You are on page 1of 37

Part I: Population Dynamics

1. A population is a group of individuals of one species living in the same area at the
same time. Populations are the biological units that natural selection acts upon and the
focus of most conservation and management actions.
2.  If a species is endangered, a common goal is to increase the size and range of its
population.
 If a species is invasive, the goal may be the reverse.
 If a species is harvested (i.e., hunted, fished, or collected for human use), we may
simply aim to maintain population size and distribution over time.
Successfully influencing these dynamics requires that we understand what makes the
size of a population rise and fall.
3. Five basic components underlie population dynamics: births, deaths, sex ratio
(relative number of males and females), age structure (the number of individuals of
each age), and dispersal (the movement of individuals from where they were born to
where they breed).

What are the factors that influence population change?

Birth and Death  The relative balance of births and deaths determines whether a
population increases (births outweigh deaths) or decreases
(deaths outweigh births) over time.
 Birth and death rates change for many reasons, and one key
parameter is the number of individuals in a population.
 The size of the population can change because of its density.
 Density-dependent factors include things like parasites,
starvation, predation, and disease.
 Weather events are another common cause of
density-independent mortality.
Sex Ratio  A small number of females is more likely to limit growth
because with fewer females, the population as a whole has a
lower reproductive potential.
 A limited number of males can find and fertilize many females,
so a reduced number of males will have less of an effect on the
population’s reproductive potential.
The importance of sex ratios is especially obvious among harvested
animals.
 Male ornaments (e.g., deer with big antlers, rhinos with large
horns, elephants with large tusks).
 Vulnerable females (e.g., sea turtles at nesting beaches).

Practices can skew sex ratio so greatly that females fail to reproduce
for lack of males, or mating is delayed.
Example Box 1. Sex ratios in Hawaiian monk seal populations (Monachus
schauinslandi)
The endangered Hawaiian monk seal is restricted to several
northwestern Hawaiian atolls. In at least two sites in the western part
of the island chain, high female mortality due to unknown factors
has led to adult males substantially outnumbering females. The
result is high competition for females, which leads to a breeding
phenomenon known as “mobbing” in which adult males kill female
seals in an attempt to mate, or injure them so severely that they fall
prey to sharks, thereby making the problem worse. A management
project introduced females into some populations, bringing overall
sex ratios into balance to reduce this behavior.

Age Structure  Age structure refers to the composition of older versus younger
individuals in a population.
 It’s important because age often relates to reproductive stage,
and the size of a breeding population is key to the entire
population’s growth or decline.
 Reproductive status depends more on size than age.

 The sexual development patterns of many species are fairly


continuous, and breeding begins at a specific age.
 Many short-lived plants and animals reach sexual maturity in
their first year and rarely deviate from this pattern.
 Longer-lived species (including species of albatrosses, tortoises,
and trees) often delay maturation for many years, using energy
early on for growth rather than reproduction.
 Those populations may consist mainly of non-breeding juveniles.
 They may also contain individuals that have lived long enough to
become infertile.
 Since such populations contain many non-breeders, the total
number of individuals can be misleading when it comes to
assessing the ability of a population to persist.
Example Box 2. The effect of poaching on the age structure of African
elephant (Loxodonta africana) populations.
The elephants in Kenya’s Samburu National Reserve are one of the
world’s most studied elephant populations. From research, we know
that elephant families revolve around experienced females that
influence where and when the family eats, which other elephants to
associate with, and how to respond to threats like lions and survive
events like droughts. Matriarchs support other females when they
give birth, and females in a family with an older matriarch
reproduce at a faster rate. This makes families larger and more
successful because pre-reproductive females assist with
“calf-sitting,” which increases calf survival. Unfortunately, poachers
target the largest, oldest individuals (both males and females), which
leaves increasing numbers of orphans. (Roughly one in five groups
in Samburu now contains no mature females.) Conservation
managers are actively studying the effects poaching has on age
structure and, in turn, on the population dynamics and survival of
the species.
Dispersal  Dispersal is defined as the movement of individuals from one
population (emigration) to another (immigration).
 Dispersal is referred to as migration, but technically speaking,
migration is an individual’s “round-trip” away from and back to
a particular place.
 In animals, dispersal usually occurs in juveniles (solitary
individuals just prior to first reproduction), and one sex usually
disperses more frequently than another.
 In plants, dispersal is a more diverse process, which can involve
pollen and seeds and is mediated most frequently by wind or
animals.
The Effect  It can help prevent inbreeding by ensuring that closely related
individuals separate instead of staying nearby and possibly
interbreeding.
 It also helps colonize previously uninhabited habitats.
 This is vital in highly changeable environments where the
alternative to dispersal is extinction.
 If only a few individuals disperse per year between populations
of thousands, the effect on overall population dynamics is small.
 Dispersal can be a key driver of population dynamics, as in the
case of isolated populations that persist in small numbers.
 As few as one or two migrants per year can have a large
influence on whether such populations persist.
Example Box 3. Importance of dispersal for small populations of bog turtles
(Glyptemys muhlenbergii)
The threatened bog turtle resides in small wetlands of the eastern
United States, and populations are exceptionally small, often with
fewer than 50 adults. A 10-year study detected no dispersal of adults
among neighboring wetlands as close as a half a kilometer apart,
leading ecologists to suspect that isolation might doom many of
these populations to imminent extinction. A follow-up genetic study
indicated that dispersal was rare only a single dispersal event every
few years, and only to wetlands within two kilometers—but was
happening. Modeling has established that even at this low rate,
dispersal can substantially increase the populations’ odds of
survival. Consequently, managers are prioritizing networks of
wetland habitats (i.e., wetlands connected and in close proximity to
other wetlands) for conservation and are considering options for
connecting more isolated populations to promote more dispersal.

How do these factors affect population dynamics?


A. Birth rate
Population dynamics are strongly influenced by the rate at which individuals
reproduce. Birth rate (sometimes referred to as “fecundity” or “natality”) refers
to number of young born or hatched per unit of time. The unit of time for
measuring birth rate is usually per year for organisms that live in seasonal
environments and breed once per year. A shorter time period, e.g., per month, is
useful for organisms that may breed several times a year.
Birth rates can be affected by factors that include individual condition
(starving or otherwise stressed individuals produce few if any offspring), age at
first reproduction (this ranges from weeks to decades after birth), and birth
interval (from once in a lifetime to every breeding season). All of these factors
combine to determine a population’s overall birth rate, which varies dramatically
among populations of different organisms.
Calculating birth rate requires two estimates: the number of the adult females
in the population and the number of offspring produced in a given unit of time.
For example, if a population of 10,000 female tortoises produces 120,000 eggs
per year, the birth rate would be 120,000/10,000 = 12 eggs/female.
B. Mortality and survival
Mortality rate (death rate) counterbalances birth rates. Mortality is the
fraction of individuals that die per unit of time (usually one year), determined by
tallying the total number of deaths and dividing it by the number of individuals
that were alive at the beginning of that time period. For example, if a pod of
whales gives birth to 100 calves and 95 are alive a year later, mortality rate for
the calves in their first year is 5/100, or 5%.
Survival is the converse of mortality and represents the fraction of
individuals that live through a particular time period. If you estimate either one of
these parameters, you can readily determine the other, as they must total 100% of
the population. For example, if the survival rate is 0.9, or 90% per year, then
mortality is 0.1, or 10% per year.
The age structure and sex ratio of populations are highly relevant to
calculating mortality and survival. Mortality rates typically vary by age and sex,
sometimes dramatically. If individuals of different ages or sexes die at different
rates, then mortality is said to be age- or sex-specific. In most species, the young
and old are more likely to die than individuals in midlife.
As with the elephants in Samburu (Box 2), younger individuals may be less
able to recognize or elude predators, find food, or locate shelter. Older individuals
may have accumulated injuries and parasite loads that likewise render them more
vulnerable. Consider sea turtles. Survival in early life is very low (perhaps one in
1,000 offspring successfully hatches and survives to reach maturity). In contrast,
adults survive at extremely high rates (> 98%/year), often living for decades.
Different species can have different patterns of age-dependent mortality and
survival, which has led to the evolution of very different reproductive “strategies”
(Box 4).
There are different patterns of sex-specific mortality and survival in different
species. Generally, the added burdens of producing, feeding, and protecting
young may predispose females to higher mortality rates. Conversely, males may
die more frequently because of the risks and exertions associated with territorial
behavior and courtship.

