 2003). 0.001), ii) taxonomic group (P = 0.124 to < 0.001),
and iii) trophic specialisation (P < 0.001) – predators
Size at maturity had smaller DD requirements than herbivores (P <
Size at maturity is a function of the rate and 0.005), presumably a reflection of the digestibility
duration of growth, and large size at maturity implies and/or nutritional quality of the different foods
a long generation time and a correspondingly large (Chapman, 1998). Life style also influences food
DD requirement. However, various other factors utilisation e.g. insects that fly have higher metabolic
influence the DD requirement associated with rates, and expend more energy than less active
growth. It is influenced by the quality and availability species, both when active and at rest (Reinhold,
of the resources required for growth, and the 1999). Consequently, DD requirements for
efficiency with which those resources are used for development to maturity ranged from < 100 °C days
growth. Honek (1996a) observed that differences in in some Aphis spp. (Tb c. 7°C–10°C) to > 4000 °C
DD requirements for insects to become adult were days in Periplaneta fuliginosa (Tb c. 12°C) (Honek,
associated more with differences in larval 1996a).
requirements than with those for the eggs or pupae. Large differences in mature mass do not signify
Honek (1999) showed that this variation in DD was exactly equivalent differences in DD because growth
Thermal time – concepts and utility 9

in plants and animals tends to be exponential rather DD requirement for development to flowering is
than linear, and is well approximated by power laws often fixed with the result that mature size decreases
(Gillooly et al., 2001; Charnov & Gillooly, 2003). as Te increases (Atkinson, 1994). Consequently, yield
For example, Ferris et al. (1996) observed that, when in temperate cereals is decreased in years with warm
cultured on bacteria, the mature mass of the summers because the crop matures more rapidly,
nematode Cruznema tripartitum (Tb 10°C) was 12- with a corresponding decrease in the duration of light
times greater than that of Caenorhabditis elegans interception (Atkinson & Porter, 1996).
(Tb c. 5°C). However, the DD requirement for Low temperatures have the opposite effect and
development to maturity of C. tripartitum was 55 mature size is often increased. In a review of the
°C days compared with 31 °C days for C. elegans. effect of temperature on the size of animals, plants,
A comparative analysis of their growth showed that, protists and bacteria, mature size increased as
up to 31 °C days, the two species were of a similar temperature decreased in 91 of 109 studies
size and that most of the greater mass of C. (Atkinson, 1994). This phenomenon is particularly
tripartitum was accumulated between 31 and 55 °C apparent in insects, e.g. black faced leaf hoppers
days (Ferris, personal communication). The rapid (Graminella nigrifrons) became c. 25% heavier
growth of C. tripartitum meant that it required only when grown at 17°C compared with 27°C (Larsen
20 °C days per µg increase in mass compared with et al., 1990). Various explanations have been offered
1300 °C days per µg for Cephalobus persignis, the for this effect, including the use of a biophysical
smallest species in the study (Ferris et al., 1996). model (Van der Have & de Jong, 1996) to show that
differences in a (the temperature coefficient for
Effects of sub-optimal resources on DD growth and differentiation) will lead to adult size
requirements and size varying with T e. However, in this model size
Sub-optimal availability of the resources required decreased sharply at both temperature extremes.
for growth may also affect the DD requirement and Gilbert (1984) and Gilbert & Raworth (1996) went
mass at maturity. However, the effects differ between one step further and, based on the effects of rearing
animals and plants. Plants, especially annuals, tend temperature on the size of Pieris rapae, proposed
to have a conserved DD requirement for that variations in adult size were due the values of
development to maturity that is largely unaffected Tb for differentiation being higher, and of a being
by the availability of resources and rates of growth. lower than that for growth. They proposed a
Consequently, when resources are scarce, mature functional explanation suggesting that, as fecundity
plant size and crop yields are greatly decreased is related to adult size, it is advantageous for
(sometimes > 100-fold; Squire, 1990; Atkinson, multivoltine insect species to maximise fecundity
1994) but the DD requirement is unchanged. In early in the year. But, as competition and predators
contrast, in insects, food scarcity increases the DD increase later in the year, reducing the generation
requirement by reducing the rate of development, time becomes most advantageous.
