Вы находитесь на странице: 1из 24

1.1. What is Population Ecology?

Population ecology relative to other ecological disciplines

Population ecology is the branch of ecology that studies the structure and dynamics of populations. Physiology studies individual characteristics and individual processes. These are use as a basis for prediction of processes at the population level. Community ecology studies the structure and dynamics of animal and plant communities. Population ecology provides modeling tools that can be used for predicting community structure and dynamics. Population genetics studies gene frequencies and microevolution in populations. Selective advantages depend on the success of organisms in their survival, reproduction and competition. And these processes are studied in population ecology. Population ecology and population genetics are often considered together and called "population biology". Evolutionary ecology is one of the major topics in population biology. Systems ecology is a relatively new ecological discipline which studies interaction of human population with environment. One of the major concepts are optimization of ecosystem exploitation and sustainable ecosystem management. Landscape ecology is also a relatively new area in ecology. It studies regional large-scale ecosystems with the aid of computer-based geographic information systems. Population dynamics can be studied at the landscape level, and this is the link between landscape- and population ecology.

The term "population" is interpreted differently in various sciences:


In human demography a population is a set of humans in a given area. In genetics a population is a group of interbreeding individuals of the same species, which is isolated from other groups. In population ecology a population is a group of individuals of the same species inhabiting the same area.

Interbreeding is seldom considered in ecological studies of populations. The exceptions are studies in population genetics and evolutionary ecology.

Populations can be defined at various spatial scales. Local populations can occupy very small habitat patches like a puddle. A set of local populations connected by dispersing individuals is called a metapopulation. Populations can be considered at a scale of regions, islands, continents or seas. Even the entire species can be viewed as a population. Populations differ in their stability. Some of them are stable for thousands of years. Other populations persist only because of continuous immigration from other areas. On small islands, populations often get extinct, but then these islands can be re-colonized. Finally, there are temporary populations that consist of organisms at a particular stage intheir life cycle. For example, larvae of dragonflies live in the water and form a hemipopulation (term of Beklemishev). The major problem in population ecology is to derive population characteristics from characteristics of individuals and to derive population processes from the processes in individual organisms:

Main axiom of population ecology: organisms in a population are ecologically equivalent. Ecological equivalency means: 1. Organisms undergo the same life-cycle 2. Organisms in a particular stage of the life-cycle are involved in the same set of ecological processes 3. The rates of these processes (or the probabilities of ecological events) are basically the same if organisms are put into the same environment (however some individual variation may be allowed).

1.2. Models as analytical tools


Population ecology is the most formalized area in biology. Model is a tool and should never be considered an ultimate goal in ecological studies. Model and reality are linked together by two procedures: abstraction and interpretation:

Abstraction means generalization: taking the most important components of real systems and ignoring less important components. Importance is evaluated by the relative effect of system components on its dynamics. For example, if we found that parasitism rate in insect pest is always below 5%, then parasitoids can be excluded from the model. Interpretation means that model components (parameters, variables) and model behavior can be related to components, characteristics, and behavior of real systems. If model parameters have no interpretation, then they cannot be measured in real systems. Most field ecologists are not good at abstraction. If they build a model they often try to incorporate every detail. Most mathematicians are not good at interpretation of their models. Usually they think of clean models and dirty reality. However, both abstraction and interpretation are necessary for successful modeling. Thus, close collaboration between ecologists and mathematicians is very important. Models are always wrong ... but many of them are useful. How it may happen that the wrong model can give a correct answer? In the same way as old maps, which assumed a flat earth and used wrong distance relations, where useful for travelers in the past. Modeling strategy: 1. 2. 3. 4. 5. Select optimal level of complexity Never plan model development for more than 1 year Avoid the temptation to incorporate all available information into the model Follow specific objectives, don't try to make a universal model If possible, incorporate already existing models

System properties and model properties 1. Many system properties are not represented in the model. Example: age structure is not represented in both exponential and logistic models. 2. Some model properties cannot be found in real systems. Example: solutions of differential equations are always smooth, while trajectories of real systems are always noisy. Example of a wrong question: Does this population have an equilibrium density? The stable equilibrium is a state to which all trajectories of the system converge infinitely close with increasing time. The model (e.g. the differential equation) may have an equilibrium density, but real populations don't have it because: 1. Population density cannot be measured with infinite accuracy.

2. Weather fluctuations always add noise to system's dynamics. 3. Time series are never long enough to talk about limits and convergence.