Box 4. Divergent reproductive strategies: frogs and elephants


Consider the rubber frog (Phrynomantis bifasciatus) and African elephants
(Loxodonta africana), two very different organisms that live in the savannas of
Gorongosa National Park. An individual rubber frog casts up to 1,500 eggs into
temporary pools formed by occasional rains and then leaves the eggs and
subsequent tadpoles to fend for themselves (no parental care). Only a few eggs may
Box 4. Divergent reproductive strategies: frogs and elephants
survive to adulthood, but that’s enough for populations to persist. In other words,
extremely high fecundity compensates for extremely low survival. In contrast,
African elephants produce one or two offspring every three to 10 years, each of
which is carefully cared for and is likely to survive to adulthood. Low fecundity is
balanced by high survival. The result—long-term population persistence—is the
same for both species but achieved by very different means.

C. Survivorship
Survivorship refers to the fraction of individuals that lives up to a certain age
(versus survival rate, which is the probability of an individual surviving a given
unit of time). Survivorship at birth is 100% (because all individuals successfully
born are born alive) but then declines in distinctly different patterns.
There are three basic patterns of survivorship in nature. If rates of
survivorship are high from birth through midlife and decline sharply in old age,
as they do among whales, bears, and elephants, those organisms have what’s
called a Type I survivorship curve. A high amount of parental care increases the
likelihood that their offspring will survive (Box 3). Other organisms, like birds
and most reptiles, die at a relatively constant rate across the lifespan. This steady
decline is called a Type II survivorship curve. At the other extreme are organisms
such as insects and many fish, which follow a Type III survivorship curve.
Mortality is high among their vulnerable young, which receive little or no
parental care, and low among older animals (as with sea turtles or insects).
D. Metapopulations
Another factor to consider in population modeling is dispersal rates, because
populations usually operate within systems called metapopulations. A
metapopulation is a group of subpopulations that are separated by inhospitable
terrain but linked when individuals occasionally move between them and
reproduce. Bog turtles, for example, inhabit patches of wetland separated by
forest, agricultural fields, roads, and housing developments. Dispersal between
these small subpopulations maintains the metapopulation. Even if some
subpopulations go extinct, the metapopulation will persist as long as dispersal
opportunities do too. That’s why maintaining connective habitat is critical to the
survival of many organisms.

Sources and sinks in a


metapopulation. Individuals
emigrate from source populations
(shaded) to sink populations
(non-shaded). Arrows indicate the
direction of movement.

How a metapopulation changes over time (metapopulation dynamics)


depends on the rate of dispersal between subpopulations. If dispersal is consistent
and frequent, the metapopulation is essentially a large, continuous population. At
the other extreme, subpopulations with no dispersal between them are essentially
individual small populations at higher risk of going extinct. More typically,
individuals emigrate from “source populations” to what are called “sink
populations” whose members do not emigrate. Source populations are usually
(but not always) larger than sink populations, and are higher density.
Wildlife managers need to be mindful, however, that sometimes the smallest
source populations are the most reproductively successful and actually sustain the
entire metapopulation. Sustaining metapopulations requires identifying those key
source subpopulations, maintaining large numbers of nearby subpopulations, and
protecting dispersal routes between them.

E. Conclusion
Population sizes change continuously because of births, deaths, and
movements of individuals. These processes are influenced by age structure and
sex ratios as well as by the population’s interactions with the environment and
with other species. Understanding these dynamics is fundamental to being able to
predict population growth, or to assess more broadly the structure of a community
or how an ecosystem functions.
Part II: Modeling Population Growth

That’s why scientists input their data into models. Mathematical models use
equations to represent natural processes, which is useful because we can’t always
conduct experiments on entire ecosystems. Models enable us to summarize and
interpret information collected from the field, and help us identify patterns in the
data and the mechanisms responsible for those patterns. Models also generate
testable predictions, such as whether the loss of an herbivore will cause a plant
population to rebound. Models are used throughout the study of ecology, but here
we will focus on modeling population growth, a practice increasingly in use by
wildlife managers.

A. Exponential growth
Although rare, exponential population growth does occur in nature when an
invasive species arrives in a place where it has no competition, for example, or in
the wake of a severe winter or drought, when the few surviving individuals enjoy
unfettered reproduction for a period of time. Notably, the human population has
grown more or less exponentially, although it is beginning to face environmental
constraints. Exponential population growth is a simple scenario that provides a
starting point for most population and community models.
Consider a simplified example. Suppose you are monitoring populations of
red flour beetles (Tribolium castaneum) and notice there is a 10% daily rate of
change in a population. So at the start of your observation you count 100 red flour
beetles (N = 100), and after one day there is a net addition of 10 individuals, for a
total of 110 beetles the following day. The next day will also have a 10% daily
rate of change for a population size of 110 beetles, so the population will grow by
11 individuals, leading to a population size of 121. Similarly, a 10% change in a
different population starting with 1,000 beetles leads to a net change of 100
individuals, for a total of 1,100 the following day. Though the growth rates are
the same, the first population increased by 10 on the first day and the other by
100—a subtle but important difference that highlights that a population increases
in proportion to its size (Figure 1).
Figure 1. Exponential growth chart. The red flour beetle (Tribolium castaneum) is a
pest species that eats flour and cereal products. They are common in homes and
grocery stores.

1. Population dynamics are strongly influenced by the rates at which individuals


reproduce and die. We can use the balance of the birth rate (b) and the death rate (d)
to calculate a population growth rate (r), expressed mathematically as:
r=b-d
2. The exponential growth of a population can be expressed as:
dN/dt = rN
dN/dt is simply the population change per unit time (per day in our example) and,
as we saw with the red beetles, is equal to the product of the growth rate (10% in our
example) and the initial population size (100).
3. The details of which we omit here, we can use an adaptation of this equation to
determine the new population size after any amount of time. The resulting equation is:
Nt = N0ert
Nt is the population size at some future time (t), e is the base of natural logarithms
(a natural constant, like pi, and approximately equal to 2.72), r is the growth rate, and
N0 is the size of the initial population.
This formulation provides a convenient way of predicting the future size of a
population. For example, what will be the likely size of a red flour beetle population
of 100 (N0 = 100) undergoing exponential growth at a rate of 10% (r = 0.1) for 10
days (t =10)?
Nt = N0ert
N(10) = (100)(2.72)(0.1)(10)
N(10) = (100)(2.72)(1)
N(10) = 272 red flour beetles after 10 days
B. Carrying capacity
After the drastic disturbance of the civil war in Gorongosa National Park and
the loss of many of its large herbivore competitors, the waterbuck population
(Kobus ellipsiprymnus) rose exponentially, At some point, however competition
for breeding sites, water, food, light, nutrients, or cover inevitably will reduce
birth rates to below the biological maximum. This type of competition in a
population affects the growth rate, especially when the population is at high
densities. Every area has a threshold, referred to as its “carrying capacity” and
often indicated by the symbol K (for the maximum number of individuals that it
can support). Incorporating this constraint into the exponential growth model
makes the model more useful and realistic.
C. Logistic growth
To model a density-dependent effect of population growth, we typically use
the so-called logistic growth equation, which involves adding a mechanism to the
exponential growth equation that slows population growth at high population
densities (r is not constant; it changes with the population size, N):
dN/dt = rN[1 - (N/K)]
As population size (N) nears carrying capacity (K), the term in brackets [1 -
(N/K)], which acts as a multiplier of population growth, gets smaller. The smaller
this factor, the larger its effect. In other words, as N approaches K, population
growth diminishes until, at N = K, it stops entirely. .

