Vous êtes sur la page 1sur 311

Plant Ecology .

"This page is Intentionally Left Blank"

j
j
j
j
j
j
j
j
j
j
j
j
j
j
j
j
j
j
j
Plant Ecology

Virendra Batra

Oxford Book Company


Jaipur, India
ISBN: 978-81-89473-65-5

First Published 2009

Oxford Book Company


267, lO-B-Scheme, Opp. Narayan Niwas,
Gopalpura By Pass Road., Jaipur-302018
Phone: 0141-2594705, Fax: 0141-2597527
e-mail: oxfordbook@sify.com
\\ eb~lte \" ~ w.abdpublisher.com

© Reserved

Typeset by:
Shivangi Computers
267, 10-B-Scheme, Opp. Narayan Niwas, •
Gopalpura By Pass Road., Jaipur-302018

Printed at Mehra Offset Press, D~lhi.

All Rights are Reserved. No part ofthis publication may be reproduced, stored in a
retrieval system. or transmitted, in any form or by any means, electronic,
mechanical, photocopying, recording, scanning or otherwise, without the prior
written permiss'ion of the copyright owner. Responsibility for the facts stated,
opinions expressed, conclusions reached and plagiarism, if any. m this volume is
entirely that of the Author, according to whom the matter encompassed in this
book has been originally created/edited and resemblance with any such
publication may be incidental. The Publisher bears no responsibility for them,
whatsoever.
Preface

Ecology involves the biological study of relationships of


organisms to their environment and to one another. Plants,
a primary unit of ecological processes, are involved in
constant interaction with the environment they inhabit,
depending upon natural resources like sunlight, air and
water for their nourishment, pr9viding oxygen and food for
other organisms, and being involved in cyclical processes
to maintain ecological and environment balance. As the
biosphere faces constant threats in the light of global
warming and pollution, it becomes even more pertinent
. upon us to comprehend the role plants play in sustaining
life through their ecological niche.
The present book has been aimed as an introductory
manual for botany students on the subject of plant ecology.
It seeks to delineate the concepts, principles, processes and
facts associated with the functioning of plants and their
interaction with other organisms, highlighting the ways in
which they lend support to maintain a stable ecological
system. In addition to stressing upon current discoveries
and breakthroughs in the field, the book provides space for
understanding the role of plants in applied ecology,
especially in the management and preservation of natural
resources and environments.

Virendra Batra
"This page is Intentionally Left Blank"
Contents

Preface v
1. Introduction 1
2. Biosphere and Plant Vegetation 15
3. Impact of Physical Environment on Plant Growth 37
4. Ecological Evolution of Plants 81
5. Ecology of Fungi 107
6. Ecology of Nonvascular Plants 125
7. Ecology of Seed Plants 153
8. Plant Community and Ecosystem Dynamics 179
9. Ecology of Weeds and Invasive Plants 197
10. Phage Ecology and Plants 227
11. Ecology of Plant Diseases 249
12. Plant Ecology and Climate Change 275
Bibliography 297
Index 301
"This page is Intentionally Left Blank"
1
Introduction

Plants are a major group of life forms and include familiar


organisms such as trees, herbs, bushes, grasses, vines, ferns,
mosses, and green algae. About 350,000 species of plants,
defined as seed plants, bryophytes, ferns and fern allies, are
estimated to exist currently. As of 2004, some 287,655
species had been identified, of which 258,650 are flowering
and 15,000 bryophytes. Green plants, sometimes called
metaphytes, obtain most of their energy from sunlight via
a process called photosynthesis.
Aristotle divided all living things between plants and
animals. In Linnaeus' system, these became the Kingdoms
Vegetabilia and Animalia. Since then, it has become clear
that the Plantae as originally defined included several
unrelated groups, and the fungi and several groups of algae
were removed to new kingdoms. However, these are still
often considered plants in many contexts, both technical
and popular. When the name Plantae or plants is applied
to a specific taxon, it is usually referring to one of three
concepts. From smallest to largest in inclusiveness, these
three groupings are:
Land plants, also known as Embryophyta or Metaphyta.
Green plants, also known as Viridiplantae, Viridiphyta
or Chlor-obionta - cOlnprise the above Embryophytes,
Charophyta (Le., primitive stoneworts), and
Chlorophyta (Le., green algae such as sea lettuce).
2 Plant Ecology

Archaeplastida, also known as Plantae sensu lato,


Plastida or Primoplantae, comprises the green plants
above, as well as Rhodophyta (red algae) and
Glaucophyta. As the broadest plant clade, this
comprises most of the eukaryotes that eons ago
acquired their chloroplasts directly by engulfing
cyanobacteria.
Informally, other creatures that carry out photosynthesis
are called plants as well, but they do not constitute a formal
taxon and represent species that are not closely related to
true plants. There are around 375,000 species of plants, and
each year more are found and described by science.
ALGAE

Most algae are no longer classified within the Kingdom


Plantae. The algae comprise several different groups of
organisms that produce energy through photosynthesis,
each of which arose independently from separate non-
photosynthetic ancestors. Most conspicuous among the
algae are the seaweeds, multicellular algae that may
roughly resemble terrestrial plants, but are classified
among the green, red, and brown algae. Each of these algal
groups also includes various microscopic and single-celled
organisms.
Only two groups of algae are considered close relatives
of land plants (embryophytes). The first of these groups is
the Charophyta (desmids and stoneworts), from which the
embryophytes developed. The sister group to the combined
embryophytes and charophytes is the other group of green
algae, and this more inclusive group is collectively referred
to as the green plants or Viridiplantae. The Kingdom
Plantae is often taken to mean this monophyletic grouping.
With a few exceptions among the green algae, all such
forms have cell walls containing cellulose, have
chloroplasts containing chlorophylls a and b, and store food
in the form of starch. They undergo closed mitosis without

/
Introduction 3

centrioles, and typically have mitochondria with flat


cristae. The chloroplasts of green plants are surrounded by
two membranes, suggesting they originated directly from
endosymbiotic cyanobacteria. The same is true of two
additional groups of algae: the Rhodophyta (red algae) and
Glaucophyta. All three groups together are generally
believed to have a common origin, and so are classified
together in the taxon Archaeplastida. In contrast, most
other algae have chloroplasts with three or four surrou-
nding membranes. They are not close relatives of the green
plants, presumably acquiring chloroplasts separately from
ingested or symbiotic green and red algae.
FUNGI

Fungi are no longer considered to be plants, though they


were previously included in the plant kingdom. Unlike
embryophytes and algae, fungi are not photosyntheoc, but
are saprotrophs: obtaining food by breaking down and
absorbing surrounding materials. Fungi are not plants, but
were historically treated as closely related to plants, and
were considered to be in the purview of botanists. It has
long been recognised that fungi are evolutionarily closer to
animals than to plants, but they still are covered more in
depth in introductory botany courses and are not
necessarily touched upon in introductory zoology courses.
Most fungi are formed by microscopic structures called
hyphae, which mayor may not be divided into cells but
contain eukaryotic nuclei. Fruiting bodies, of which
mushrooms are most familiar, are the reproductive
structures of fungi. They are not related to any of the
photosynthetic groups, but are close relatives of animals.
Therefore, the fungi are in a kingdom of their own.
PLANT DIVERSITY