1.3. Population system


Population system (or a life-system) is a population with its effective environment. The term "life-system" was introduced by Clark et al. (1967). Later, Berryman (1981) suggested another term "population system" which is definitely better. A short review of the life-system methodology was published by Sharov (1991). Major components of a population system 1. Population itself. Organisms in the population can be subdivided into groups according to their age, stage, sex, and other characteristics. 2. Resources: food, shelters, nesting places, space, etc. 3. Enemies: predators, parasites, pathogens, etc. 4. Environment: air (water, soil) temperature, composition; variability of these characteristics in time and space. Temporal and spatial structure of a population system Temporal structure Diurnal cycles Seasonal cycles Long-term cycles Spatial structure Spatial distribution Habitat structure Metapopulations

Dynamics of Population Systems Factor-Effect concept: Environmental factors affect population density. When factors change then population density changes respectively. Advantage: Causal explanation of population change Disadvantage: It is useful for one-step instant effects but becomes confusing if the effect is simultaneously a factor that causes soemthing else, or if there are time delays in the effect. Factor-Process concept: Environmental factors do not affect population density directly, instead they affect the rate of various ecological processes (mortality, reproduction, etc.) which finally result in a change of population density. Processes can change the value of factors; thus, feedback loops become possible. Advantage: Can handle complex dynamic interactions among components of population systems including time-delays, negative and positive feedbacks. In 50-s and 60-s there was a discussion about population regulation between two schools in population ecology. An agreement could not be reached because these schools used different concepts of population dynamics. Andrewartha and Birch (1954) used the factor-effect concepts

whereas Nicholson (1957) used the factor-process concept. The factor-process concept works not only in population ecology but in any kind of dynamic systems, e.g. in economic systems. Forrester (1961) formalized the factor-process concept and applied it to industrial dynamics. Later this formalism became very popular in ecology and is widely used especially in ecosystem modeling. The model of Forrester is based on tank-pipe analogy. The system is considered as a set of tanks connected by pipes with vents which can regulate the "flow" of liquid from one tank to another. The flow of "liquid" between tanks is considered as "material flow". However there is also "information flow" that regulates the vents. Vents are equivalent to processes; and the amount of liquid in a tank is a variable or a factor because it can affect processes via information flow.

The figure above is the Forrester diagram for an insect population system with 4 stages: eggs, larvae, pupae, and adults. Transition between these states (development) is regulated by temperature. Influx of eggs depends on the number of adults. Mortality occurs in all stages of development. Larval and pupal mortality is density-dependent. Diagrams of Forrester can be easily transformed into differential equation models. Each process becomes a term in a differential equation that determines the dynamics of the variable. For example, the number of larvae in the figure above is affected by 3 processes: egg development ED(T), larval development LD(T), and larval mortality LM(N). Egg and larval development rates are functions of temperature T; whereas larval mortality is a function of larval numbers N. The equation for larval dynamics is the following: dN/dt = ED(T) - LD(T) - LM(N) Here the term ED(T) is positive because egg development increases the number of larvae ("liquid" influx). Terms LD(T) and LM(N) are negative because larval development (molting into pupae) and mortality reduce the numbers of larvae. Limitations of Forrester's model: 1. Distinction between information and material flows was not clear because information cannot be transferred without any matter. For example: egg or seed production is not only an information flow, there is a flow of matter too. 2. Only one type of processes is considered in which "fluid" moves from one tank to another. This is good representation of organisms changing their state. However, it is impossible to

apply the model to the processes that involve two or more participants. For example, when a parasite enters a host then it is impossible to make a "pipe" that starts from two "tanks": host and parasite, and ends in the "tank" of "parasitized host". 3. Only one level of processes is considered. In some cases it is important to consider processes at two spatial levels, e.g., the dynamics of phytophagous insects can be considered within a host plant and within a population of host plants. Other models has been developed to represent the factor-process concept. Petri nets consider interactions of one two, or more participants. However, there is no generic model with no limitations. Factors and processes The dynamics of a system is always viewed as a sequence of states. State is an abstraction like an arrow hanging in the air but it helps to understand the dynamics. Because systems are built from components, the state can be represented as a combination of states of all its components. For example, the state of the predator-prey system can be characterized by the density of prey (component #1) and density of predators (component #2). As a result, the state of the system can be considered as a point in a 2-dimensional space with coordinates that correspond to system components. This space is usually called the phase space Components of the system will be called factors because they affect system dynamics. The state of the component is the value of the factor. Factors are considered with as many details as necessary for understanding system's dynamics. Examples of factors:

In simple models (e.g., exponential or logistic), there is only one factor (=component): the population itself. Its value is population density. In age-structured populations, each age class is a factor and its value is equal to the density of individuals in that class. Natural enemies and various resources can be considered as additional factors. Then, the abundance of predators, parasites, or food will be values of these factors. Weather can be viewed as a set of factors: temperature, precipitation, etc. Each of these factors has a numerical value. Factors may have a hierarchical structure. If the population occupies multiple patches in space, then all factors (e.g., population density, density of natural enemies and resources, temperature) become patch-specific.

Any detectable change in the population system is considered as an event. Events can be classified according to the components involved in these events. The process can be defined as a class of identical events. The rate of a process can be measured by the number of events that occur in the system per unit time. Also, the specific rate of processes is often used which is the number of events per time unit per one organism involved in the event. Examples of processes

The birth of an organism is an event. The birth rate is the number of births per unit time. The specific birth rate (= reproduction rate) is the birth rate per 1 female (or per 1 parent). Death of an organism is an event. Mortality on a specific stage from a specific cause (e.g.,

parasitism, predation, starvation) is a process. Mortality rate is the number of deaths per unit time. Specific mortality rate is the number of deaths per unit time per organism. Other processes are: growth, development, consumption of resources, dispersal, entering diapause, etc.

Interaction of factors and processes Factors affect the rate of processes as shown below:

On the other hand, processes change the values of factors:

A process may be affected by multiple factors. For example, mortality caused by predation may depend on the prey density, predator density, number of refuges, temperature (if it changes the activity of organisms), etc. The value of a factor may change due to multiple processes. For example, the number of organisms on a specific stage changes due to development (entering and exiting this stage), dispersal, and mortality due to predation, parasitism, and infection. Thus, there is no one-to-one correspondence between factors and processes. Life-tables show the rate of various mortality processes, but they do not show the effect of factors on these processes. For example, parasitism may be mostly determined by weather; viral infection may be determined by host plant chemistry. But life-tables do not show the effect of either weather or and host plant chemistry. It is dangerous to judge on the role of factors (e.g., biotic vs. abiotic) from life-tables. For example, a life-table may show that 90% mortality of an insect is caused by parasitism. This may lead to an erroneous conclusion that parasites rather than weather are most important in the

change of population density. It may appear that the synchrony between life cycles of the host and parasite depends primarily on weather. Life tables show various mortality processes in the population but they do not indicate the role of factors. To analyze the role of factors, it is necessary to vary these factors experimentally and examine how they affect various mortality processes. These experiments may show, for example, that the rate of parasitism depends on the density of parasites, density of hosts, and temperature. To understand population dynamics, it is necessary to know both the effect of factors on the rate of various processes, and the effect of processes on various ecological factors (e.g., on population density). This information is integrated via modeling. References Berryman A. A. (1981) Population systems: a general introduction. New York: Plenum Press. Clark L. R., P. W. Geier, R. D. Hughes, and R. F. Morris (1967) The ecology of insect populations. London: Methuen. Sharov, A. A. 1992. Life-system approach: a system paradigm in population ecology. Oikos 63: 485-494. Get a reprint! (PDF)

Lecture 2. Estimation of population density and size 2.1. Censusing a whole population
This method works only if organisms can be easily observed, their numbers are not too big, and the area is well bounded and is not too large. Examples: trees in a small isolated forest; all bird nests are censused in the New-York state (see Pielou 1977). Migrating populations can be counted using aerial photography. This method is used when the population has seasonal migration. Examples: sandhill cranes and caribou.

2.2. Simple Random or Systematic Sampling


Two possible objectives for sampling: 1. Estimate average population density, 2. Make a map of population density. Traditionally, random sampling plans were preferred over systematic sampling plans because random sampling helped to avoid subjective selection of sample locations. However, systematic sampling has no elements of subjectivity if sample location is selected prior to examining the area. For example, there is no subjective decisions if we sample every tenth potato plant and count Colorado potato beetles. Moreover, systematic sampling has an advantage over random sampling if the number of samples is large because of more uniform coverage of the entire sampling area. It is especially important for making population maps. Random sampling can be used if the objective is to estimate the mean population density and the number of samples is not large (<100).