thereby increasing the feeding duration of each larval Where vital functions (e.g. feeding) are impaired,
stage, and sometimes the number of larval stages the relation between body size and temperature is
(Mukerji & Guppy, 1970; Honek, 1987; Chapman, sometimes reversed at values of Te close to Tb. This
1998). Even so, adult size may be reduced by up to is evident in some insects where maximum adult
two thirds because the increased duration of feeding size is achieved at median temperatures (David et
does not fully compensate for the decreased rates of al., 1994).
food intake and the increased maintenance
requirement. Growth and development of armyworm Decreasing the DD Requirement
(Pseudaetia unipuncta) and of the spider Linyphia
triangularis were linearly related to rates of food As is demonstrated by C. elegans, and by many
intake with consequential effects on mature size, other species with short generation times, rapid
survival rates and fecundity (Turnbull, 1962; Mukerji development is a consequence of a small overall DD
& Guppy, 1970). The quality of the food will also requirement. Decreasing the number of
affect the DD requirement and mature size e.g. black developmental steps can also reduce the DD
faced leaf hoppers (Graminella nigrifrons) grew requirement. Tylenchid nematodes undergo their first
larger on maize than oats (Larsen et al., 1990). moult in the egg, in Meloidogyne spp. (root-knot
nematodes) the third and fourth juvenile stages are
Effects of temperature on mature size also non-functional, and in the dorylaim nematode
Within the range Tb to To body size/mass typically Xiphinema americanum the number of juvenile
decreases as Te increases. This is because rates of stages has been decreased from four to three
differentiation are determined by thermal time but (Halbrendt & Brown, 1992). Many Meloidogyne
rates of resource acquisition and growth are also spp. are parthenogenetic, eliminating the time
influenced by calendar time. In annual plants the required for mating. Similarly, aphids have a small
10 D L TRUDGILL ET AL.

DD requirement for development to maturity (mean has important practical implications (e.g. it has
127 °C days; Honek, 1996a). This is achieved by implications for the effectiveness of plant breeding
being parthenogenetic, by larval development aimed at adapting genotypes to different thermal
beginning within the body of the mother even before environments). But, whilst these results help to
her moult to adult, by producing live young (all of provide a theoretical basis for the thermal time
which allows ‘telescoping’ of generations; Dixon, approach, in practice poikilothermic organisms must
1987), by small size, and by feeding almost function as a whole and their responses to
continuously. temperature are unlikely to be dependent on a single,
rate-limiting enzyme.
Discussion and Conclusions Before discussing the ecological consequences of
the linear relationship we need to consider the
Thermal time provides a means of determining confounding factors originating from experimental
and comparing rates of development of error. Estimates of DD and Tb are both derived from
poikilothermic organisms at temperatures within the a (the slope of the regression) and, consequently,
range Tb to To. It has wide utility because it applies are inversely correlated. Hence, an over-estimate of
to many plants and invertebrate animals. Differences a will result in an over-estimate of T b and a
in Tb indicate when organisms can become active in corresponding under-estimate of DD (and vice versa
a variable thermal environment and, together with – see Appendix 1). We explored the effects of such
To and Tm, their temperature ranges. In particular, experimental errors in a series of simulations and
the values of, and trade-off between Tb and DD, the results (not shown) indicate that the trade-off
provide insight into the adaptation and the ecological between DD and Tb observed in Figs 3 to 5 is much
strategies of organisms in relation to related species greater than expected from correlated measurement
and their thermal environment. errors. In particular, in the simulated results, the
The approach to thermal time described here estimates of Tb varied proportionately much more
differs from that of the unifying model of Gillooly than that of DD, whereas in the biological data the
et al. (2001) in four ways. Firstly, our approach reverse was observed. Additionally, we were able
differs in terms of scale. Gillooly et al. (2001) to demonstrate that the ranges of some of the data
assume a universal temperature relationship over the sets we examined (e.g. for Collembola) were distinct
whole of the animal kingdom. Consequently, and therefore the differences were unlikely to be due
differences within taxa are averaged out, whereas to error.