About 350,000 species of plants, defined as seed plants,


bryophytes, ferns and fern allies, are estimated to exist
4 Plant Ecology

currently. As of 2004, some 287,655 species had been


identified, of which 258,650 are flowering plants, 16,000
bryophytes, 11,000 ferns and 8,000 green algae.
Embryophytes
Most familiar are the multicellular land plants, called
embryophytes. They include the vascular plants, plants
with full systems of leaves, stems, and roots. They also
include a few of their close relatives, often called
bryophytes, of which mosses and liverworts are the most
common.
All of these plants have eukaryotic cells with cell walls
composed of cellulose, and most obtain their energy
through photosynthesis, using light and carbon dioxide to
synthesize food. About three hundred plant species do not
photosynthesize but are parasites on other species of
photosynthetic plants. Plants are distinguished from green
algae, which represent a mode of photosynthetic life similar
to the kind modem plants are believed to have evolved
from, by having specialised reproductive organs protected
by non-reproductive tissues.
Bryophytes first appeared during the early Palaeozoic.
They can only survive where moisture is available for
significant periods, although some species are desiccation
tolerant. Most species of bryophyte remain small
throughout their life-cycle. This involves an alternation
between two generations: a haplOid stage, called the
gametophyte, and a diploid stage, called the sporophyte.
The sporophyte is short-lived and remains dependent on its
parent gametophyte.
Vascular plants first appeared during the Silurian
period, and by the Devonian had diversified and spread
into many different land environments. They have a
number of adaptations that allowed them to overcome the
limitations of the bryophytes. These include a cuticle
Introduction 5

resistant to desiccation, and vascular tissues which


transport water throughout the organism. In most the
sporophyte acts as a separate individual, while the \
gametophyte remains small.
The first primitive seed plants, Pteridosperms (seed
ferns) and Cordaites, both groups now extinct, appeared in
the late Devonian and diversified through the
Carboniferous, with further evolution through the Permian
and Triassic periods. In these the gametophyte stage is
completely reduced, and the sporophyte begins life inside
an enclosure called a seed, which develops while on the
parent plant, and with fertilisation by means of pollen
grains. Whereas other vascular plants, such as ferns,
reproduce by means of spores and so need moisture to
develop, some seed plants can survive and reproduce in
extremely arid conditions.
Early seed plants are referred to as gymnosperms
(naked seeds), as the seed embryo is not enclosed in a
protective structure at pollination, with the pollen landing
directly on the embryo. Four surviving groups remain
widespread now, particularly the conifers, whicJ:t are
dominant trees in several biomes. The angiosperlns,
comprising the flowering plants, were the last major group
of plants to appear, emerging from within the·
gymnosperms during the Jurassic and diversifying rapidly
during the Cretaceous. These differ in that the seed embryo
(angiosperm) is enclosed, so the pollen has to grow a tube
to penetrate the protective seed coat; they are the
predominant group of flora in most biomes today.

Fossils
Plant fossils include roots, wood, leaves.! seeds, fruit, pollen,
spores, phytoliths, and amber (the fossilised resin produced
by some plants). Fossil land plants are recorded in
terrestrial, lacustrine, fluvial and nearshore marine
sediments. Pollen, spores and algae (dinoflagellates and
6 Plant Ecology

acritarchs) are used for dating sedimentary rock sequences.


The remains of fossil plants are not as common as fossil
animals, although plant fossils are locally abundant in
many regions worldwide.
The earliest fossils clearly assignable to Kingdom
Plantae are fossil green algae from the Cambrian. These
fossils resemble calcified multicellular members of the
Dasycladales. Earlier Precambrian fossils are known which
resemble single-cell green algae, but definitive identity
with that group of algae is uncertain.
The oldest known trace fossils of embryophytes date
from the Ordovician, though such fossils are fragmentary.
By the Silurian, fossils of whole plants are preserved,
including the lycophyte Baragwanathia longifolia. From
the Devonian, detailed fossils of rhyniophytes have been
found. Early fossils of these ancient plants show the
individual cells within the plant tissue. The Devonian
period also saw the evolution of what many believe to be
the first modern tree, Archaeopteris. This fern-like tree
combined a woody trunk with the fronds of a fern, but
produced no seeds.
The Coal Measures are a major source of Palaeozoic
plant fossils, with many groups of plants in existence at this
time. The spoil heaps of coal mines are the best places to
~ollect; coal itself is the remains of fossilised plants, though
structural.detail of the plant fossils is rarely visible in coal.
kt the Fossil Forest at Victoria Park in Glasgow, Scotland,
the stumps of Lepidodendron trees are found in their
origina~ growth positions.
The fossilized remains of conifer and angiosperm roots,
stems and brancl1.es may be locally abundant in lake and
inshore sedimentary rocks from the Mesozoic and
Caenozoic eras. Sequoia and its allies, magnolia, oak, and
palms are often found.
Petrified wood is common in some parts of the world,
and is most frequently found in arid or desert areas where
Introduction 7

it is more readily exposed by erosion. Petrified wood is


often heavily silicified (the organic material replaced ~y
silicon dioxide), and the impregnated tissue is often
preserved in fine detail. Such specimens may be cut and
polished using lapidary equipment. Fossil forests of
petrified wood have been found in all continents.
Fossils of seed ferns such as Glossopteris are widely
distributed throughout several continents of the southern
hemisphere, a fact that gave support to Alfred Wegener's
early ideas regarding Continental drift theory.
PLANT GROWIH

Most of the solid material in a plant is taken from the


atmosphere. Through a process known as photosynthesis,
plants use the energy in sunlight to convert carbon dioxide
from the atmosphere into simple sugars. These sugars are
then used as building blocks and form the main structural
component of the plant. Plants rely on soil primarily for
support and water (in quantitative terms), but also obtain
nitrogen, phosphorus and other crucial elemental nutrients.
- For the majority of plants to grow successfully they also
require oxygen in the atmosphere and around their roots
for respiration. However, a few specialised vascular plants,
such as Mangroves, can grow with their roots in anoxic
conditions.