Preferential sampling of specific areas (e.g., high-density areas) was always considered unacceptable. However, modern geostatistical methods and stratified sampling can take advantage of preferential sampling. This shows that the methodology of sampling evolves and old textbooks may give obsolete recipes. Traditional statistical methods include estimation of the mean population density (M), standard deviation (S.D.), and standard error (S.E.), which is the standard deviation of the sample mean.

The equation for standard error is derived assuming that all samples are independent. This is a very strong assumption which is unrealistic in many situations. Samples separated by small distance are often positively correlated. Before using standard statistics it is important to test if samples are correlated. Spatial correlations are examined using geostatistics. The simplest geostatistical test for spatial autocorrelation is the omnidirectional correlogram:

where z1 and z2 are organism numbers in two samples separated by lag distance h, summation is performed over all pairs of samples separated by distance h; Nh is the number of pairs of samples separated by distance h; Mh and sh are the mean and the standard deviation of samples separated by distance h (each sample is weighted by the number of pairs of samples in which it is included). Correlation decreases with distance between samples as shown below.

The range of correlogram is the lag distance h at which correlation reaches (or becomes close to) zero. Standard statistics can be applied only if inter-sample distance exceeds the range of the correlogram.

2.3. How Many Samples to Take?


There are two major methods for planning the number of samples: 1. two-step sampling, and 2. sequential sampling:

Two-step sampling

The number of samples, N, required to achieve specific accuracy level can be estimated from equations for standard error (S.E.) and accuracy, A:

where M is sample mean and S.D. is standard deviation. Here the third equation is derived from the first two equations. Standard deviation, S.D., is usually not known before sampling. Thus, the first step is to take N1 samples and to estimate N using the equation above. Then, at the second step, take N1 = N - N1 samples. Taking samples in two steps is possible only if population numbers don't change between two sampling dates.

Sequential sampling
The main idea of sequential sampling is to take samples until some condition (which is easy to check) is met. The first example is the sampling plan targeted at achieving specific accuracy. It is based on the Taylor's power law:

Coefficients a and b can be estimated using linear regression from several pairs of M and S.D. estimated in different areas with different average population density. Combining two previous equations we get:

N= Mean (M) equals to the total number of recorded individuals (S) in all samples divided by the number of samples (N). Now, we substitute M by S/N, and solve this equation for S:

Stop-lines for accuracy levels of A = 0.1; 0.07; and 0.05 are plotted below:

The blue line shows the total number of captured individuals in all samples. Sampling terminates when this line crosses the stop line for selected accuracy level. The second example is the sequential sampling plan used for decision-making in pest management. This method was developed by Waters (1955; Forest Sci. 1:68-79). It is described in Southwood (1978).

Here the blue line again shows the total number of captured individuals in all samples. While the blue line is between magenta inclined lines, sampling continues. If the blue line crosses the upper magenta line, then sampling stops and pesticides are applied against the pest population. If the blue line crosses the lower magenta line, then sampling stops and pesticides are not applied. Deriving the solution of this problem it is too complicated. Thus, we will consider the final result only. If the population has a negative binomial distribution (see next lecture), then stop lines correspond to the linear equation:

where:

Confidence interval (c.i.) is the interval where the population mean can be found with probability of (1 - P), where P is error probability (e.g., P = 0.05). The number of degrees of freedom d.f. = N - 1 (one d.f. goes for estimation of sample mean). Precision of sample mean is A = S.E. / M There is an empirical rule that precision should be below 0.05 (or 0.1). However, this rule is not universal The only thing that matters in statistics is testing hypotheses. If null-hypothesis is rejected then it does not matter whether A was above or below 0.05. However, in each specific research area, it is useful to find a precision level which is usually sufficient for rejecting nullhypotheses.

Example:. Insect pest population should be suppressed if its density exceeds the economic injury level (EIL). A null-hypothesis is tested that the average density M is equal to EIL. If EIL is within the c.i. for M, then the null-hypothesis cannot be rejected and no decision can be made. In this case, more samples should be taken. If the EIL is outside of the c.i., then nullhypothesis is rejected, and population is suppressed if M > EIL, or not suppressed if M < EIL.