our approach focuses on such differences because Based on a linear relationship, we have shown that
of their agricultural and ecological importance. there is an inverse relationship between Tb and DD
Secondly, our approach applies only to associated with latitude and/or habitat that adapts
poikilothermic invertebrates rather than to all each species to its thermal environment (Figs 3 to
animals. Thirdly, it can also be applied to plants. 5). A trade-off between Tb and DD, especially one
And fourthly, based on enzyme kinetics, Gillooly et where the rate of change of DD is greater than that
al. (2001) assume that the rate of a process will of Tb, makes physiological and ecological sense
increase exponentially rather than linearly with because it leads to each species and its ecological
temperature. Furthermore, having seen an early draft strategy being adapted to its particular thermal
of this paper, Charnov & Gillooly (2003) environment. This is demonstrated by species of
subsequently proposed that those linear relations that bacterial-feeding nematodes that co-existed at one
have been described are ‘linear approximations of site (Ferris et al., 1996), whereby differences in
an exponential response’. However, a linear thermal requirements appear to help position the
relationship between temperature and rate of development of each species in relation to that of
development is fundamental to the thermal time the other species within the community. Ferris et
model. We believe that there are too many examples al. (1995) linked physiology, thermal time and
of a linear relationship between rate of development ecology by relating rates of development to those of
and Te for all of them to be ‘linear approximations’ CO2 evolution for eight species of bacterial-feeding
as proposed by Charnov & Gillooly (2003). nematodes grown at five temperatures. They showed
Crucially, therefore, we provide an explanation for that, at each temperature, unit metabolic rates varied
the linear relationship by suggesting that the with the thermal adaptation of the species and
proportion of enzyme molecules in an active state concluded that such differences may contribute to
progressively decreases as the temperature the changes in predominance of species at different
approaches Tb and To. We also demonstrate a negative times and soil depths.
relationship between the estimates of HA for a Our analyses of the trade-off between Tb and DD
hypothetical, rate limiting enzyme and DD, and a are based on comparisons between related species.
positive relationship with Tb. Such a relationship can Within species, the DD requirement for development
be predicted on theoretical grounds and, if correct, to maturity may differ due to differences in size and/
Thermal time – concepts and utility 11

or growth rates, but much of the available evidence many species, water availability in relation to
indicates Tb is less variable. Jarosik et al. (2003) transpiration demand is an important factor for
observed that Tb was usually the same for each plants, pore space availability is important for soil-
developmental step within a species of mite and dwelling nematodes, etc. Also, although the thermal
insect. Gomi (1996) examined the differentiation in time model applies equally to aquatic organisms and
populations in Japan of the fall webworm to those that live above- and below-ground, many
(Hyphantria cunea), an introduction from North above-ground organisms are able to increase their
America. Populations in the south-west had shifted body temperature using radiant (solar) heat or by
from two (bivoltine) to three (trivoltine) generations activity. For above-ground species, Te frequently
per year. Tb values were similar for both bi- and varies greatly and, for periods, may be below Tb or
trivoltine populations, but the results indicated that above To. Such seasonal or diurnal variations in Te
in the trivoltine population the DD requirement for complicate the interaction between organisms and
development to maturity had decreased (particularly their environment and, in some species, possibly
for the larval stages). This change was also because of heterogeneity, there may be a lack of
accompanied by a change in the photoperiodic linearity close to Tb that could be important early in
response controlling pupal diapause (Gomi & the growing season and when modelling dynamic
Takeda, 1996). Stacey & Fellows (2002) observed processes e.g. arthropod population dynamics
no significant differences in T b within three (Baker, 1972; Stinner et al., 1974; Diekkruger &
populations of Thrips major, three of T. tabaci or Roske, 1995). The ecological strategies of
two of Frankliniella occidentalis, but there were poikilothermic species are further fine-tuned by
significant differences in body size. Using published many other factors including photoperiodism,
data, they found a negative relationship between Tb dormancy and diapause that may be much more
and DD for F. occidentalis and T. tabaci, and important than temperature in determining onset of
suggested that a trade-off between Tb and DD may activity, or development. Physiological condition
constrain adaptation to local conditions. Even so, also has an influence on rates of processes (e.g. old
Craufurd et al. (1996) reported differences in Tb in seed with reduced viability will germinate more
cultivars of cowpea, and correlated differences in slowly than new seed), and heterogeneity,
DD in two of three experiments, and Mohamed et particularly in DD requirements for certain
al. (1988) observed differences in Tb, To and Tm in processes, is probably widespread (Trudgill et al.,
four cultivars of groundnut and four of pearl millet. 2000). Consequently, thermal time information is
Size at maturity is of practical and theoretical most valuable when used in conjunction with
interest because of its relationship to crop yields, to knowledge of the many other relevant environmental
generation time, to the fecundity of weeds and pests, and biological factors.