Factors Affecting Growth


The genotype of a plant affects its growth, for example
selected varieties of wheat grow rapidly, maturing within
110 days, whereas others, in the same environmental
conditions, grow more slowly and mature within 155 days.
Growth is also determined by environmental factors,
such as temperature, available water, available light, and
available nutrients in the soil. Any change in the
availability of these external conditions will be reflected in
8 Plant Ecology

the plants growth. Biotic factors (living organisms) also


affect plant growth.
flants compete with other plants for space, water, light
and nutrients. Plants can be so crowded that no single
individual makes normal growth.
Many plaIi.ts rely on birds and insects to effect
pollination.
Grazing animals may affect vegetation.
Soil fertility is influenced by the activity of bacteria and
fungi.
Bacteria, fungi, viruses, nematodes and insects can
parasitise plants.
Some plant roots require an association with fungi to
maintain normal activity (mycorrhizal association).
Simple plants like algae may have short life spans as
individuals, but their populations are commonly seasonal.
Other plants may be organised according to their seasonal
growth pattern:
Annual: live and reproduce within one growing season.
B,iennial: live for two growing seasons; usually
reproduce in second year.
Perennial: live for many growing seasons; continue to
reproduce once mature.
Among the vascular plants, perennials include both
evergreens that keep their leaves the entire year, and
deciduous plants which lose their leaves for some part of
it. In temperate and boreal climates, they generally lose
their leaves during the winter; many tropical plants lose
their leaves during the dry season.
The growth rate of plants is extremely variable. Some
mosses grow less than 0.001 mm/h, while most trees grow
0.025-0.250 mm/h. Some climbing species, such as kudzu,
which do not need to produce thick supportive tissue, may
grow up to 12.5 mm/h. Plants protect themselves from frost
Introduction 9

and dehydration stress with antifreeze proteins, heat-shock


proteins and sugars. LEA (Late Embryogenesis Abundant)
protein expression is induced by stresses and protects other
proteins from aggregation as a result of desiccation and
freezing.
Vascular plants differ from other plants in that they
transport nutrients between different parts through
specialised structures, called xylem and phloem. They also
have roots for taking up water and minerals. The xylem
moves water and minerals from the root to the rest of the
plant, and the phloem provides the roots with sugars and
other nutrient produced by the leaves.
ECOLOGICAL RELATIONSHIPS

The photosynthesis conducted by land plants and algae is


the ultimate source of energy and organic material in nearly
all ecosystems. Photosynthesis radically changed the
composition of the early Earth's atmosphere, which as a
result is now 21% oxygen. Animals and most other
organisms are aerobic, relying on oxygen; those that do not
are confined to relatively rare anaerobic environments.
Plants are the primary producers in most terrestrial
ecosystems and form the basis of the food web in those
ecosystems. Many animals rely on plants for shelter as well
as oxygen and food.
Land plants are key components of the water cycle and
several other biogeochemical cycles. Some plants have
coevolved with nitrogen fixing bacteria, making plants an
important part of the nitrogen cycle. Plant roots play an
essential role in soil development and prevention of soil
erosion.
Plants are distributed worldwide in varying numbers.
While they inhabit a multitude of biomes and ecoregions,
few can be found beyond the tundras at the northernmost
regions of continental shelves. At the southern extremes,
plants have adapted tenaciously to the prevailing
10 Plant Ecology

condi~ions. Plants are often the dominant physical and


structural component of habitats where they occur. Many
of the Earth's biomes are named for the type of vegetation
because plants are the dominant organisms in those biomes,
such as grasslands and forests.
Numerous animals have coevolved with plants. Many
animals pollinate flowers in exchange for food in the form
of pollen or nectar. Many animals disperse seeds, often by
eating fruit and passing the seeds in their feces.
Myrmecophytes are plants that have coevolved with ants.
The plant provides a home, and sometimes food, for the
ants. In exchange, the ants defend the plant from herbivores
and sometimes competing plants. Ant wastes provide
organic fertiliser.
The majority of plant species have various kinds of
fungi associated with their root systems in a kind of
mutualistic symbiosis known as mycorrhiza. The fungi help
the plants gain water and mineral nutrients from the soil,
while the plant gives the fungi carbohydrates
manufactured in photosynthesis. Some plants serve as
homes for endophytic fungi that protect the plant from
herbivores by producing toxins. The fungal endophyte,
Neotyphodium coenophialum, in tall fescue does
tremendous economic damage to the cattle industrys.
Various .forms of parasitism are also fairly common
among plants, from the semi-parasitic mistletoe that merely
takes some nutrients from its host, but still has
photosynthetic leaves, to the fully parasitic broomrape and
toothwort that acquire all their nutrients through
connections to the roots of other plants, and so have no
chlorophyll. Some plants, known as myco-heterotrophs,
parasitize mycorrhizal fungi, and hence act as epiparasites
on other plants.
Many plants are epiphytes, meaning they grow on other
plants, usually trees, without parasitizing them. Epiphytes
may indirectly harm their host plant by intercepting
Introduction 11

mineral nutrients and light that the host would otherwise


receive. The weight of large numbers of epiphytes may
break tree limbs. Many orchids, bromeliads, ferns and
mosses often grow as epiphytes. Bromeliad epiphytes
accumulate water in leafaxils to form phytotelmata,
complex aquatic food webs. A few plants are carnivorous,
such as the Venus flytrap and sundew. They trap small
animals and digest them to obtain mineral nutrients,
.;?Specially nitrogen.
IMPORTANCE OF PLANTS

The study of plant uses by people is termed economic


botany or ethnobotany. They are often used as synonyms
but some consider economic botany to focus mainly on uses
.)f modern cultivated plants, while ethnobotany studies
uses of indigenous plants by native peoples. Human
cultivation of plants is part of agriculture, which is the basis
.)f human civilisation. Plant agriculture is subdivided intu
agronomy, horticulture and forestry.
Virtually all human nutrition depends on land plants
directly or indirectly. Much of human nutrition depends on
cereals, especially maize or corn, wheat and rice or other
staple crops such as potato, cassava, and legumes. Other
parts from plants that are eaten include fruits, vegetables,
nuts, herbs, spices and edible flowers. Beverages from
plants include coffee, tea, wine, beer and alcohol. Sugar is
obtained mainly from sugar cane and sugar beet. Cooking
oils and margarine come from corn, soybean, canola,
safflower, sunflower, olive and others. Food additives
include gum arabic, guar gum, locust bean gum, starch and
pectin.
Wood is used for buildings, furniture, paper, cardboard,
musical instruments and sports equipment. Cloth is often
made from cotton, flax or synthetic fibers derived from
cellulose, such as rayon and acetate. Renewable fuels from
plants include firewood, peat and many other biofuels.
12 Plant Ecology

Medicines derived from plants include aspirin, taxol,


morphine, quinine, reserpine, colchicine, digitalis and
vincristine. There are hundreds of herbal supplements such
as ginkgo, Echinacea, feverfew, and Saint John's wort.
Pesticides derived from plants include nicotine, rotenone,
strychnine and pyrethrins. Drugs obtained from plants
include opium, cocaine and marijuana. Poisons from plants
include ricin, hemlock and curare. Plants are the source of
many natural products such as fibers, essential oils, dyes,
pigments, waxes, tannins, latex, gums, resins, alkaloids,
amber and cork. Products derived from plants include
soaps, paints, shampoos, perfumes, cosmetics, turpentine,
rubber, varnish, lubricants, linoleum, plastics, inks,
chewing gum and hemp rope. Plants are also a primary
source of basic chemicals for the industrial synthesis of a
vast array of organic chemicals. These chemicals are used
in a vast variety of studies and experiments.
Thousands of plant species are cultivated to beautify the
human environment as well as to provide shade, modify
temperatures, reduce windspeed, abate noise, provide
privacy and prevent soil erosion. People use cut flowers,
dried flowers and house plants indoors. Outdoors, they use
lawngrasses, shade trees, ornamental trees, shrubs, vines,
herbaceous perennials and bedding plants. Images of
plants are often used in art, architecture, humor, language
and photography and on textiles, money, stamps, flags and
coats of arms. Living plant art forms include topiary,
bonsai, ikebana and espalier.
Ornamental plants have sometimes changed the course
of history, as in tulipomania. Plants are the basis of a multi-
billion dollar per year tourism industry which includes
travel to arboretums, botanical gardens, historic gardens,
national parks, tulip festivals, rainforests, forests with
colorful autumn leaves and the National Cherry Blossom
Festival. Venus fly trap, sensitive plant and resurrection
plant are examples of plants sold as novelties.
Introduction 13