2.4. Elements of geostatistics


Geostatistics is a collection of statistical methods which were traditionally used in geo-sciences. These methods describe spatial autocorrelation among sample data and use it in various types of spatial models. Geostatistical methods were recently adopted in ecology (landscape ecology) and appeared to be very useful in this new area. Geostatistics changes the entire methodology of sampling. Traditional sampling methods don't work with autocorrelated data and therefore, the main purpose of sampling plans is to avoid spatial correlations. In geostatistics there is no need in avoiding autocorrelations and sampling becomes less restrictive. Also, geostatistics changes the emphasis from estimation of averages to mapping of spatially-distributed populations. Spatial autocorrelation can be analyzed using correlograms, covariance functions and variograms (=semivariograms). For simplicity, here we will use correlograms only. Covariance functions and variograms are discussed in the next lecture. In brief, geostatistical analysis usually has the following steps: 1. 2. 3. 4. Estimation of correlogram Estimation of parameters of the correlogram model Estimation of the surface (=map) using point kriging, or Estimation of mean values using block kriging

Detailed description of most geostatistical methods can be found in Isaaks and Srivastava (1989). Here we will discuss only the most important elements of geostatistics.

Estimation of Correlogram
Correlogram is a function that shows the correlation among sample points separated by distance h. Correlation usually decreases with distance until it reaches zero. Correlogram is estimated using equation:

where z1 and z2 are organism numbers in two samples separated by lag distance h, summation is performed over all pairs of samples separated by distance h; Nh is the number of pairs of samples separated by distance h; Mh and sh are the mean and the standard deviation of samples separated by distance h (each sample is weighted by the number of pairs of samples in which it is included). Notes:

1. This is an omnidirectional correlogram; "omnidirectional" means that we don't care about the direction of lag h. 2. Serious geostatistical analysis often includes estimation of directional correlograms (see next lecture). It may happen that points are more closely correlated in some direction (e.g., NE-SW) than in other directions. If correlogram depends on direction, then the spatial pattern is called anisotropic. If no anisotropy detected, then it is possible to use the omnidirectional correlogram. 3. Correlogram equation works only if there are no trends in population density in the study area. If a trend exists, then a non-ergodic correlogram should be used instead (see next lecture).

Estimation of parameters of correlogram model


The correlogram can be approximated by some mathematical model. Two models are used most often: 1. Exponential model:

2. Spherical model:

where c1 is sill, and a is range. These parameters can be found using the non-linear regression.

Estimation of the surface (=map) using point kriging (ordinary kriging)


The value z'o at unsampled location 0 is estimated as a weighted average of sample values z2 at locations i around it:

Weights depend on the degree of correlations among sample points and estimated point. The sum of weights is equal to 1 (this is specific to ordinary kriging):

Weights are estimated individually for each point in a regular spatial grid using the system of linear equations:

where is the Lagrange parameter; is the correlation between points i and j which is estimated from the variogram model using distance, h, between points i and j; 0 is the estimated point; 1,...,n are sample points. Using matrix notation this system can be re-written as:

The solution of this matrix equation is:

Now, weights are found, and thus, it is possible to estimate the value z'o. When these values are estimated for all points in a regular grid, then we get a surface of population density. The variance of local estimation is equal to:

Estimation of the mean value using block kriging


The only difference of block kriging from point kriging is that estimated point (0) is replaced by a block. Consequently, the matrix equation includes "point-to-block" correlations:

Point-to-block correlation is the average correlation between sampled point i and all points within the block (in practice, a regular grid of points within the block is used, as shown in the figure).

The variance of block estimation

is equal to:

Where is the average correlation within the block (average correlation between all pairs of grid nodes within the block).

Advantages of kriging:

It handles spatial autocorrelation It is not sensitive to preferential sampling in specific areas It estimates both: local population densities and block averages. It can replace stratified sampling if the size of aggregations is larger than the intersample distance.

2.4. Stratified sampling


Stratified sampling is used if sampled area (or volume) is heterogeneous (Pielou, p.107). If patch size is much larger than inter-sample distance, then kriging can be used instead of stratified sampling. In stratified sampling program, the area (volume) is subdivided into 2 or more portions which are sampled separately. Example: pine sawflies prefer to spin their cocoons close to the tree; thus, the area adjacent to trees (within 1 m radius) can be sampled separately from the rest of the area. 1. Density of samples in each stratum should be proportional to the variance of organism counts in the stratum -- in this case the maximum precision is reached. Taylor's power law can be used to predict variance from the mean. Variance usually increases with mean, and thus, the stratum with higher organism density should be sampled more intensively. 2. The mean population density, M, is estimated as a weighted mean of mean densities, Mi, in each stratum, i, with weights, wi equal to the area covered by stratum i:

3. The standard error, SE, of the mean is equal to

where SEi is the standard error for the mean in stratum

2.6. Capture-recapture and Removal Methods


Capture-Recapture Method
Suppose the population is of size N, so that N in the number we wish to estimate. Suppose, M organisms were captured, marked (or tagged) and released back into the population. After some time that should be sufficient for organisms to mix, n organisms were captured, and m of these appeared to be were marked. The proportion of recaptured organisms is assumed to be the same as the proportion of marked organisms:

Population size can be found as: N = nM/m (the Lincoln Index). The following conditions should be met: 1. No immigration, emigration, births or deaths between the release and recapture times. 2. The probabilities of being caught are equal for all individuals (including marked ones). 3. Marks (or tags) are not lost and are always recognizable. The first two conditions are often non-realistic, and thus, several modifications of this method has been developed that loosen these conditions. Perhaps the most popular is the Jolly-Seber method which requires capturing and marking of animals at regular time intervals. Animals, marked and released each time, should have different marks so that it is possible to distinguish between individuals marked on different dates. Algorithm is given in Southwood (1978). Jolly-Seber method gives an estimate of population size on each specific date; the first condition can be violated. However, the second condition is still required. It is also possible to estimate mortality+emigration rate and birth+immigration rate on each specific day. These rates are assumed to be constent for all individuals (including marked individuals). There are numerous other models for capture-recapture experiments, which are specific for a particular population. For example, age structure of the population may be important, or some individuals may have higher probability to be caught than others. Another problem arises if the population has no boundaries. In this case, a grid of traps can be established, and only the central portion of the grid is used for analysis (because traps near the edges may be influenced by migration). The area covered by the grid should be much larger than the average distance of animal dispersal. Because the biology of different species is variable, it may be necessary to modify the capturerecapture model.

Removal method
Removal method is based on intensive trapping of animals in an isolated area. Migration is prevented by some kind of barriers. It is assumed that there are no births or natural deaths of organisms. The proportion of animals captured each day is the same. Therefore, population numbers and the number of captured individuals declines exponentially:

The model of removal: dN/dt = -aN, where a is the removal rate. The solutionof this differential equation is N = Noexp(-at), where No are initial numbers.

Then the number of animals captured per unit time is equal to A(t) = aNoexp(-at) Parameters a and No can be estimated using the non-linear regression. Pielou (p. 127) used a different method for estimating parameters which is based on the analysis of 2 first time intervals only. For example, if captures in the first 2 time intervals were 29 and 18 animals, then

a = (29-18)/29 = 0.38.
This model can be generalized assuming the recruitment of organisms (e.g., emergence of adult insects from the soil).

Lecture 5. Exponential and Logistic Growth


Course: Quantitative Population Ecology Dept. of Entomology, Virginia Tech, Blacksburg, VA Alexei Sharov Self-reproduction is the main feature of all living organisms. This is what distinguishes them from non-living things. Any model of population dynamics include reproduction. We will discuss two most important models of population growth based on reproduction of organisms: exponential and logistic models. Exponential and logistic models help to solve different kinds of problems in ecology, here are some examples: 1. How long will it take for a population to grow to a specific size? 2. What will be population size after n years (or generations)? 3. How long the population can survive at non-favorable conditions?

5.1 Exponential model 5.2. Logistic model 5.3. Discrete-time analogs of the exponential and logistic models 5.4. Questions and assignments

Additional external links: Logistic model (St. Olaf College, MN)

Exponential Model
Exponential model is associated with the name of Thomas Robert Malthus (1766-1834) who first realized that any species can potentially increase in numbers according to a geometric series. For example, if a species has non-overlapping populations (e.g., annual plants), and each organism produces R offspring, then, population numbers N in generations t=0,1,2,... is equal to:

When t is large, then this equation can be approximated by an exponential function:

There are 3 possible model outcomes:

1. Population exponentially declines (r < 0) 2. Population exponentially increases (r > 0)

3. Population does not change (r = 0) Parameter r is called:


Malthusian parameter Intrinsic rate of increase Instantaneous rate of natural increase Population growth rate