and to the relative performance of interacting Information on thermal constants helps provide
components of ecosystems. We have reported how insights into the position of species within
the availability of resources affects size at maturity communities. Values of Tb and DD for each stage in
and why, because of their conserved DD a life cycle define time to reproduction and longevity
requirement, plant size may vary greatly. This is of at any given temperature. If size at maturity is
major importance to agriculture because of its known, then relative rates of growth and their
implications for crop yields. However, in most resource requirements can be estimated. Pachepsky
invertebrate animals differentiation to the next larval/ et al. (2001) proposed that in natural communities
juvenile stage depends, at least partly, on an increase there is a trade-off between time to reproduction (i.e.
in mass and, when resources are limiting, DD) and fecundity that sustains diversity and
development is prolonged allowing more time for governs the form of the relative abundance and
feeding. Consequently, adult size varies within distribution. Hence, thermal time information has
narrower limits than for plants. Temperature has a value as a basis for characterising ecological
different influence on size. Increasing temperatures strategies, including modelling the relative growth
tend to decrease mature size in many plants and rates of interacting components of ecosystems – e.g.
invertebrate animals (Atkinson, 1994), probably Griffith et al. (1996) used it to model the interaction
because as Te increases the rate of differentiation and between an endoparasitic nematode and its clover
development tends to increase proportionally faster host.
than the rate of growth, with the result that the time Throughout this paper we assume that the thermal
available for resource capture is decreased. characteristics of species have been honed by
The thermal time model applies only to organisms selection to maximise their physiological fitness in
during their active stages and our analysis is far from relation to the thermal environment to which they
comprehensive (e.g. the range T o –T m is not are adapted. We suggest that the linear thermal time
considered). Tolerance of temperature extremes is model has wide relevance amongst invertebrates and
likely to be a major factor determining the ranges of plants and is conceptually and practically useful. The
12 D L TRUDGILL ET AL.

linear relation between temperature and rate 12:97–109.


simplifies the mathematics associated with Covell S, Ellis R H, Roberts E H, Summerfield R J. 1986.
The influence of temperature on seed germination rate in grain
calculating durations, and is the basis for estimating legumes. I. A comparison of chickpea, lentil, soyabean and
the cardinal values (Tb, To, DD). It has numerous cowpea at constant temperatures. Journal of Experimental
applications e.g. agriculturalists have used thermal Botany 37:705–715.
time information to schedule sowing and harvesting Craufurd P Q, Ellis R H, Summerfield R J, Menin L. 1996.
and to assess the risks posed by the accidental Development in cowpea (Vigna unguiculata). I. The influence
of temperature on seed germination and seedling emergence.
introduction of pests (e.g. Lahtinen et al., 1988). Experimental Agriculture 32:1–12.
Such thermal time analyses demonstrate that the David J R, Moreteau B, Gauthier J P, Petavy G, Stockel A,
population dynamics of many species are sensitive Imasheva A G. 1994. Reaction norms of size characteristics
to small differences in temperature, especially in in relation to growth temperature in Drosophila melanogaster:
geographical regions close to the edge of their range. an isofemale line analysis. Genetics Selection Evolution
26:229–251.
Consequently, global warming perhaps provides the Dean G J. 1974. Effect of temperature on the cereal aphid
greatest incentive for agriculturalists and ecologists Metapolodium dirhodum (Wlk.), Rhopalosiphum padi (L.)
to utilise thermal time. and Macrosiphum avenae (F.) (Hem., Aphididae). Bulletin
of Entomological Research 63:401–409.
Acknowledgements Dejoux J-F, Recous S, Meynard J-M, Trinsoutrot I, Leterme
P. 2000. The fate of nitrogen from winter-frozen rapeseed
leaves: mineralization, fluxes to the environment and uptake
Drs H Ferris for supplying data on nematode sizes by rapeseed crop in spring. Plant and Soil 218:257–272.
and rates of development, G Squire for his Diekkruger B, Roske H. 1995. Modelling the population
encouragement and support, D Mackerron and B dynamics of Isotoma notabilis (Collembola) on sites of
Marshall for help with mathematical modelling, and different agricultural usage. Pedobiologia 39:58–73.