Tree rings are an important method of dating in


archeology and serve as a record of past climates. Basic
biological research has often been done with plants, such
as the pea plants used to derive Gregor Mendel's laws of
genetics. Space stations or space colonies may one day rely
on plants for life support. Plants are used as national and
state emblems, including state trees and state flowers.
Ancient trees are revered and many are famous. Numerous
world records are held by plants. Plants are often used as
memorials, gifts and to mark special occasions such as
births, deaths, weddings and holidays.
Plants figure prominently in mythology, religion and
literature. The field of ethnobotany studies plant use by
indigenous cultures which helps to conserve endangered
species as well as discover new medicinal plants.
Gardening is the most popular leisure activity in the U.S.
Working with plants or horticulture therapy is beneficial
for rehabilitating people with disabilities. Certain plants
contain psychotropic chemicals which are extracted and'
ingested, including tobacco, cannabis and opium.
Weeds are plants that grow where people do not want
them. People have spread plants beyond their native ranges
and some of these introduced plants become invasive,
damaging existing ecosystems by displacing native species.
Invasive plants cause billions of dollars in crop losses
annually by displacing crop plants, they increase the cost
of production and the use of chemical means to control
them affects the environment.
Plants may cause harm to people. Plants that produce
windblown pollen invoke allergic reactions in people who
suffer from hay fever. A wide variety of plants are
poisonous. Several plants cause skin irritations when
touched, such as poison ivy. Certain plants contain
psychotropic chemicals, which are extracted and ingested
or smoked, including tobacco, cannabis (marijuana),
cocaine and opium, causing damage to health or even
14 Plant Ecology

death. Both illegal and legal drugs derived from plants have
negative effects on the economy, affecting worker
productivity and law enforcement costs. Some plants cause
allergic reactions in people and animals when ingested,
while other plants cause food intolerances that negatively
affect health.
REFERENCES

Evans, L. T. (1998). Feeding the Ten Billion - Plants and Population Growth.
Cambridge University Press. Paperback, 247 pages.
Kenrick, Paul & Crane, Peter R. The Origin and Early Diversification of Land
Plants: A Cladistic Study. Washington, D. c.: Smithsonian
Institution Press. 1997.
Raven, Peter H., Evert, Ray F., & Eichhorn, Susan E. Biology of Plants (7th
_ ed.). New York: W. H. Freeman and Company. 2005.
Taylor, Thomas N. & Taylor, Edith L. The Biology und Evolution of Fossil
Plants. Englewood Cliffs, NJ: Prentice Hall. 1993.
2
Biosphere and Plant Vegetation

Plants occur in almost every conceivable habitat on Earth


-submerged on lake bottoms, exposed on wind-
sweptmountain tops, hidden within polar rocks, or perched
perilously on branches in the rain forest canopy. They can
be microscopic or enormous like sequoias and eucalypts
that may tower more than a hundred meters tall. Their
flowers may span nearly a meter across or extend the height
of a human, and be almost any color of the rainbow.
Plants comprise more than 99 percent of all the Earth's
living matter. The history of the biosphere is largely the
history of the origin and diversification of plants. Without
plants, conditions on Earth-including temperature, types
of rocks, the composition of the atmosphere, and even the
chemical composition of the oceans-would be vastly
different. While plant ecology is generally defined as lithe
study of relationships between plants and the
environment," plants do not, as this definition implies,
merely inhabit environments. Plants also modify the
environments, and they may even control them.
PLANTS AND BIOSPHERE

The word biosphere, which refers to that relatively thin


layer on the surface of the Earth within which life exists,
is now rather familiar to students. Yet the concept,
according to Hutchi.nson, was introduced into science
16 Plant Ecology

rather casually by the Australian geologist, Eduard Suess


in 1875. The idea was largely overlooked until the Russian
mineralogist, Vladimir Vernadsky, published La Biosphe're
in 1929. The word has now attained a general usage and
significance that Vernadsky probably could not have
imagined.
The biosphere has conditions that are rare in the
universe as a whole-liquid water in substantial quantities,
an external energy source, and temperatures at which there
are interfaces between solid, liquid, and gaseous forms of
water. Liquid water exists under a rather narrow range of
conditions of temperature and pressure. It was once
abundant present on Mars and may still occur beneath the
ice on Jupiter's moon Europa. New information is
continually emerging from interplanetary space probes. At
one end of the galactic temperature gradient there are
temperatures of trillions of degrees inside stars, and, at the
other end, there are conditions near absolute zero in the
vastness of space. Neither extreme provides the conditions
where biological chemistry, at least as humans understand
itl, can occur.
ENERGY FLOW

For life to exist, energy flow is required. Such a requirement


is met when a planet is situated near enough to a star for
sufficient energy released by solar fusion to pass the planet
before dissipating into outer space. This is the case for our
particular planet, situated near a star we know as the Sun.
While it is not known how often life occurs, it may not be
infrequent, given the enormous size of the universe our
own galaxy has some 100 billion suns, and there now
appears to be convincing evidence that some of these suns
have their own solar systems. This provides many
opportunities for other possible planets to be affected by
flowing energy. Proximity to a source of solar energy is
essential for life because that energy flow, by itself,
Biosphere and Plant Vegetation 17

organises matter. Life, at least as it is presently understood,


is matter that has been organised by energy flow. Morowitz
has examined the relationships among energy flow,
thermodynamics, and life asserting that in order to
properly understand life, one must look at the relationship
between physical laws and biological systems. He
demonstrates that flowing energy can create complexity out
of simplicity. Once the requirement for energy is met, life
then requires resources. This begs the question of what
those early resources might have been. One way to answer
such a question is to ask what conditions would have
existed in the early Earth's atmosphere before there was
life, since the early atmosphere would likely have been one
source of resources for the precursors of living cells.
Determining what the early atmosphere was like, however,
requires considerable detective work. It seems that this
atmosphere would have come, in part, from volcanic out-
gassings. For clues about its composition one can measure
the current composition of volcanic gases.
The early atmosphere would likely have been
composed of water, carbon dioxide, and sulfur. It was an
atmosphere rather different from that of today. Yet, some
billions of years later, these basic molecules remain as the
principal constituents of cellulose, the dominant structural
molecule of plants, and the most abundant molecule in the
biosphere.
Morowitz presents thermodynamic calculations
il'lustrating how energy flow stimulates chemical
interactions and creates molecules with higher potential
energy. Morowitz demonstrates mathematically that, with
energy flow and simple mixtures of gases, increasingly
complex molecules are formed. For example, a gaseous
mixture of ~arbon, hydrogen, nitrogen, and oxygen at 500
"C yields mo!:>'tly water and CO2 with smaller amounts of
other molecules, such as methane and ethane, which have
higher potential energy. The latter molecules are less likely
18 Plant Ecology