"Instantaneous rate of natural increase" and "Population growth rate" are generic terms because they do not imply any relationship to population density. It is better to use the term "Intrinsic rate of increase" for parameter r in the logistic model rather than in the exponential model because in the logistic model, r equals to the population growth rate at very low density (no environmental resistance). Assumptions of Exponential Model: 1. Continuous reproduction (e.g., no seasonality) 2. All organisms are identical (e.g., no age structure) 3. Environment is constant in space and time (e.g., resources are unlimited) However, exponential model is robust; it gives reasonable precision even if these conditions do not met. Organisms may differ in their age, survival, and mortality. But the population consists of a large number of organisms, and thus their birth and death rates are averaged. Parameter r in the exponential model can be interpreted as a difference between the birth (reproduction) rate and the death rate:

where b is the birth rate and m is the death rate. Birth rate is the number of offspring organisms produced per one existing organism in the population per unit time. Death rate is the probability of dying per one organism. The rate of population growth (r) is equal to birth rate (b) minus death rate (m). Applications of the exponential model

microbiology (growth of bacteria), conservation biology (restoration of disturbed populations), insect rearing (prediction of yield), plant or insect quarantine (population growth of introduced species), fishery (prediction of fish dynamics).

Logistic Model
Logistic Model
Logistic model was developed by Belgian mathematician Pierre Verhulst (1838) who suggested that the rate of population increase may be limited, i.e., it may depend on population density:

At low densities (N < < 0), the population growth rate is maximal and equals to ro. Parameter ro can be interpreted as population growth rate in the absence of intra-specific competition. Population growth rate declines with population numbers, N, and reaches 0 when N = K. Parameter K is the upper limit of population growth and it is called carrying capacity. It is usually interpreted as the amount of resources expressed in the number of organisms that can be supported by these resources. If population numbers exceed K, then population growth rate becomes negative and population numbers decline. The dynamics of the population is described by the differential equation:

which has the following solution:

Three possible model outcomes 1. Population increases and reaches a plateau (No < K). This is the logistic curve. 2. Population decreases and reaches a plateau (No > K) 3. Population does not change (No = K or No = 0) Logistic model has two equilibria: N = 0 and N = K. The first equilibrium is unstable because any small deviation from this equilibrium will lead to population growth. The second equilibrium is stable because after small disturbance the population returns to this equilibrium state.

Logistic model combines two ecological processes: reproduction and competition. Both processes depend on population numbers (or density). The rate of both processes corresponds to the mass-action law with coefficients: ro for reproduction and ro/K for competition. Interpretation of parameters of the logistic model Parameter ro is relatively easy to interpret: this is the maximum possible rate of population growth which is the net effect of reproduction and mortality (excluding density-dependent mortality). Slowly reproducing organisms (elephants) have low ro and rapidly reproducing organisms (majority of pest insects) have high ro. The problem with the logistic model is that parameter ro controls not only population growth rate, but population decline rate (at N > K) as well. Here biological sense becomes not clear. It is not obvious that organisms with a low reproduction rate should die at the same slow rate. If reproduction is slow and mortality is fast, then the logistic model will not work. Parameter K has biological meaning for populations with a strong interaction among individuals that controls their reproduction. For example, rodents have social structure that controls reproduction, birds have territoriality, plants compete for space and light. However, parameter K has no clear meaning for organisms whose population dynamics is determined by the balance of reproduction and mortality processes (e.g., most insect populations). In this case the equilibrium population density does not necessary correspond to the amount of resources; thus, the term "carrying capacity" becomes confusing. For example, equilibrium density may depend on mortality caused by natural enemies.

Discrete-time analogs of the exponential and logistic models


Exponential model analog:

where t is time measured in generations, and R is net reproduction rate. For monovoltine organisms (1 generation per year), R is the average number of offsprings per one parent. For example, in monovoltine insects with a 1:1 sex ratio, R = Fecundity/2. The dynamics of this model is similar to the continuous-time exponential model. Logistic model analog (Ricker):

The dynamics of this model is similar to the continuous-time logistic model if population growth rate is small (0 < ro < 0.5). However, if the population growth rate is high, then the model may exhibit more complex dynamics: damping oscillations, cycles, or chaos (see Lecture 9). An example of damping oscillations is shown below:

. Complex dynamics results from a time delay in feed-back mechanisms. There are no intermediate steps between time t and time t+1. Thus, overcompensation may occur if the population grows or declines too fast passing the equilibrium point. In the continuous-time logistic model, there is no delay because the rate of population growth is updated continuously. Thus, the population density cannot pass the equilibrium point.

Вам также может понравиться