Dixon A F G. 1987. Parthenogenetic reproduction and the rate
C Scrimgeour for discussing enzyme kinetics. of increase in aphids. In Aphids – Their Biology, Natural
Enemies and Control. Vol. A, pp. 269–287. Eds A K Minks
References and P Harrewijn. Amsterdam, Oxford, New York & Tokyo:
Elsevier.
Allee W C, Emerson A E, Park T, Schmidt K P. 1949. Ellis R H, Covell S, Roberts E H, Summerfield R J. 1986.
Principles of Animal Ecology. Philadelphia and London: W. The influences of temperature on seed germination rate in
B. Saunders Co. grain legumes. II Intraspecific variation in chickpea (Cicer
Anderson R V, Coleman D C. 1982. Nematode temperature arietinum L.) at constant temperature. Journal of
responses: a niche dimension in populations of bacterial- Experimental Botany 37:1507–1515.
feeding nematodes. Journal of Nematology 14:69–75. Ellis R H, Summerfield R J, Edmeades G O, Roberts E H.
Angus J F, Cunningham R B, Moncur M W, MacKenzie D 1992. Photoperiod, temperature, and the interval from sowing
H. 1981. Phasic development in field crops I. Thermal to tassel initiation in diverse cultivars of maize. Crop Science
response in the seedling phase. Field Crops Research 3:365– 32:1225–1232.
387. Ferris H. 1976. Development of a computer-simulation model
Atkinson D. 1994. Temperature and organism size – A biological for a plant-nematode system. Journal of Nematology 8:255–
law for ecotherms. Advances in Ecological Research 25:1– 263.
59. Ferris H, Lau S S, Venette R C. 1995. Population energetics
Atkinson D, Porter J R. 1996. Temperature, plant development of bacterial-feeding nematodes: respiration and metabolic
and crop yields. Trends in Plant Science 1:119–123. rates based on CO2 production. Soil Biology and Biochemistry
Baker C R B. 1972. An approach to determining potential pest 27:319–330.
distribution. EPPO Bulletin 3:5–22. Ferris H, Eyre M, Venette R C, Lau S S. 1996. Population
Birkemoe T, Leinaas H P. 2000. Effects of temperature on the energetics of bacterial-feeding nematodes: stage-specific
development of an arctic Collembola (Hypogastrura development and fecundity rates. Soil Biology and
tullbergi). Functional Ecology 14:693–700. Biochemistry 28:271–280.
Black C, Ong C. 2000. Utilisation of light and water in tropical Fisher P R, Lieth H J. 2000. Variability in flower development
agriculture. Agricultural and Forest Meteorology 104:25–47. of Easter lily (Lilium longiflorum Thunb.): model and
Brittain J E, Mutch R A. 1984. The effect of water temperature decision-support system. Computers and Electronics in
on the egg incubation period of Mesocapnia oenone Agriculture 26:53–64.
(Plecoptera) from the Canadian Rocky Mountains. Canadian Forcella F, Banken K R. 1996. Relationships among green
Entomologist 116:549–554. foxtail (Setaria viridis) seedling development, growing degree
Brittain J E, Lillehammer A, Saltweit S A. 1984. The effect days, and time of nicosulfuron application. Weed Technology
of temperature on intraspecific variation in egg biology and 10:60–67.
nymphal size in the stonefly, Capnia atra (Plecoptera). Garcia-Huidobro J, Monteith J L, Squire R C. 1982. Time,
Journal of Animal Ecology 53:161–169. temperature and germination of Pearl Millet. Journal of
Chapman R F. 1998. The Insects: Structure and Function. 4th Experimental Botany 33:288–296.
edn. Cambridge: Cambridge University Press. Gilbert N. 1984. Control of fecundity in Pieris rapae II.
Charnov E L, Gillooly J F. 2003. Thermal time: body size, Differential effects of temperature. Journal of Animal Ecology
food quality and the 10 °C rule. Evolutionary Ecological 53:589–597.
Research 5:43–51. Gilbert N, Raworth D A. 1996. Insects and temperature – a
Clifton-Brown J C, Neilson B, Lewandowski I, Jones M B. general theory. The Canadian Entomologist 128:1–13.