Figure 2. Carrying capacity. Logistic population growth produces an S-shaped


(sigmoid) curve in which the carrying capacity (K, dotted line) limits the number of
individuals in the population. When the population exceeds the carrying capacity, the
growth rate trends downwards. The population size may oscillate around the carrying
capacity, but unless something about the system changes (e.g., a disturbance or
community shift), the population size averages out to be a constant number (K).
For example, suppose the carrying capacity of a population is 100 individuals (K
= 100) and the population is seven (N = 7). The unused portion of the carrying
capacity is [1 - (7/100)] = 0.93. The relatively uncrowded population is growing at
93% of the exponential growth rate [r(0.93)N]. On the other hand, suppose the
population is close to its carrying capacity (N = 98), the unused carrying capacity of
the environment is small: [1 - (98/100)] = 0.02. In this case, the population grows at
only 2% of the exponential growth rate [r(0.02)N]. If the population should ever
exceed carrying capacity (N > K), the term in parentheses becomes negative, which
means that the growth rate is less than zero, and the population declines toward K.
This shows how density-dependent birth and death rates effectively curb exponential
population growth.
Estimating an area’s carrying capacity is challenging, however, and it can change
from one year to the next. We usually have to make many assumptions about how the
system works and narrow our focus to particular parts of the system to come up with
population estimates. As long as we’re aware of these limitations, the logistic growth
equation is a useful tool for modeling population growth (Box 1).

Box 1. Using logistic growth curves in fisheries management: optimal yields


By the early 1920s, commercial stocks of many fish species had started to decline in
response to overfishing. Fisheries managers turned to modeling population sizes to
understand how to avoid future population crashes. To estimate the maximum number
of fish they could harvest without causing a crash (maximum sustained yield), they
used the logistic growth model. A key property of this model is that the maximum
number of individuals added to a population per unit of time is at the inflection point
(half the carrying capacity, K/2) (see Figure 3). Below K/2, population growth rates
are high because the multiplier 1 - (N/K) is nearly equal to 1 and thus has no limiting
effect on population growth, but the population base upon which the growth rates is
acting is small. Thus, relatively fewer individuals are added to the population at < K/2
than at K/2. Above K/2, the population base that the growth rate can act upon is large,
but the rate of growth is diminished because the multiplier 1 - (N/K) approaches zero
and dramatically reduces population growth. Thus, K/2 is the population size at which
growth is highest. At that point, the maximum sustained annual harvest (the maximum
annual number of individuals that can be removed) is about (rK)/4.
Figure 3. Inflection point. An hypothetical point of maximum population growth
(the difference in population size from one year to the next) highlighted on the
logistic growth curve at the point K/2, or one half of carrying capacity.

Note that harvesting at the “sustainable maximum” does not guarantee that a
population will be stable. Natural disturbances can have long-lasting effects on
populations and should be taken into account in management decisions.

D. Conclusion
In practice, scientists continuously modify and refine basic models of
population growth, incorporating the variations that occur in natural populations
and their environments to better reflect and be applicable to populations scientist
uses the logistic growth model to compare different herbivore species in National
Park, but modifies it to include information about age structure and respective
differences in birth, maturation, and death rates. Still simplified, the model does a
decent job describing what is happening in a complex system. When a model fits,
we can use it to predict population sizes and growth by changing different
parameters (e.g., how long would it take for a population to rebound after a
disease outbreak?). When a model does not fit, we go back to our assumptions
and see how we can modify our estimates, or even the equations, to refine the
model. The simplest model that fits is usually the best model. But more complex
models, ones that allow us to add parameters such as birth rate, mortality, age
structure, and dispersal are necessary to refine our understanding of the complex
interactions that shape population dynamics. Conservation biologists, who are
tasked with recovering, reducing, or maintaining populations, can only succeed
by understanding these key and interconnected processes.
Communities and the Interaction Between Species

A. Introduction to community composition


Communities are assemblages of populations of different species that inhabit
the same place at the same time. Many factors affect the composition of a given
community, and no single factor explains its patterns. We’ll use the
recolonization of an area after a disturbance to consider what factors shape
community composition.

Figure 1. Processes and factors that affect species composition of a local


community. The species found in a regional species pool are determined by
historical events, physiological constraints, and evolutionary processes. The local
community composition is a nonrandom subset of the regional species pool based
on the interspecific interactions, habitat selection, and dispersal ability of the
species found within the regional species pool. Adapted from Morin (1999).

The community that colonizes a disturbed area will be a nonrandom subset of


the regional species pool. Dispersal ability is key. For example, a plant species
that disperses via water will be unable to colonize an area with no waterways.
Habitat selection can also shape the community. If there’s no appropriate habitat,
no caves for cave-dwelling frogs, for example, a species may arrive but be unable
to persist.

B. Interspecific interactions
Species interactions include competition, predation, herbivory, parasitism,
mutualism, commensalism, and amensalism (defined below). We’ll use symbols
to characterize the relationships + (positive), - (negative), and 0 (zero, no effect)
to indicate the net effect of one population on the population size of the other
species (Figure 2).
Figure 2. Interspecific interactions. The direct effects of some interspecific
interactions between two species (Species 1 and Species 2). For example, in
Predation, Species 1 has a direct negative effect on Species 2 (Species 2 dies or is
harmed), while Species 2 has a direct positive effect on Species 1 (Species 2
provides energy to the consumer).

C. Competition
When individuals of different species compete for a resource necessary for
their survival or growth, it is known as interspecific competition. It is a -/-
interaction--negative for both species. Not all resources are limited (in short
supply). For instance, unless the environment is anoxic, oxygen is rarely in short
supply and therefore organisms rarely compete for it.
In 1934, the Russian ecologist G. F. Gause conducted laboratory experiments
with two closely related species of single-celled organisms, Paramecium aurelia
and Paramecium caudatum, to study what happens when two species directly
compete for limited resources. When Gause grew the two species in separate test
tubes, adding a consistent amount of food and water every day, each population
grew rapidly until a certain point in time when the numbers leveled off (the
apparent carrying capacity). The population grew as large as the resource base
allowed (logistic growth). But when Gause grew the two species together, P.
caudatum numbers plummeted to close to zero within the tubes. Gause concluded
that, in the absence of disturbance, two species cannot coexist indefinitely when
competing for the same limited resources, especially if one species has a
competitive advantage. In this case, P. aurelia outcompeted P. cadatum for food
and used the resource more efficiently, allowing them to reproduce more rapidly.
This slight reproductive advantage will eventually lead to what is known as
competitive exclusion.
So, how can similar species coexist in a community? The answer lies in their
specific ecological niches. A niche is the sum of the biotic and abiotic resources
(including time and space) that a species uses, and how they are used.
Coexistence occurs when niches differ in one or more significant ways. For
example, a warbler’s niche consists of the time of year it nests, the time of the
day when it’s active, the size of the branches on which it perches, and the sizes
and types of insects and berries it eats along with other components.
How can two or more warbler species live on the same tree? In a study of
five species of Cape May warblers in 1958, ecologist Robert H. MacArthur noted
slight preferences in habitat, feeding position, behavior, and nesting date, all of
which reduced competition. Most likely the warblers’ niches differentiated over
time because of natural selection (niche differentiation), and resources were
divided effectively enough (resource partitioning) to prevent competitive
exclusion. For example, food (primarily insects on and around the tree) is
partitioned based on preferred feeding locations (e.g., myrtle warblers feed most
regularly at the base of the tree, while black-throated green warblers feed on outer
branches near the top of the tree).

D. Predation and herbivory


Predation is a +/- interaction between species in which the predator eats the
prey (a positive interaction) and the prey is killed (a negative interaction).
Herbivory differs from predation in that a herbivore often eats only part of the
plant, algae, or fungi it feeds on, and doesn’t necessarily kill it, but it’s still a +/-
interaction since the species being eaten is harmed. Both predators/herbivores and
prey/plants have adaptations that evolved by natural selection as a result of the
relationship. For example, adaptations such as teeth, claws, stingers, or poison
help predators catch and subdue prey. Similarly, many herbivores have
specialized teeth or digestive systems that are effective at breaking down the
cellulose found in many plant species. Some predators hunt collaboratively in
packs, and generally those that pursue prey are fast and agile, while camouflage
often assists those that lie in ambush.
In turn, prey can have defensive adaptations. Behavioral adaptations of
animals include hiding, fleeing, forming herds, and giving alarm calls. Just as
predators can use camouflage, some prey have cryptic coloration making
themselves more difficult to locate or capture. Other animals and plants have
mechanical or chemical defenses (think porcupines, rose thorns, skunks, and
poisonous mushrooms). When the interaction between predators/herbivores and
their prey/plants acts as a major selective force, it is an example of coevolution.
This is often referred to as an “arms race”: as predators/herbivores evolve better
weapons, prey/plants develop stronger defenses.
E. Parasitism
Parasitism is a relationship in which one organism (the parasite) gets its
nourishment from another organism (the host) and the host is harmed by the
interaction. Like predation, this is a +/- interaction, but unlike predator/prey
interactions, parasites and hosts live in symbiosis (close association). In some
cases parasites have complex life cycles involving multiple hosts and/or periods
of free living (with no host).
Parasites cause a direct or indirect negative effect on hosts by significantly
altering their survival, reproduction, and/or the density of their populations.
Additionally, some parasites manipulate the behavior of their hosts to increase the
probability of being transferred from one host to another, which may have an
additional negative effect on the host. Parasites are understudied, considering that
some scientists have estimated that one-third of the Earth’s biodiversity are
parasites, but ecologists are continually learning new things about parasite/host
interactions.