to form because they are larger and therefore more energy


is required to create them. As energy flows through the
molecular system, however, the energy distribution shifts
upward toward more and more complicated molecules.
Morowitz postulates that energy flow through the early
atmosphere yielded similar results: starting off with simple
low energy molecules such as water, CO2, and nitrogen,
more complex molecules were produced. The production of
molecules was driven by the external energy source, which
on Earth is the Sun. While some authors suggest that the
origin of life by such means contradicts the second law of
thermodynamics, what they fail to appreciate is that the
second law applies to closed systems. The biosphere is an
open system where, so long as energy flow occurs,
organisation will increase.
Another important physical condition of the early
environment on Earth was the abundance of water. It is not
surprising that water is still a major constituent of the
bodies of living organisms. Given the probable
temperatures on Earth at that time, water would be
evaporating from some areas, condensing in the
atmosphere, and then falling as rain. As it flowed back into
the sea, water would dissolve elements from the rocks-
elements that would rise in concentration as water
evaporated from the ocean again. These elements could
interact in solution, and concentrate in locations where
seawater was evaporating most rapidly.
Of course, while energy flow tends to produce larger
and more complex molecules, there is a natural
countervailing tendency-complex molecules will also
tend to fall apart into simpler molecules. But here is the
crucial point-some molecules will be more stable than
others. These stable ones will tend to persist and
accumulate. They will steadily become more common than
those other molecules that are unstable. It does not require
any great scientific insight to appreciate this, nor does it
Biosphere and Plant Vegetation 19

require us to imagine any sort of magical complexity or life


force-this process is simply a logical consequence of what
we mean by the terms "stable" and "unstable". Nothing
lasts forever.
Some things fall apart quickly, some things fall apart
slowly. So long as both kinds of things are being steadily
built by energy flowi, the long-lived ones will tend to
become more common than the short-lived ones. It is so
very simple-yet note that even at the chemical level, long
before there is anything that one might be tempted to call
life, there is a crude p~ocess of natural selection. Some
things are surviving longer than others, and hence are
becoming more common. Ammonia and methane are two
such molecules that likely accumulated in.the Earth's early
atmosphere.
Once a reservoir of larger and more stable molecules
forms, these molecules can in turn interact with each other,
yielding molecules with greater complexity and higher
levels of potential energy. Like the simpler molecules, these
more complex molecules will have varying degrees of
stability. Again, molecules that are unstable will fall apart
and those that are stable will accumulate. Imagine this
process continuing, with increasingly complex molecules
forming as a consequence of external energy flow. In this
simple scenario, there is ongoing natural selection
for stability and persistence, even at the molecular level
(Figure 1).
Such ideas are based upon thermodynamic calculations,
simple chemistry, and logic. Experimental work nicely
complements them. In an early experiment, Miller and Urey
set up a simple atmospheric system with a hydrological
cycle. Water was evaporated and then cooled and
condensed while sealed within glass tubes. Miller and Urey'
then let the hydrological cycle run, created electrical sparks
to simulate lightning, and found that primitive amino acids
formed.
20 Plant Ecology

, \
,I \

I
\
I
\

Figure 1. Solar energy creates high-energy molecules out of simpler low-


energy molecules.

This classic piece of work was done in the early 1950s, and
it is worth emphasizing that it was done by a graduate
student. Miller was fishing around for a research project to
do for graduate work and had already tried one project that
did not work. Then he and his advisor heard a seminar
about early conditions on Earth that stimulated them to try
their experiment. This single study led to a large series of
experiments wherein researchers created all manner of
artificial atmospheres and utilised different types of energy
flow to explore what kinds of molecules could be produced.
One could ask what factors might allow complex
molecules to further increase in stability and further
accumulate. Such factors would likely include:
Biosphere and Plant Vegetation 21

protective walls,
the direct use of sources of energy such as sunlight, a,nd
. the ability to form larger aggregations to buffer against
short-term periods of unsuitable conditions.
Consciousness would be another step, but this is not a step
that plants have taken. In The Selfish Gene, Dawkins argues
that consciousness can be thought of as the ability to
develop predictive models for future events. For example,
if an organism knows that certain conditions are likely to
bring winter, then it can store up fo('-:i. Such ideas will not
be explored further here, but Dawkins does raise other
issues, one of them being the way in which molecules that
copy themselves will proliferate.
Let us try to mentally reconstruct the circumstances on
Earth some 4 billion years ago. Pools of increasingly
complex molecules are accumulating as water evaporates
and energy flow stimulates chemical interactions.
Molecules that are stable are accumulating, those that are
unstable are falling apart. Now consider the possibility of
replication. Anymolecule that tends to create copies of itself
will accumulate more rapidly than other molecules.
Dawkins suggests that the occurrence of such
replicators was a critical event in the origin of life. Although
he uses the word "replication," "reproduction" is the
analogous biological term. From this perspective, then,
molecular stability is survival, and molecular replication is
reproduction. Thus, in a very basic and non-living
molecular system, it is possible to find the sorts of ecological
and evolutionary processes that occur in whole organisms.
Further, one can also find larger ecological processes such
as competition and predation. Margulis and Sagan describe
the circumstances on Earth at this time:
The ponds, lakes andwarmshallowseas of the early Earth,
exposed as they were to cycles of heat and cold, ultraviolet
light and darkness, evaporation and rain, harbored their
chemical ingredients through the gamut of energy states.
22 Plant Ecology

Combinations of molecules formed, broke up, and reformed,


their molecular links forged by the constant energy input of
sunlight. As the Earth's various microenvironments settled
into more stable states, more complex molecule chains
formed, and remained intact for longer periods. By
connecting to itself five times, for example, hydrogen
a
cyanide (HCN), molecule created in interstellar space and
a deadly poison to modem oxygen-breathing life, becomes
adenine (HsCsNs), the main part of one of the universal
nucleotides which make up DNA, RNA and ATP.