2000. The modelled productivity of Miscanthus × giganteus Gillooly J F, Dodson S I. 2000. The relationship of egg size
(GREEF et DEU) in Ireland. Industrial Crops and Products and incubation temperature to embryonic development time
in univoltine and multivoltine aquatic insects. Freshwater
Thermal time – concepts and utility 13

Biology 44:595–604. from South Georgia at different temperatures. Polar Biology


Gillooly J F, Brown J H, West G B, Savage V M, Charnov E 3:73–76.
L. 2001. Effects of size and temperature on metabolic rate. Milanovic M, Andjelkovic M, Stamenkovic-Bojic G. 1989.
Science 293:2248–2251. Adaptive significance of amylase polymorphism in
Gomi T. 1996. A mechanism for the decreased developmental Drosophila IV. A comparative study of biochemical properties
period of a trivoltine population of Hyphantria cunea of the alpha-amylase in Drosophila melanogaster, D. hydei,
(Lepidoptera: Arctiidae). Applied Entomology and Zoology D. subobscura, and D. busckii. Comparative Biochemistry
2:217–223. and Physiology 93B:629–634.
Gomi T, Takeda M. 1996. Changes in life-history traits in the Mohamed H A, Clark J A, Ong C K. 1988. Genetic differences
fall webworm within half a century of introduction to Japan. in the temperature responses of tropical crops. II. Seedling
Functional Ecology 10:384–389. emergence and leaf growth of groundnut (Arachis hypogaea
Granier C, Tardieu F. 1998. Is thermal time adequate for L.) and pearl millet (Pennisetum typhoides S. & H.). Journal
expressing the effects of temperature on sunflower leaf of Experimental Botany 39:1129–1135.
development? Plant, Cell and Environment 21:695–703. Mukerji M K, Guppy J J. 1970. A quantitative study of food
Griffith G S, Cook R, Mizen K A. 1996. Modelling consumption and growth of Pseudaletia unipuncta
development of stem nematode populations in white clover (Lepidoptera: Noctuidae). Canadian Entomologist 102:1179–
stolons. In Aspects of Applied Biology 46, Modelling in 1188.
Applied Biology: Spatial Aspects, pp. 253–256. Mullens B A, Paine E O, Velten R K. 1995. Temperature effects
Gutierrez A P. 1996. The role of abiotic factors. In Applied on survival and development of Heleidomermis magnapapula
Population Ecology: A Supply Demand Approach, pp. 27– in the laboratory. Journal of Nematology 27:29–35.
41. New York: John Wiley & Sons. Mutch R A, Pritchard G. 1986. Development rates of eggs of
Halbrendt J M, Brown D J F. 1992. Morphometric evidence some Canadian stoneflies (Plecoptera) in relation to
for three juvenile stages in some species of Xiphinema temperature. Journal of the North American Benthological
americanum sensu lato. Journal of Nematology 24:305–309. Society 5:272–277.
Honek A. 1987. Regulation of body size in the heteropteran NeSmith S D. 1997. Summer squash (Cucurbita pepo L.) leaf
bug, Pyrrhocoris apterus. Entomologica experimentalis et number as influenced by thermal time. Scientia Horticulturae
Applicata 44:257–262. 68:219–225.
Honek A. 1996a. The relationship between thermal constants Pachepsky E, Crawford J W, Brown J L, Squire G. 2001.
for insect development: a verification. Acta Societas Towards a general theory of biodiversity. Nature 410:923–
Zoologicae Bohemicae 60:115–152. 926.
Honek A. 1996b. Geographical variation in thermal Pearson S, Wheeler T R, Hadley P, Wheldon A E. 1997. A
requirements for insect development. European Journal of validated model to predict the effects of environment on the
Entomology 93:303–312. growth of lettuce (Lactuca sativa L.): Implications for climate
Honek A. 1999. Constraints on thermal requirements for insect change. Journal of Horticultural Science 72:503–517.
development. Entomological Science 2:615–621. Pritchard G, Harder L D, Mutch R A. 1996. Development of
Honek A, Kocourek F. 1988. Thermal requirements for aquatic insect eggs in relation to temperature and strategies
development of aphidophagous Coccinellidae (Coleoptera), for dealing with different thermal environments. Biological
Chrysopidae, Hemerobiidae (Neuroptera), and Syrphidae Journal of the Linnean Society 58:221–244.