Box 1. Manipulative parasitic worm creates caterpillar mimics in snail eyestalks


Leucochloridium is a parasitic worm that has two hosts. It begins its life cycle as
an egg in bird droppings. If an amber snail ingests the droppings, the egg hatches
and develops into a sporocyst, which absorbs the nutrients and resources the snail
would normally use for reproduction. The sporocyst also places brood sacs filled
with larvae into the snail’s eyestalks, which causes them to swell and pulsate,
mimicking appetizing caterpillars upon which birds are likely to prey (Figure 4).
These snail hosts are mainly nocturnal, while the birds are diurnal, so the parasite
further manipulates the snail. The mechanism is still poorly understood, but it is
most likely a chemical cue that confuses the snail and causes it to be active during
the day and put itself in harm’s way. Fortunately for the snail, most birds simply
pluck off the infected eyestalk, which the snail can regenerate, but this requires
energy that would otherwise go toward reproduction. Once ingested,
Leucochloridium develops into a reproductive worm in the bird’s gut (where it
ingests digestive waste, affecting the bird only minimally). The life cycle is
completed when the worm releases eggs in the bird’s droppings.

F. Mutualism
Mutualism is a +/+ interaction between symbiotic partners that benefits both
organisms. For example, some lichens (composite organisms comprised of both a
fungus and an alga or cyanobacterium) are considered mutualistic symbionts with
the fungus receiving sugars from the alga, and in turn providing the algae a
protective habitat in which to grow. For some lichens, at least one species
(normally the fungus) has lost the ability to survive without its mutualistic partner.
This is known as obligate mutualism. Another example is the mutualism between
termites and the microorganisms in their digestive system: termites would not be
able to digest their food (wood) without the enzymes produced by the
microorganisms. In other cases, both species can survive alone, but the
relationship benefits both. For example, ants protect Acacia trees from browsing
herbivores by swarming and stinging the animals, and in return feed on the tree’s
nectar, but both can live without this relationship.
Mutualism may also involve adaptations that affect the survival, reproduction,
or dispersal of the other species. For example, most flowering plants produce
nectar or fruit that attracts animals, which pollinate them or disperse their fruit. In
turn, many animals have adaptations that help them to either find (with a
heightened sense of smell, for example) or consume the fruit or nectar.
Another examples of mutualistic relationships:
1. A goby fish and a pistol shrimp are unlikely companions. The blind pistol
shrimp creates a burrow that both it and the goby fish use for protection. In
exchange for this safe haven, the goby fish keeps an eye out for predators and
alerts the shrimp when danger is near.
2. Cleaner wrasses (smaller fish) are known to set up cleaning “stations” in
parts of a coral reef where fishes, sharks, and sea turtles present themselves
for cleaning. The cleaner wrasses clean parasites, dead and damaged scales,
and mucus from the organisms that visit these stations. The cleaner wrasses
get a meal, and the visitors’ health may improve.
3. In 1862, Darwin wrote that the Angraecum sesquipedale orchid measured
“eleven and a half inches long, with only the lower inch and a half filled with
sweet nectar. The Xanthopan morganii praedicta moth was discovered and
later confirmed as the mutualistic pollinator of the remarkable orchid.

G. Commensalism and ammensalism


Commensalism is a +/0 interaction that benefits one of the species and
neither harms nor helps the other. Ammensalism occurs when one species is
harmed, and the other is neither harmed nor helped (-/0). One example of a
commensal interaction might be “hitchhiking” species, like algae, living on the
protective shells of aquatic turtles. Although the algae appear to have little effect
on the turtles, these hitchhikers may affect its ability to move efficiently or to
absorb solar radiation, possibly reducing its energy and in turn its reproductive
success. Conversely, the algae may provide a camouflaging benefit to the turtles.
It’s hard to determine the net effect, but one species is likely benefiting more than
the other, the algae in this case. An example of ammensalism would be a giraffe
trampling the grass around a Acacia tree while grazing on its branches. The grass
doesn’t seem to affect the giraffe directly, but the animal’s hooves harm the grass.
H. Complexity of Interactions
The relationships between species in an ecosystem are less clear-cut than
these categories suggest. The direct and indirect effects of one species on another
can change over time and depend on webs of relationships with cascading effects.
For instance, we would expect Acacia trees to do better without giraffes grazing
on them, but ecologist Todd Palmer noticed the opposite. His research team
discovered that Acacia trees surrounded by fences were less healthy. Because of
reduced herbivory, trees produced less nectar and fewer thorns resources used by
ants.
The ant colonies shrank or abandoned the fenced-in trees, and the dominant
species changed from an ant species (Crematogaster mimosae) more reliant on
the trees’ thorns for nesting sites to an ant species (Crematogaster sjostedti)
reliant on stem cavities made by the larvae of long-horned beetles. While C.
mimosae ants prevent a beetle infestation, C. sjostedti ants somehow stimulate it,
and this infestation appears to harm the trees. In other words, over time, a change
in community composition (loss of large herbivores) affected the abundances and
distribution of the mutualistic ant species. For trees dominated by C. sjostedti ants,
the net effect shifted from positive (fences protected trees against herbivores) to
negative (fences encouraged beetles).

I. Conclusion
Communities are dynamic and interconnected assemblages of species.
Monitoring them after disturbances whether natural, like hurricanes, volcanic
eruptions, and landslides, or human-induced (anthropogenic), like deforestation
or war can reveal their complexities. Relationships between species and the
composition of communities are not always obvious, as in the indirect effect of
large herbivores on ant species and ants on Acacia trees in turn.
Community Structure and Dynamics

A. Trophic structure
The structure and dynamics of a community depend to a large extent on what
eats what. The trophic structure of a community can be described by the number
of different trophic levels it has: primary producers, primary consumers,
secondary consumers, and so forth. Another way to describe a community is
through food chains: linear representations of how the food energy that originates
in plants and other autotrophic organisms (primary producers) moves through
herbivores (primary consumers) to carnivores (secondary, tertiary, and quaternary
consumers).
A food chain with four trophic
levels (from left to right:
primary producer, primary
consumer, secondary consumer,
and tertiary consumer).

Food chains aren’t isolated. It was English biologist Charles Elton who
recognized in the 1920s that food webs connect them. Food webs can be complex.
Often it is easier to simplify these webs by either assigning species with similar
trophic relationships to broad functional groups (trophic levels), or isolating
portions that don’t interact much with the rest of the community.