ORIGIN OF BIOSPHERE

Life began during the first billion years of an Earth history


which is 4.5 billion years old. The illustration depicts an
early Earth in which volcanoes, a gray, lifeless ocean, and
a turbulent atmosphere dominated the landscape. Vigorous
chemical activity is represented by the heavy clouds, which
were fed by volcanoes and penetrated both by lightning
discharges and solar radiation. The ocean received organic
matter from the land and the atmosphere, as well as from
infalling meteorites and comets. Here, substances such as
water, carbon dioxide, methane, and hydrogen cyanide
formed key molecules such as sugars, amino acids, and
nuc1eotides. Such molecules are the building blocks of
proteins and nucleic acids, compounds ubiquitous to all
living organisms.
A critical early triumph was the development of RNA
and DNA molecules, which directed biological processes
and preserved life's "operation instructions" for future
generations. RNA and DNA are depicted in the illustration,
first as fragmets and then as fully assembled helices. These
helices formed some of the living threads, as shown in the
illustration, however, other threads derived from planetary
processe~$uch as ocean chemistry a~d volcanic activity.
This evolving bundle of threads thus arose from a variety
of sources, illustrating that the origin of life was triggered
not only by special molecules such as RNA or DNA, but
Biosphere and Plant Vegetation 23

also by the chemical and physical properties of the Earth's


primitive environments.
Most of life's history involVled the biochemical
evolution of single-celled microorganisms. We find
individual fossilized microbes in rocks 3.5 billion years 'old,.
yet we can conclusively identify multicelled fossils only in
rocks younger than 1 billion years. The oldest microbial
communities often constructed layered mound-shaped
deposits called stromatolites, whose structures suggest that
those organisms sought light and were therefore
photosynthetic. These early stromatolites grew along
ancient seacoasts and endured harsh sunlight as well as
episodic ~etting and drying by tides. Thus it appears that,
even as early as 3.5 billion years ago, microorganisms had
become remarkably durable and sophisticated.
Many important events mark the interval between 1
and 3 billion years ago. Smaller volcanic terrains were
joined by larger, more stable granitic continents. Life
learned how to release oxygen from water, and it populated
the newly expanded continental shelf regions. The
illustration depicts these events, both in the abundant
mound-shaped stromatolites along the shoreline and in the
greater variety of filamentous and spherical microbes in the
foreground. Finally, between 1 and 2 billion years ago, the
eukaryotic cells with their complex system of organells and
membranes d~veloped and began to experiment with
multicelled body structures.
PLANT VEGETATION

Vegetation is a general term for the plant life of a region;


it refers to the ground cover life forms, structure, spatial
extent or any other specific botanical or geographic
characteristics. It is broader than the term flora which refers
exclusively to species composition. Perhaps the clos~s~
synonym is plant community, but vegetation can, and often
does, refer to a wider range of spatial scales. Primeval
24 Plant Ecology

redwood forests, coastal mangrove stands, sphagnum bogs,


desert soil crusts, roadside weed patches, wheat fields,
cultivated gardens and lawns; are all encompassed by the
term vegetation. Vegetation supports critical functions in
the biosphere, at all possible spatial scales.
First, vegetation regulates the flow of numerous
biogeochemical cycles, most critically those of water,
carbon, and nitrogen; it is also of great importance in local
and global energy balances. Such cycles are important not
only for global patterns of vegetation but also for those of
climate. Second, vegetation strongly affects soil
characteristics, including soil volume, chemistry and
texture, which feed back to affect various vegetational
characteristics, including productivity and structure. Third,
vegetation serves as wildlife habitat and the energy source
for the vast array of animal species on the planet.
Vegetation is also critically important to the world
economy, particularly in the use of fossil fuels as an energy
source, but also in the global production of food, wood, fuel
and other materials. Perhaps most importantly, and often
overlooked, global vegetation has been the primary source
of oxygen in the atmosphere, enabling the aerobic
metabolism systems to evolve and persist. Lastly,
vegetation is psychologically important to humans, who
evolved in direct contact with, and dependence on,
vegetation, for food, shelter, and medicine.

Vegetation Classification
Much of the work on vegetation classification comes from
European and North American ecologists, and they have
fundamentally different approaches. In North America,
vegetation types are based on a combination of the
following criteria: climate pattern, plant habit, phenology
and/ or growth form, and dominant species. In the current
US standard (adopted by the Federal Geographic Data
Committee (FGDC), and originally developed by UNESCO
Biosphere and Plant Vegetation 25

and The Nature Conservancy), the classification is


hierarchical and incorporates the non-floristic criteria into
the upper (most general) five levels and limited floristic
criteria only into the lower (most specific) two levels.
In Europe, classification often relies much more heavily,
sometimes entirely, on floristic (species) composition alone,
without explicit reference to climate, phenology or growth
forms. It often emphasizes indicator or diagnostic species
which separate one type from another.
In the FGDC standard, the hierarchy lev~ls, from most
general to most specific, are: system, class, subclass, group,
formation, alliance, and association. The lowest level, or
association, is thus the most precisely defined, and
incorporates the names of the dominant one to three
(usually two) species of the type.
Structure
A primary characteristic of vegetation is its three-
dimensional structure, sometimes referred to as its
physiognomy, or architecture. Most people have an
understanding of this idea through their familiarity with
terms like "jungle", "woods", "prairie" or "meadow"; these
terms conjure up a mental image of what such vegetation
looks like. So, meadows are grassy and open, tropical
rainforests are dense, tall and dark, savannas have trees
dotting a grass-covered landscape, etc.
Obviously, a forest has a very different structure than
a desert or a backyard lawn. Vegetation ecologists
discriminate structure at much more detailed levels than
this, but the principle is the same. Thus, different types of
forests can have very different structures; tropical
rainforests are very different from boreal conifer forests,
both of which differ from temperate deciduous forests.
Native grasslands in South Dakota, Arizona, and Indiana
are visibly different from each other, low elevation
chaparral differs from that at high elevations, etc.
26 Plant Ecology

Structure is determined by an interacting combination


of environmental and historical factors, and species
composition. It is characterised primarily by the horizontal
and vertical distributions of plant biomass, particularly
foliage biomass. Horizontal distributions refer to the
pattern of spacing of plant stems on the ground. Plants can
be very uniformly spaced, as in a tree plantation, or very
non-uniformly spaced, as in many forests in rocky,
mountainous terrain, where areas of high and low tree
density alternate dependi~g on the spatial pattern of soil
and climatic variables.
Three broad categories of spacing are recognised:
uniform, random and clumped. These correspond directly
to the expected variation in the distance between randomly
chosen locations and the closest plant to such locations.
Vertical distributions of biomass are determined by the
inherent productivity of an area, the height potential of the
dominant species, and the presence/absence of shade
tolerant species in the flora. Communities with high
productivities and in which at least one shade tolerant tree
species is present, have high levels of biomass because of
their high foliage densities throughout a large vertical
distance.
Although this discussion centers on biomass, it is
difficult to measure in practice. Ecologists thus often
measure a surrogate, plant cover, which is defined as the
percentage of the ground surface area that has plant
biomass vertically above it. If the vertical distribution of the
foliage is broken into defined height layers, cover can be
estimated for each layer, and the total cover value can
therefore be over 100; otherwise the values range from zero
to 100. The measure is designed to be a rough, but useful,
approximation of biomass.
In some vegetation types, the underground distribution
of biomass can also discriminate different types. Thus a
sod-forming grassland has a more continuous and
Biosphere and Plant Vegetation 27

connected root system, while a bunchgrass community's is


much less so, with more open spaces between plants
(though often not as drastic as the openings or spacings in
the above-ground part of the community, since root
systems are generally less constrained in their horizontal
growth patterns than are shoots). However, below-ground
architecture is so much more time-consuming to measure,
that vegetation structure is almost always described in
relationship to the above-ground parts of the community.