(Diptera): some general trends. Oecologia 76:455–460. Qi A, Wheeler T R, Keatinge D H, Ellis R H, Summerfield
Honek A, Kocourek F. 1990. Temperature and development R J, Craufurd P Q. 1999. Modelling the effects of
time in insects: a general relationship between thermal temperature on the rates of seedling emergence and leaf
constants. Zoologische Jahrbucher Systematik 117:401–439. appearance in legume cover crops. Experimental Agriculture
Jarosik V, Honek A, Dixon A F G. 2003. Development rate 35:327–344.
isomorphy in insects and mites. American Naturalist 4:497– Ramin A A. 1997. The influence of temperature on germination
510. of taree Irani (Allium ampeloprasum L. ssp. iranicum W.).
Kaukoranta T. 1996. Impact of global warming on potato late Seed Science and Technology 25:419–426.
blight: risk, yield loss and control. Agricultural and Food Reinhold K. 1999. Energetically costly behaviour and the
Science in Finland 5:311–327. evolution of resting metabolic rate in insects. Functional
Kehat M, Wyndham M. 1972. The influence of temperature Ecology 13:217–224.
on development, longevity, and fecundity of the rutherglen Scholander P F, Flagg W, Walters V, Irving L. 1953. Climatic
bug, Nysius vinitor (Hemiptera: Lygaeidae). Australian adaptation in arctic and tropical poikilotherms. Physiological
Journal of Zoology 20:67–78. Zoology 26:67–92.
Lahtinen A E, Trudgill D L, Tiilikkala K. 1988. Threshold Schoolfield R M, Sharpe P J H, Magnuson C E. 1981. Non-
temperature and minimum thermal time requirements for the linear regression of biological temperature-dependent rate
complete life cycle of Meloidogyne hapla from Northern models on absolute rate theory. Journal of Theoretical Biology
Europe. Nematologica 34:443–451. 88:719–731.
Larsen K J, Madden L V, Nault L R. 1990. Effect of Sharpe P J H, De Michele D W. 1977. Reaction kinetics of
temperature and host plant on the development of the poikilotherm development. Journal of Theoretical Biology
blackfaced leafhopper. Entomological experimentalis et 64:649–670.
Applicata 55:285–294. Smith S R, Woods V, Evans T D. 1998. Nitrate dynamics in
Leach L, Trudgill D L, Gahan P B. 1986. Influence of food biosolids-treated soils. II. Thermal-time models of the
and temperature on the development and moulting of different nitrogen pools. Bioresource Technology 66:151–160.
Goodeyus ulmi (Nematoda: Rhabditida). Nematologica Somero G N. 1986. Protein adaptation and biogeography:
32:216–221. threshold effects on molecular evolution. Tree 1:124–127.
Li D. 1998. A linear model for description of the relationship Soto A G, Apablaza J H, Norero A S, Estay P P. 1999. Thermal
between the lower threshold temperature and thermal constant requirements for development of Trialeurodes vaporariorum
in spiders (Araneae: Arachnida). Journal of Thermal Biology (Hemiptera: Aleyrodidae) reared on tomato (Lycopersicon
23:23–30. esculentum). Ciencia E Investigacion Agraria 26:37–42.
Meyer-Arndt S. 1984. Growth and development of Spano D, Cesaraccio C, Duce P, Snyder R L. 1999.
Hydromedion sparsutum Mueller (Coleoptera: Perinylopidae) Phenological stages of natural species and their use as climate
14 D L TRUDGILL ET AL.

indicators. International Journal of Biometeorology 42:124– Watt W B, Cassin R C, Swan M S. 1983. Adaption at specific
133. loci. III. Field behaviour and survivorship differences among
Squire G R. 1990. The Physiology of Tropical Crop Production. Colias PGI genotypes are predicatable from in vitro
Wallingford, UK: CAB International for ODA. biochemistry. Genetics 103:725–739.