An example of a food web


focusing on a subset of species
in a community. Note how
some species can be
categorized into more than one
trophic level. For example, the
mouse is both a primary
consumer (eating a primary
producer: wheat) and a

B. Bottom-up and top-down controls


Once we've assigned species to trophic levels, an approach to understanding
community structure is to model the relationships between the trophic levels. For
example, let’s consider three possible relationships between vegetation and
herbivores
Relationships between plants and herbivores. Three possible relationships
between vegetation (PP, for primary producers) and herbivores (PC, for primary
consumers). The horizontal arrows depict how the change of biomass in one
trophic level will lead to a change in the other trophic level. PP → PC shows the
unidirectional relationship of an increase in primary producer biomass leading to
an increase in primary consumer biomass. Conversely, PP ← PC represents a
unidirectional relationship of an increase in primary consumer biomass leading
to a decrease in primary producer biomass. When each trophic level responds to
changes of biomass of the other, this can be represented with a double-headed
arrow.
The first two relationships in the picture above underlie what we call the
bottom-up and top-down controls of communities, but we should keep in mind
that most communities combine both. The unidirectional influence seen in the PP
→ PC relationship shows a bottom-up model where lower trophic levels
influence higher trophic levels. For example, if fertilizer from a farm makes its
way into a body of water, it could stimulate the growth of phytoplankton, leading
to an increase of biomass in higher trophic level. In a bottom-up community, if
you add or remove predators, lower trophic levels shouldn’t be affected because
the effects are unidirectional.
The top-down model operates in the opposite direction: tertiary consumers
limit secondary consumers, secondary consumers limit primary consumers, and
primary consumers limit primary producers (Figure below). These are examples
of the direct effects of one species on another. Because food webs connect so
many species, the direct effect of one species on another may set off a chain of
events that affects other species in the community (indirect effects). In figure
below, this is depicted with the direct effect represented by the solid arrow and
the indirect effect represented by the dashed arrow: the number of tertiary
consumers indirectly affects the number of primary consumers. When a predator
initiates indirect effects on species lower in the food chain, it’s called a trophic
cascade.
Bottom-up (A) and top-down (B) controls of communities. In (A), the addition of
nutrients into the system (from agricultural runoff, for example) may increase the
number of producers, which in turn increases the abundances of each successive
trophic level. In (B), the number of predators in the highest trophic level has
direct effects on its prey, which has both cascading direct effects (solid arrows)
and indirect effects (dashed arrows) on the rest of the food chain.

C. Species with large effects


Certain species can affect entire communities because their ecological roles,
or niches, are pivotal. These type of species are known as keystone species.
Sometimes a species’ impact on a community not due to its trophic interactions,
but to its sheer abundance. Dominant species powerfully influence the occurrence
and distribution of other species because of their prevalence or collectively large
biomass. For example, mangrove trees and shrubs dominate and characterize the
coastlines of many tidal marshes in the tropics. These mangrove trees may have
become dominant in these areas because they are more able to sustain harsh
conditions and better exploit limited resources such as fresh water. These
dominant trees have abiotic effects on the ecosystem, including stabilizing soils
and shading the water, which in turn affects which other species live in those
areas. Dominant species might also become dominant because they succeed at
avoiding predation or disease. We see this when invasive species are introduced
to an area free of predators and disease agents, exploit this competitive advantage
over native species, and attain a high biomass.
Some organisms exert their influence not through their abundance but by
affecting the physical environment. They may create, the way corals build reefs
by secreting calcium carbonate; significantly alter; maintain; or destroy their
environments. Species that dramatically alter their physical environment are
called “ecosystem engineers” or “foundation species.”
D. Community stability and alternative stable states
A disturbance can change the community composition, species abundances,
and the relationships between species. If the disturbance is an event like a
landslide, the community may regain its original structure over time. This ability
is called community stability and is measured by observing how much a
community changes after a disturbance (resistance) and if, or how long, it takes to
return to its original state (resilience).
When a community does not bounce back, a new and stable community may
emerge that is resistant to further change. For example, if a beaver dams a stream
that runs through a wetland, most of the native vegetation may die. When
resources are reduced, the beaver will abandon its den, and when the dam gives
way, the water will drain, exposing sediment-rich soils. If the invasive common
reed (Phragmites spp.) is present and dominates this open niche, the community
will shift from a one characterized by diverse native vegetation to a stable state
dominated by dense common reed monocultures. Alternative stable states are not
necessarily bad for ecosystems, and later in the course we will discuss an ongoing
debate about the pros and cons of these “new” or “novel” ecosystems.

E. Conclusion
The effect of an invasive species on the forests of Guam offers a dramatic
example of these interconnections. In the mid-1940s, the brown tree snake (Boiga
irregularis), was unintentionally introduced to the island, where it encountered no
predators and ample food. To date, the snake has caused the local extinction of 12
native bird species, resulting in a trophic cascade in the food chain and indirect
effects in how species interact. Trees in the forest rely on birds to disperse their
seeds (known as a fruit-frugivore mutualism), and Guam has seen a 61% to 92%
decline in tree seedling recruitment. Given our biodiversity crisis, the need to
investigate the relationships between species is ongoing and urgent.
Flow of Energy and Matter in an Ecosystem

A. Conservation of energy and mass


Energy is the capacity to cause change, or to do work (to move matter against
an opposite force). In ecology, we focus principally on solar radiation, which is
used by autotrophs (primary producers such as plants and algae) to build organic
compounds. These compounds can then be passed to heterotrophs (consumers
such as herbivores and carnivores) in food, and dissipated as heat. As you may
remember from the first law of thermodynamics, energy cannot be created or
destroyed, only transferred or transformed.
Matter is anything that has mass and takes up space. In ecology, we focus on
common chemical elements that organisms use principally carbon, nitrogen, and
phosphorus. Chemical elements are mostly recycled, but smaller amounts also
flow in and out of an ecosystem’s boundaries over time. Autotrophic organisms
assimilate inorganic chemical elements from the air, soil, and/or water, and
transform them into organic form by incorporating them into their biomass. Some
of these chemical elements are added to the biomass of other species when
consumed via the food web. Detritivores (organisms like millipedes, burying
beetles, clams, and small crabs) and decomposers (e.g., fungi, some bacteria, and
some snails and insects) break down organic wastes (e.g., remains of dead
organisms, excrement, fallen foliage) and return them to the environment in
inorganic form. This picture below will shows how both energy and chemical
elements are transferred in ecosystems. Note that, unlike chemical elements,
energy cannot be recycled. Instead it enters, flows through, and exits an
ecosystem (entering most frequently as solar radiation and leaving as heat during
processes such as cellular respiration).

Energy (in orange) enters, flows


through, and exits and
ecosystem as heat. Chemical
nutrients (in blue) cycle within
an ecosystem--through the
trophic levels and eventually
back to the primary
producers--although, nutrients
can leave an ecosystem through
processes like erosion and
leaching (not shown). Primary
producers are autotrophs and
consumers are heterotrophs.
B. Primary productivity
An ecosystem’s primary production consists of the amount of energy that
autotrophs (producers) convert to chemical bonds in organic compounds during a
given time period. (The compounds are subsequently broken down to generate
energy-storing molecules like ATP.) The vast majority of this energy enters an
ecosystem through solar radiation and is processed via photosynthesis. (However,
in some ecosystems such as in some caves and deep sea hydrothermal vents,
autotrophic communities use chemical energy instead of sunlight, a process called
chemosynthesis.)
Ecologists identify two types of primary productivity: gross primary
productivity (GPP) and net primary productivity (NPP). GPP is the total amount
of energy converted via photosynthesis in a given area and period of time. GPP is
often expressed in Joules (J) or kiloJoules (kJ) per square meter per year. Not all
of this energy is stored, because the autotrophs use some for metabolic processes,
which are fueled by cellular respiration. The difference is the NPP: the energy
captured by primary producers in a given area and period of time (GPP), minus
the energy used by the primary producers for cellular respiration.
NPP = GPP - Respiration