Dynamism in Vegetation
Like all biological systems,_plant communities are
temporally and spatially dynamic; they change at all
possible scales. Dynamism in vegetation is defined
primarily as changes in either or both of species
composition and vegetation structure.
Temporal Dynamics
Temporally, a large number of processes or events can
cause change, but for sake of simplicity they can be
categorised roughly as either abrupt or gradual. Abrupt
changes are generally referred to as disturbances; these
include things like wildfires, high winds, landslides, floods,
avalanches and the like. Their causes are usually external
to the community-they are natural processes occurring
independently of the natural processes of the community.
Such events can change vegetation structure and
species composition very quickly and for long time periods,
and they can do so over large areas. Very few ecosystems
are without some type of disturbance as a regular and
recurring part of the long term system dynamic. Fire and
wind disturbances are particularly common throughout
many vegetation types worldwide. Fire is particularly
potent because of its ability to destroy not only living
plants, but also the spores and seeds representing the
28 Plant Ecology

potential next generation, and because of fire's impact on


faunal populations and soil characteristics.
Temporal change at a slower pace is ubiquitous; it
Icomprises the field of ecological succession. Succession is
/ the relatively gradual change in structure and composition
J that arises as the vegetation itself modifies various
environmental variables, including light, water and
nutrient levels over time. These modifications change the
suite of species most adapted to grow, survive and
reproduce in an area, causing floristic changes. These
floristic changes contribute to structural changes that are
already inherent in plant growth even in the absence of
species changes, causing slow and broadly predictable
changes in the vegetation.
Succession can be interrupted at any time by
disturbance, setting the system either back to a previous
state, or off on another trajectory altogether. Because of this,
successional processes mayor may not lead to some static,
final state. Moreover, accurately predicting the
characteristics of such a state, even if it does arise, is not
always possible. In short, vegetative communities are
subject to many and unpredictable variables that limit
predictability.

Spatial Dynamics
As a general rule, the larger an area under consideration,
the more likely the vegetation will be heterogeneous across
it. Two main factors are at work. First, the temporal
dynamics of disturbance and succession are increasingly
unlikely to be in synchrony across any area as the size of
that area increases. That is, different areas will be at
different developmental stages due to different local
histories, particularly their times since last major
disturbance. This fact interacts with inherent
environmental variability, which is also a function of area.
Environmental variabi11ty constrains the suite of species
Biosphere and Plant Vegetation 29

that can occupy a given area, and the two factors together
interact to create a mosaic of vegetation conditions across
the landscape. Only in agricultural or horticultural systems
does vegetation ever approach perfect uniformity. In
natural systems, there is always heterogeneity, although its
scale and intensity will vary widely. A natural grassland
may seem relatively homogeneous when compared to the
same area of partially burned forest, but highly diverse and
heterogeneous when compared to the wheat field next to
it.

Global Vegetation Patterns


At regional and global scales there is predictability of
certain vegetation characteristics, especially physiognomic
ones, which are related to the predictability in certain
environmental characteristics. Much of the variation in
these global patterns is directly explainable by
corresponding patterns of temperature and precipitation.
These two factors are highly interactive in their effect on
plant growth, and their relationship to each other
throughout the year is critical. Such relationships are
shown graphically in climate diagrams. By graphing the
long term monthly averages. of the two variables against
each other, an idea is given as to whether or not
precipitation occurs during the warm season, when it is
most useful, and consequently the type of vegetation to be
expected. For example, two locations may have the same
average annual precipitation and temperature, but if the
relative timing of the precipitation and seasonal warmth are
very different, so will their vegetation structure and growth
and development processes be.

Scientific Study on Vegetation


Vegetation scientists study the causes of the patterns and
processes observed in vegetation at various scales of space
and time. Of partiCular interest and importance are
30 Plant Ecology

questions of the relative roles of climate, soil, topography,


and history on vegetation characteristics, including both
species composition and structure. Such questions are often
large scale, and so cannot easily be addressed by
experimentation in a meaningful way. Observational
studies supplemented by knowledge of botany,
paleobotany, ecology, soil science etc, are thus the rule in
vegetation science.
Vegetation science has its origins in the work of
botanists and/or naturalists of the 18th century, or earlier
in some cases. Many of these were world travelers on
exploratory voyages in the Age of Exploration, and their
work was a synthetic combination of botany and geography
that today we would call plant biogeography. Little was
known about worldwide floristic or vegetation patterns at
the time, and almost nothing about what determined them,
so much of the work involved collecting, categorising, and
naming plant specimens. Little or no theoretical work
occurred until the 19th century. The most productive of the
early naturalists was Alexander von' Humboldt, who
collected 60,000 plant specimens on a five year voyage to
South and Central America from 1799 to 1804.
Humboldt was one of the first to document the-
correspondence between climate and vegetation patterns,
in his massive, life-long work Voyage to the Equinoctial
Regions of the New Continent, which he wrote with Aime
Bonpland, the botanist who accompanied him. Humboldt
also described vegetation in physiogonmic terms rather
than just taxonomically. His work presaged intensive work
on environment-vegetation relationships that continues to
this day.
The beginnings of vegetation study as we know it today
began in Europe and Russia in the late 19th century,
particularly under Jozef Paczoski, a Pole, and Leonty
Ramensky, a Russian. Together they were much ahead of
their time, introducing or elaborating on almost all topics-
Biosphere and Plant Vege~tion 31

germane to the field today, well before they were so in the


west. These topics included plant community analysis, or
phytosociology, gradient analysis, succession, and topics in
plant ecophysiology and functional ecology. Due to
language and/or political reasons, much of their work was
unknown to much of the world, especially the English-
speaking world, until well into the 20th century.
In the United States, Henry Cowles and Frederic
Clements developed ideas of plant succession in the early
1900s. Clements is famous for his now discredited view of
the plant community as a superorganism. He argued that,
just as all organ systems in an individual must work
together for the body to function well, and which develop
in concert with each other as the individual matures, so the
individual species in a plant community also develop and
cooperate in a very tightly coordinated and synergistic way,
pushing the plant community towards a defined and
predictable end state. Although Clements did a great deal
of work on North American vegetation, his devotion to the
superorganism theory has hurt his reputation, as much
work since then by numerous researchers has shown the
idea to lack empirical support.
In contrast to Clements, several ecologists have since
demonstrated the validity of the individualistic hypothesis,
which asserts that plant communities are simply the sum
of a suite of species reacting individually to the
environment, and co-occurring in time and space.
Ramensky initiated this idea in Russia, and in 1926, Henry
pleason developed it in a paper in the, 'Pnited States.
Gleason's ideas were categorically rejected for 'many years,
so powerful was the influence of Clementsian ideas.
However, in the 1950s and 60s, a series of well-designed
studies by Robert Whittaker provided strong evidence for
Gleason's arguments, and against those of Clements.
Whittaker, considered one of the brightest and most
productive of American plant ecologists, was a developer
32 Plant Ecology