Stacey D A, Fellows M D E. 2002. Temperature and the Wilson D R, Cloughley C G, Sinton S M. 1999. Model of the
development rates of thrips: Evidence for a constraint on local influence of temperature on the elongation rate of asparagus
adaptation? European Journal of Entomology 99:399–404. spears. Acta Horticulturae 479:297–304.
Steinmaus S J, Prather T S, Holt J S. 2000. Estimation of Yeh D M, Atherton J G, Craigon J. 1999. A thermal time model
base temperatures for nine weed species. Journal of of post-initiation flower development in the shade plant
Experimental Botany 51:275–286. cineraria. Annals of Applied Biology 136:335–340.
Stinner R E, Gutierrez A P, Butler G D. 1974. An algorithm Yeh D M, Lin Y R, Atherton J G. 2000. A thermal time model
for temperature-dependent growth rate simulation. The for predicting time to aerial shoot elongation in Variegated
Canadian Entomologist 106:519–524. Solomon’s Seal. Annals of Applied Biology 136:69–75.
Sykes M T, Prentice I C. 1996. Climate change, tree species
distributions and forest dynamics: A case study in the mixed
conifer/northern hardwoods zone of Northern Europe.
Appendix 1
Climatic Change 34:161–177.
Thomas J A, Rose R J, Clarke R T, Thomas C D, Webb N R. The equation for a linear relationship between Te
1999. Intraspecific variation in habitat availability among and rate of development (for values of Te between
ectothermic animals near their climatic limits and their centres Tb and To) is:
of range. Ecology 13:55–64.
Trudgill D L. 1995a. Why do tropical poikilothermic organisms
tend to have higher threshold temperatures for development y = ax + b (Eqn 1)
than temperate ones? Functional Ecology 9:136–137.
Trudgill D L. 1995b. An assessment of the relevance of thermal where y is the development rate (1/duration of
time relationships to nematology. Fundamental and Applied development), x is the environment temperature (Te,
Nematology 18:407–417.
Trudgill D L, Perry J N. 1994. Thermal time and ecological
in °C), b is the intercept on the y axis, and a is the
strategies - a unifying hypothesis. Annals of Applied Biology temperature coefficient (slope) of the regression. Tb
125:521–532. is estimated by back extrapolation of the linear
Trudgill D L, Squire G R, Thompson K. 2000. A thermal regression and is the value of Te where development
time basis for comparing the germination requirements of ceases, and To is determined by the loss of linearity
some British herbaceous plants. New Phytology 145:107–114.
Turnbull A L. 1962. Quantitative studies of the food of Linyphia
as Te is increased (Fig. 1). It should be noted that Tb
triangularis Clerck (Araneae: Linyphiidae). Canadian = –b/a and that Eqn 1 can be rewritten in a
Entomologist 94:1233–1249. mathematically equivalent form, but with more
Tzortzakakis E A, Trudgill D L. 1996. A thermal time based biologically meaningful parameters as:
method for determining the fecundity of Meloidogyne
javanica in relation to modelling its population dynamics.
Nematologica 42:347–353.
Rate = (Te – Tb)/DD (Eqn 2)
Valdes L A, Moral M. 1998. Time-series analysis of copepod
diversity and species richness in the southern Bay of Biscay The thermal constant (DD) is the reciprocal of
off Santander, Spain, in relation to environmental conditions. slope (1/a, Eqn 1) and the duration of development
ICES Journal of Marine Science 55:783–792. for all values of Te between To and Tb is given by :-
Van der Have T M, de Jong G. 1996. Adult size in ectotherms:
temperature effects on growth and differentiation. Journal of
Theoretical Biology 185:329–340. Duration = DD/(Te – Tb) (Eqn 3)
Van Straalen N M. 1983. Physiological time and time-
invariance. Journal of Theoretical Biology 104:349–357.
Van Straalen N M. 1994. Adaptive significance of temperature (Revised version accepted 19 October 2004;
responses of Collembola. Acta Zoologia Fennica 195:135–
142.
Received 30 June 2004)
Washitani I, Takenaka A. 1984a. Mathematical description
of the seed germination dependency on time and temperature.
Plant, Cell and Environment 7:359–362.
Washitani I, Takenaka A. 1984b. Germination responses of a
non-dormant seed population of Amaranthus patulus Bertol.
to constant temperatures in the sub-optimal range. Plant, Cell
and Environment 7:353–358.

View publication stats

You might also like