Flow of energy from the Sun to GPP to NPP. Approximately 99 percent of the
solar energy that reaches plants (primary producers) is reflected or passes
through them without being absorbed. Only one percent is harnessed by
photosynthesis and contributes to GPP. The producer then uses approximately
60% of that energy for cellular respiration, leaving 40% available for other
organisms (NPP). This means that almost all life on the planet (excluding
organisms that get their energy from chemosynthesis and all the organisms that
depend on them) is supported by less than 1% of the solar energy that reaches
Earth.
C. Secondary production and efficiency
NPP is the energy that moves through the food web. Its movement and
distribution depends on how efficiently food energy is converted to biomass at
each step in each food chain within a food web. In most ecosystems, herbivores
only consume a small fraction of the NPP (available plant material). Moreover,
they digest only part of the plant material (assimilation) and excrete the rest as
waste. The amount of chemical energy in food that consumers convert to new
biomass (through growth or reproduction) during a given period is called
secondary production.
Hypothetical example. Say a giraffe eats an Acacia tree leaf containing 200 J
of energy: 50% of the energy is lost through its feces (100 J) and approximately
48.5% is used in cellular respiration (97 J). This leaves 1.5% (3 J) for growth
(addition of new biomass). Some detritivores may eat the feces, but most of this
energy flows out of the ecosystem as heat. This is why we say energy flows
through ecosystems rather than cycles within them. We can use the following
equation to measure how efficiently animals transform energy:
Production efficiency = Net secondary production x 100% / Assimilation of
primary production. In the case of our hypothetical giraffe, net secondary
production is 3 Joules and assimilation (the amount of energy not lost through
feces) is 100 Joules. Thus its production efficiency is 3 x 100% / 100 = 3%.
Animals’ efficiencies vary widely. For example, mammals (like giraffes) and
birds typically range between 1-3% because it takes a lot of energy to maintain a
high body temperature. In contrast, some insects and microorganisms can have
efficiencies that average 40% or more.
How do we measure energy flow within an ecosystem? Instead of calculating
the efficiencies of each organism within the food web, we can simplify the food
web by calculating the trophic efficiencies across it. As mentioned in the
Community Structure and Dynamics essay, the trophic structure of a community
can be described in terms of its trophic levels: primary producers, primary
consumers, secondary consumers and so forth. Although different ecosystems
have different trophic efficiencies, the average trophic efficiency is about 10%,
meaning that about 10% of the energy stored in the biomass at each trophic level
is converted to biomass at the next level. This means that 100 kg of vegetation
biomass can support about 10 kg of herbivore biomass, which can support about
1 kg of carnivore biomass. A useful way to visualize the difference in energy at
each level is to draw a trophic pyramid where the relative energy in each trophic
level is represented by different sized bars, stacked from primary producer up to
the highest trophic level in that community.
An idealized terrestrial trophic pyramid of net production. In this trophic pyramid,
trophic levels are visualized by the amount of net production (in Joules) for each
trophic level with the assumption that trophic efficiency between each level is 10%.
Note, the exact efficiency percentages vary among species and ecosystems.

D. Limits to production
Net primary productivity differs a great deal among ecosystems around the
world, in part because the amounts of sunlight and precipitation that reach
terrestrial ecosystems limit primary production (Figure 4). For example, little
sunlight reaches tundra regions at high latitudes—little energy enters the
system—so they are one of the least productive biomes. Tropical rainforests, on
the other hand, which are found around the equator and receive intense sunlight
year-round, are approximately 20 times more productive. On a more local scale,
production depends on the relative availability of key soil nutrients. Chemical
elements like nitrogen and phosphorus are the building blocks of organic
molecules that organisms need in order to grow and reproduce. As we’ll see in
the next section, animals can also moderate NPP.

E. The carbon cycle


The carbon cycle is closely tied to energy flow. Why? Photosynthesis
depends upon it. Plants, algae, and other photosynthetic organisms transform
solar energy into chemical energy. During this process, these organisms use solar
energy, water (H2O), and CO2 in the atmosphere (or dissolved in water) to create
carbon rich carbohydrates, such as glucose (C6H12O6), trapping the solar energy
in the bonds that hold these atoms together. Thus, you might see NPP expressed
in terms of carbon (grams of C per m2 per year -- or day in the case of Figure 4).
The chemical energy stored in an organism is passed through the food chain and
metabolized through cellular respiration. In the process, the waste product of CO2
is released back into the atmosphere or water, completing the cycle.
However, not all carbon cycles through organisms and directly back to the
atmosphere. A variety of processes operating at different rates and timescales
contribute to the global carbon cycle. For example, carbon can be sequestered for
varying lengths of time in soil, forests, the ocean, rocks and fossil fuels, and
released in a short period of time through natural and human-caused events.

Box 1. Wolves as part of the solution to global climate change?


Since the Industrial Revolution, humans have released large amounts of carbon
into the atmosphere through activities such as burning fossil fuels, deforestation,
industrialized farming and manufacturing. (This increase of carbon in the
atmosphere has not only altered the balance of the carbon cycle, but is responsible
for many other cascading effects that contribute to global climate change, such as
global warming, changes in water cycles that include droughts, floods, and
melting polar ice sheets. Sometimes these effects, in combination with other
factors, impact ecosystems in unforeseen ways--see Climate Stress and Coastal
Food Webs video). Scientists are studying how animals are mediating the carbon
cycle, and how our understanding of these processes can help us manage a system
more holistically. In a study published in 2014, Schmitz and others quantified the
animal contributions to regional carbon cycles and compare those numbers to the
anthropogenic fossil fuel emissions from the same region. The results were
striking. For example, experimental research in the boreal systems of Isle Royale,
Michigan, has shown that areas with high moose density can cause declines in
CO2 uptake and storage. Moose affect the system directly, by eating
photosynthetic plant tissues, and indirectly, by reducing tree growth. Shorter trees
may not form a closed canopy, which in turn increases evaporation and
temperature; these conditions are more conducive to fires, which release tons of
carbon (Figure 5). If extrapolated throughout Canada’s boreal forests, this overall
increase in CO2 in the atmosphere is equivalent to roughly 42 to 95% (depending
on moose density) of Canada’s total annual CO2 release from fossil-fuel
emissions. Therefore ecological managers can significantly reduce CO2 emissions
by establishing and promoting healthy wolf populations to keep moose
populations at lower densities. As we continue to address the challenges of
climate change, it is important for us to learn about how the integrity of food
webs can affect the carbon cycle, and then make informed management decisions
that reflect this knowledge.
F. Nitrogen and phosphorus cycles

The relationship between wolves, moose, trees, carbon, and global warming. This
causal loop diagram visually represents series of events and relationships that occur
in a boreal system when wolves are not present to keep moose populations at lower
densities. For the purposes of clarity, we have chosen to show only some of the
components and relationships within the system. For example, moose negatively
influence ecosystem carbon uptake/storage through browsing of trees and vegetation
(as shown) but they also increase CO2 emissions through priming soil microbial
activity through fecal deposition, and through direct respiration (not shown because
collectively, soil and moose respiration is orders of magnitude less than tree carbon
uptake/storage). Although not all feedbacks within this system are represented, we
have highlighted with double lined arrows example feedbacks that are expected to
occur. For example, up to a certain point, the reduction of trees/vegetation will lead
to increases in fire frequency or severity, which will further decrease tree/vegetation
biomass--a feedback loop that reinforces the effects of reduced carbon uptake/storage
and reduced canopy.

As mentioned earlier, NPP differs greatly among ecosystems, in part because the
amounts of sunlight and precipitation, but also because nitrogen or phosphorus
availability can limit production in ecosystems. Primary producers need nitrogen and
phosphorus for growth and reproduction because they are the building blocks of many
critical biochemicals such as ATP, amino acids, proteins, and nucleic acids (which
DNA is made of). Therefore, if there isn’t enough available nitrogen in an ecosystem
(in a boreal forest, for example) or phosphorus (in a freshwater lake, for example) for
all the organisms in need of these resources, growth will be limited.
Some organisms have developed adaptations to living in areas with low nutrient
availability (in part because of soil composition). For example, various bacteria and
fungi are able to use more forms of nitrogen (in both inorganic and organic chemical
compounds) than plants, which has led to the evolution of mutualistic relationships
between fungi/bacteria and plants. In the case of the bacteria Rhizobium and legumes,
the bacteria makes more nitrogen available to legumes than what’s available to them
in the soil, and in turn the legumes provide sugars and protection for the Rhizobium.
Our understanding of nutrient limitations and chemical cycles is important
for sustaining high agricultural yields. Agriculture removes nutrients from
ecosystems (plants uptake nutrients from the soil and then are harvested and
shipped outside of the ecosystem, and the nutrients do not cycle back into the soil
where the plants grew). To counteract this loss of nutrients, large supplements
(e.g., fertilizer) are added to the soil. Although these added nutrients lead to better
and consistent crop growth, the nutrient runoff from the fields can pollute aquatic
ecosystems (eutrophication), and stimulate excess algal growth .

G. Human activities dominate chemical cycles


As Earth’s population has grown, human activities have disrupted the basic
function of ecosystems: their species composition, energy flow, and chemical
cycling. Largely because of industrialized forms of agriculture and our growing
reliance on fossil fuels, human activities are not influencing most chemical cycles
more than they are influenced by natural processes. If we continue at this rate, we
may be altering Earth systems to the point that is unsustainable for humans.
The Stockholm Resilience Center has identified nine “planetary boundaries”
thresholds that could generate unacceptable change. This framework helps us
assess our risk of exceeding the capabilities of the global system. For example,
the biodiversity crisis we discussed at the beginning of the course is represented
by the red ‘genetic diversity’ zone in the figure under ‘biodiversity loss’. We are
also approaching or already beyond the sustainable change that the nitrogen and
phosphorus cycles are able to sustain.
Risk status for nine planetary boundaries. For each category of concern, the
threshold for a “safe-operating space for humanity” is represented by the planetary
boundary that lies at the intersection of the green and yellow zones. Categories
depicted in yellow and red have already exceeded the planetary boundary. Those in
yellow are in a zone of increasing risk and uncertainty, while those in red are in a
zone of high risk and uncertainty. Processes for which planetary boundaries have yet
to be quantified are represented by gray wedges; these are atmospheric aerosol
loading (i.e., particulates in the atmosphere), novel entities (i.e., new substances and
modified life forms that have the potential for unwanted geophysical and/or biological
effects), and functional diversity (i.e., the value, range, distribution, and relative
abundance of functions performed by organisms within an ecosystem).

Ecosystem Services, Functioning, and Resilience

A. Ecosystem function

The function of a particular ecosystem emerges from the collective activities


of the specific plants, animals, and microbes that inhabit it, and their effect on
physical and chemical conditions. A wetland, for example, produces certain kinds
of plant biomass, stores carbon, and cycles certain nutrients all differently than a
tropical forest does. These functions emerge from that particular system.
Ecosystem functions are varied and operate on vastly different scales, from a
fallen tree decomposing in the woods to extensive forests regulating weather and
climate.

B. Ecosystem resilience and the link to biodiversity

Some species and ecosystems are more resistant to disturbance, and


experience less change, while those that experience change but can recover
quickly are considered more resilient. For example, the Hudson River
experienced major changes in species composition following its invasion by
zebra mussels. While it was not resistant to this disturbance in the short term, as
you’ll see in the next essay, it has exhibited considerable resilience over time.

C. Ecosystem services

Regardless of who we are and where we live, functioning ecosystems


generate services from which every human being benefits. Ecosystem services are
the components of nature that humans consume, enjoy, or that contribute directly
or indirectly to human well-being. There are various ways to classify ecosystem
services. In 2000 the United Nations inaugurated the Millennium Ecosystem
Assessment (MEA), an ongoing assessment of the world’s ecosystems and the
services they provide.
The MEA grouped ecosystem services into four broad categories:

1. Provisioning services: goods that directly benefit people, such as fruits and
vegetables, and drinking water. Other types include timber, fisheries, fuelwood,
medicinal plants, natural building materials, fibers, and other forest products.
2. Regulating services: processes that moderate natural phenomena, including water
regulation for flood control, soil stabilization, removal of pollution from the air
and water, disease control, pollination, and regulation of climate through the
storage of carbon. These services help to make ecosystems sustainable, functional,
and resilient.
3. Cultural services: a non-material benefit that contributes to people’s intellectual,
cultural, and social development. This includes the role of ecosystems in people’s
history and education, and in their spiritual, religious, artistic and recreational
lives.
4. Supporting services: Supporting services have indirect or very long-term impacts
on people, but underlie other ecosystem services, particularly provisioning
services. Examples include the formation of soils, water and nutrient cycles, and
photosynthesis (i.e., oxygen production).

D. Accounting for ecosystem services

Regardless of how we define and classify ecosystem services, rapid


population growth and increasing demands for food, fresh water, timber, fiber,
and fuel have presented unprecedented threats. Human well-being and economic
development have greatly improved over the last two centuries, but at a cost:
many ecosystem services have degraded, and Earth’s biodiversity has suffered
substantial and largely irreversible losses. The MEA estimates that more than 60
percent of the world’s ecosystems are being used unsustainably.

Markets for ecosystem services like biodiversity and carbon sequestration are
beginning to emerge around the world, both voluntary and legally enforced.
Collectively known as Payments for Ecosystem Services (PES), these programs
offer incentives for landowners or stewards to restore or maintain ecosystems and
one or more of the services they provide.

The Anthropocene

A. Anthropogenic threats to biodiversity and ecosystems

Human activities are the leading threat to Earth’s biodiversity. Subsumed for
agriculture, tropical dry forests and grasslands have almost completely
disappeared. Many species are barely surviving on a fraction of their former
ranges, in increasingly fragmented landscapes. Dams disrupt freshwater
ecosystems, while overfishing, pollution, and habitat destruction threaten coastal
regions and the oceans. As we have seen, plants and animals transported around
the globe by humans can change entire ecosystems. No corner of the globe is
untouched: pollutants are carried tens of thousands of kilometers through the
atmosphere and contaminate both poles. We are changing Earth’s atmosphere
through the industrial release of carbon dioxide, which is dramatically altering the
climate.
B. Habitat loss and fragmentation

Habitat loss and fragmentation are related processes, typically occurring


simultaneously, and together are the primary threat to biodiversity globally.
Habitat loss occurs when an environment is modified to the extent that a
particular organism can no longer survive there, while ecosystem loss refers to
the disappearance of an entire ecosystem. Conservation measures often focus on
protecting threatened ecosystems, such as the acacia forests in the Senegal River
basin, or threatened species such as orangutans. Fragmentation is usually a
product of habitat loss and is best thought of as the subdivision of a formerly
contiguous landscape into smaller units.

In Sumatra and Borneo (home to the last populations of the critically


endangered orangutan), forests are converted primarily for agriculture (palm oil)
but also for mining, road development, and logging. When habitat is fragmented,
the species that depend on it not only lose part (or all) of their range, but also face
new risks, such as higher exposure to predators, invasive species, and being
hunted or collected by humans (e.g., young orangutans are targeted for the illegal
pet trade). Fragmentation also isolates populations, disrupting the dispersal and
migration of individual plants and animals (and their genetic material) across a
landscape. Ecosystems that are especially vulnerable to loss or fragmentation are
typically those in greatest demand for human use, for example for agriculture,
resource extraction, or other development (e.g., roads, housing). Conservation
goals often compete with these demands

C. Invasive species

Invasive species are arguably the second most important threat to


biodiversity globally, threatening individual species and even entire ecosystems.
The frequency and scope of this threat has increased along with the evolution of
transportation and commerce. While not all introduced species have a detrimental
effect, others can cause local or global extinctions and significantly disrupt
ecosystems, as we saw with the zebra mussels’ impact on the Hudson River. The
brown tree snake is directly credited with the local extinction of over half of
Guam’s native bird and lizard species and two (of only three) native bat species.
Chestnut blight, a fungus introduced from Japan, is estimated to have killed some
4 billion American chestnut trees in North America in the first half of the 20th
century. Rabbits have created major environmental disruption in Australia, where
they were introduced as a game animal, also costing millions per year in lost
agricultural production.

D. Overexploitation

Humans use wildlife and other natural resources in order to survive, but
overexploitation is a critical problem and plays a powerful role in biodiversity
loss. Overharvesting, unsustainable use, and illegal trade threaten not only the
continued survival of some plants and animals, but also that of ecosystems,
communities, and local economies that depend upon those species.
ECOLOGY
Name : Wulan Nur Fitriani

Topic : Adaptation of Organisms


NOTES

QUESTIONS NOTES
1. Define the following  Ecology is the study of how organisms interact with each other in their
word(s): biosphere, environment.
biogeographic regions,
Ecology consists of two components, the living and the non-living.
biomes, zones, habitat,

environment, ecological
niche.
Ecology can be divided into Aute-ecology and Synecology:
1. Aute-Ecology
2. What common features
do biomes have? Refers to the study of single species of plants and animals and focuses on the
population, distribution and seasons of abundance.

3. Give some examples of 2. Synecology


biomes? The study of the complete ecosystem and includes the study of many
different species of plants and animals interacting with each other and their
surroundings.
4. Find two countries that
are in the same line of Distribution and Adaptation of Organisms
latitude and compare the 1. Biosphere: The region region surrounding the earth that can support life.
types of vegetation or
biomes. 2. Biome: Geographical region that has it own unique sets of conditions
that support only particular flora and fauna. Biome is a place where
organisms adapt and live naturally.

 Biomes are divided into units called zones.


Ex: A closer look at the tropical rainforest would show distinct habitat zones
such as underground, understorey, canopy, emergent and so on, each
inhabited by specific plants and animals.
 Each particular zone (habitat) has its own set of physical conditions. In
other words, habitats are places with a particular set of conditions and an
adapted community of organisms.
Types of habitats include aquatic, for any water habitats, and terrestrial for
those in land.
*Aquatic : marine (sea), fresh water (lakes, rivers, ponds, streams).
*Terrestrial : boreal (on or in trees), ground, underground.
Terrestrial Biomes:
1. Alpine: an ecosystem that doesn’t contain trees due to its high altitude.
These biomes are found in mountainous regions across the globe. Their
elevation normally ranges between 10,000 feet (3,000 meters) and the
area where a mountain’s snow line begins. Animals include sheep,
mountain goats, alpaca, grasshopper.
2. Polar: located at the very top and very bottom of the Earth. They are
cold, windy and have a lot of snow and ice. It’s even too cold for trees to
grow. Polar habitats have tundra, which is ground that is nearly always
frozen. The very top of it will thaw in the summer months so grasses and
mosses can grow, but there isn’t enough room for tree roots in the non-
frozen soil. Because it is too cold for trees to grow in arctic habitats,
animals find other places to live such as holes in the ground, or in caves
made from snow.
3. Tundra: the coldest biome, treeless, receives only about 60 growing days
and low precipitation. Plants are mostly scrubs, mosses, lichens, sedges.
Animal include lemmings, mosquitos, migratory birds, fish. Ex: Canada,
Alaska, Iceland.
4. Coniferus forest: the areas are very cold or dry, sandwiched in between
the tundra to the north and the deciduous forest to the south. Coniferus
forest consist mostly of conifers, which are trees that grow needles
instead of leaves and cones instead of flowers.
5. Taiga (boreal forest) : long dry winter, cool, wet summer, 130 growing
day season, hosts coniferous trees, and low plants. Animalsspecies
include bear, moose, deer, lynx, hares.
6. Chaparral : scrubland and few trees,25-30 inch of rain annually, chiefly
in winter, dry summer mean dormancy for many plants, ex: California,
Mexico.
7. Desert : experience very high temperature, plants are adapt for low
rainfall, animal have nocturnal activity.
8. Grassland : dominated by grass, hot tropical savanna, moderate
precipitation, rich soils, a few trees grow along the rivers. Animals
species include deer, gazelles, birds, insects, and large predators.
9. Savanna: The savanna biome is often described as an area of grassland
with dispersed trees or clusters of trees. The lack of water makes the
savanna a difficult place for tall plants such as trees to grow. Grasses and
trees that grow in the savanna have adapted to life with little water and
hot temperatures. Animal species include elephants, lions, cheetah,
zebra, rhinos.
10. Tropical rainforest : contains the world’s greatest biodiversity, make up
one of earth’s largest biomes, known for its dense canopies, located
lowland near the equator, no cold and dry seasons, warm temperature,
200 inches of rain annually.
11. Tropical dry forests (TDFs) : characterized by a long dry season with
little rainfall. Although they once accounted for over 40% of all tropical
forests, they are highly threatened by climate change and agricultural
pressures.
12. Temperate Deciduous forest: hot summers and cold winters, go through
four seasons, the vegetation are broadleaf trees, shrubs, parennial herbs,
and mosses. Temperate deciduous forest are located in the mid-latitude
area which means that they are found between the polar regions and the
tropics.
Aquatic Biomes:

Intertidal zone Hugs the shoreline and is greatly affected by tides and
waves, the exoskeletons of shoreline crustaceans (such as
the shore crab, Carcinus maenas) are tough and protect
them from desiccation (drying out) and wave damage.
Another consequence of the pounding waves is that few
algae and plants establish themselves in the constantly
moving rocks, sand, or mud.
Neritic zone Extends to the continental shelf. Enough sunlight
penetrates for photosynthesis to take place, well-
oxygenated, low in pressure, and stable in temperature.
Seaweeds are often found here.
Pelagic zone Extends farther and experiences a mix of temperatures
due to current. Large fish and sea mammals ply this zone.
Nutrients are scarce and this is a relatively less productive
part of the marine biome. The majority of organisms in
the aphotic zone include sea cucumbers (phylum
Echinodermata) and other organisms that survive on the
nutrients contained in the dead bodies of organisms in the
photic zone.
Abyssal zone Represents the deepest ocean zone. High pressure, cold
temperatures, and low nutrients.There are a variety of
invertebrates and fishes found in this zone, but the abyssal
zone does not have plants because of the lack of light.
Cracks in the Earth’s crust called hydrothermal vents are
found primarily in the abyssal zone. Around these vents
chemosynthetic bacteria utilize the hydrogen sulfide and
other minerals emitted as an energy source and serve as
the base of the food chain found in the abyssal zone.
Benthic realm Deep region beyond the continental shelf. Here sea stars,
fish and sponges line the ocean floor.

 High biodiversity and the structures created by invertebrates that live in


warm, shallow waters within the photic zone of the ocean.
 The coral organisms (members of phylum Cnidaria) are colonies of
saltwater polyps that secrete a calcium carbonate skeleton. These
calcium-rich skeletons slowly accumulate, forming the underwater reef.
3.
 Biomes that occur where a source of fresh water, such as a river, meets
the ocean. Therefore, both fresh water and salt water are found in the
same vicinity; mixing results in a diluted (brackish) saltwater.
 Estuaries form protected areas where many of the young offspring of
crustaceans, mollusks, and fish begin their lives.
 Salinity is a very important factor that influences the organisms and the
adaptations of the organisms found in estuaries. The salinity of estuaries
varies and is based on the rate of flow of its freshwater sources.

Temperature is an important abiotic factor affecting living things found


in lakes and ponds.
Thermal stratification of lakes and ponds occurs when the upper layer of
water is warmed by the sun and does not mix with deeper, cooler water.
Light can penetrate within the photic zone of the lake or pond.
Organism in this area are phytoplankton providing the base of the food
web of lakes and ponds, then zooplankton consume these
phytoplankton.
5.
 Continuously moving bodies of water that carry large amounts of water
from the source, or headwater, to a lake or ocean.
 The source water is usually cold, low in nutrients, and clear.
 Plants and animals have adapted to this fast-moving water.

 Environment is the collective conditions and surrounding in which


organisms live.
 Conditions may include both living and non-living things, e.g., light,
temperature, water, depth, acidity, soil, and other organisms.
Ecoclogical niche:
 Refers to the interrelationship of a species with all the biotic and abiotic
factors affecting it.
 Generally considered to pertain to how an organism or a population
responds to, as well as alters, competition and the distribution of
resources.
 It particularly describes the relational position of an organism or a
population in a particular ecosystem.
 A niche may be influenced by biotic and abiotic factors of an ecosystem.
 However, the niche of a species in a particular ecosystem will help set
the features of its environment as these features will be crucial to its
survival.

Regions that share the same line of latitude tend to have similar biomes
because of the similar climatic conditions. For instance equatorial regions
such as the top part of South America, middle Africa, Northern Australia,
Indonesia, Malaysia and Papua New Guinea share the same line of latitude
and all have tropical rainforests. There are two major types of wet tropical
rainforests: equatorial evergreen rainforests and moist forests, which includes
monsoon forests and montane/cloud forests.

More information based of the topic:

Climate change and human activity pose dual threats to the long-term survival of the world’s coral reefs. As
global warming due to fossil fuel emissions raises ocean temperatures, coral reefs are suffering. The
excessive warmth causes the reefs to expel their symbiotic, food-producing algae, resulting in a phenomenon
known as bleaching. When bleaching occurs, the reefs lose much of their characteristic color as the algae and
the coral animals die if loss of the symbiotic zooxanthellae is prolonged.

Rising levels of atmospheric carbon dioxide further threaten the corals in other ways; as CO2 dissolves in
ocean waters, it lowers the pH and increases ocean acidity. As acidity increases, it interferes with the
calcification that normally occurs as coral animals build their calcium carbonate homes.

You might also like