and proponent of gradient analysis, in which the


abundances of individual species are measured against
quantifiable environmental variables or their well-
correlated surrogates. In studies in three very different
montane ecosystems, Whittaker demonstrated strongly
that species respond primarily to the environment, and not
necessarily in any coordination with other, co-occurring
species. Other work, particularly in paleobotany, has lent
support to this view at larger temporal and spatial scales.
Since the 1960s, much research into vegetation has
revolved around topics in funCtional ecology. In a
functional framework, taxonomic botany is relatively less
important; investigations center around morphological,
anatomical and physiological classifications of species, with
the aim of predicting how particular groups thereof will
respond to various environmental variables. The
underlying basis for this approach is the observation that,
due to convergent evolution and adaptive radiation, there
is often not a strong relationship between phylogenetic
relatedness and environmental adaptations, especially at
higher levels of the phylogenetic taxonomy, and at large
spatial scales.
Functional classifications arguably began in the 1930s
with Raunkiaer's division of plants into groups based on
the location of their apical meristems relative to the ground
surface. Functional classifications are crucial in modeling
vegetation-environment interactions, which has been a
leading topic in vegetation ecology for the last 30 or more
years. Currently, there is a strong drive to model local,
regional and global vegetation changes in response to
global climate change, particularly changes in temperature,
precipitation and disturbance regimes. Functional
:lassifications such as the examples above, which attempt
to categorise all plant species into a very small number of
groups, are unlikely to be effective for the wide variety of
different modeling purposes that exist or will exist.
Biosphere and Plant Vegetation 33

It is generally recognised that simple, all-purpose


classifications will likely have to be replaced with more
detailed and function-specific classifications for the
modeling purpose at hand. This will require much better
understanding of the physiology, anatomy, and
developmental biology than currently exists, for a great
number of species, even if only the dominant species in
most vegetation types are considered.
ECOLOGICAL SUCCESSION

Ecological succession, a fundamental concept in ecology,


refers to more-or-Iess predictable and orderly changes in
the composition or structure of an ecological community.
Succession may be initiated either by formation of new,
unoccupied habitat or by some form of disturbance of an
existing community. Succession that begins in areas where
no soil is initially present is called primary succession,
whereas succession that begins in areas where soil is
already present is called secondary succession.
The trajectory of ecological change can be influenced by
site conditions, by the interactions of the species present,
and by more stochastic factors such as availability of
colonists or seeds, or weather conditions at the time of
disturbance. Some of these factors contribute to
predictability of successional dynamics; others add more
probabilistic elements. In general, communities in early
succession will be dominated by fast-growing, well-
dispersed species. As succession proceeds, these species
will tend to be replaced by more competitive species.
Trends in ecosystem and community properties in
succession have been suggested, but few appear to be
general. For example, species diversity almost necessarily
increases during early succession as new species arrive, but
may decline in later succession as competition eliminates
opportunistic species and leads to dominance by locally
superior competitors. Net Primary Productivity, biomass,
34 Plant Ecology

and trophic level properties all show-variable patterns over


succession, depending on the p~rticular system and site.
Ecological succession was formerly seen as having a
stable end-stage called the climax, sometimes referred to as
the 'potential vegetation' of a site, shaped primarily by the
local cli:i.nate. This idea has been largely abandoned by
modem ecologists in favor of nonequilibrium ideas of how
ecosystems function. Most natural ecosystems experience
disturbance at a rate that makes a "climax" community
unattainable. Climate change often occurs at a rate and
frequency sufficient to prevent arrival at a climax state.
Additions to available species pools through range
expansions and introductions can also continually reshape
communities.
Many species are specialised to exploit disturbances. In
forests of northeastern North America trees such as Betula
alleghaniensis and Prunus serotina are particularly well-
adapted to exploit large gaps in forest canopies, but are
intolerant of shade and are eventually replaced by other
species in the absence of disturbances that create such gaps.
The development of some ecosystem attributes, such as
pedogenesis and nutrient cycles, are both influenced by
community properties, and, in turn, influence further
community development. This process may occur only over
centuries or millennia. Coupled with the stochastic nature
of disturbance events and other long-term changes, such
dynamics make it doubtful whether the 'climax' concept
ever applies or is particularly useful in considering actual
vegetation.
The idea of ecological succession goes back to the 19th
Century. The French naturalist Adolphe Dureau de la Malle
was the first to make use of the word succession about the
vegetation development after forest clear-felling. In 1860
Henry David Thoreau read an address called "The
Succession of Forest Trees" in which he described
successIon in an Oak-Pine forest.
Biosphere and Plant Vegetation 35

Henry Chandler Cowles, at the University of Chicago,


developed a more formal concept of succession. Inspired by
the studies of Danish dunes done by Eugen Warming,
Cowles studied vegetation development sand dunes on the
shores of Lake Michigan. He recognised that vegetation on
sand-dunes of different ages might be interpreted as
different stages of a general trend of vegetation develop-
ment on dunes, and used his observations to propose a
particular sequence and process of primary succession. His
paper, "The ecological relations of the vegetation of the
sand dunes of Lake Michigan" in 1899 in the Botanical
Gazette is one of the classic publications in the history of
the field of ecology.
Understanding of succession was long dominated by
theories of Frederic Clements, a contemporary of Cowles,
who held that successional sequences of communities, were
highly predictable and culminated in a climatically
determined stable climax. Clements and his followers
developed a complex taxonomy of communities and
successional pathways.
A contrasting view, the Gleasonian framework, is more
complex, with three items: invoking interactions between
the physical environment, population-level interactions
between species, and disturbance regimes, in determining
the composition and spatial distribution of species. It differs
most fundamentally from the Clementsian view in
suggesting a much greater role of chance factors and in
denying the existence of coherent, sharply bounded
community types.
Gleason's ideas, first published in the early 20th
century, were more consistent with Cowles' thinking, and
were ultimately largely vindicated. However, they were
largely ignored from their publication until the 1960s.
About Frederic Clements' distinction between primary
succession and secondary succession, Cowles wrote:
36 Plant Ecology

This classification seems not to be of fundamental value,


since it separates such closely related phenomena as those
of erosion and deposition, and it places together such unlike
things as human agencies and the subsidence of land.

Beginning with the work of Robert Whittaker and John


Curtis in the 1950s and 1960s, models of succession have
gradually changed and become more complex. In modem
times, among North American ecologists, less stress has
been placed on the idea of a single climax vegetation, and
more study has gone into the role of contingency in the
actual development of communities.
REFERENCES

Archibold, O. W. Ecology of World Vegetation. New York: Springer


Publishing, 1994.
Barbour, M. G. and W. D. Billings (eds). North American Terrestrial
Vegetation. Cambridge: Cambridge University Press, 1999.
Barbour, M.G, J.H. Burk, and W.D. Pitts. "Terrestrial Plant Ecology".
Menlo Park: Benjamin Cummings, 1987.
Breckle, S-W. WaIter's Vegetation of the Earth. New York: Springer
Publishing, 2002.
Burrows, C. J. Processes of Vegetation Change. Oxford: Routledge Press,
1990.
Feldmeyer-Christie, (et.al). Modern Approaches In Vegetation Monitoring.
Budapest: Akademiai Kiado, 2005.
Gleason, H.A. 1926. The individualistic concept of the plant association.
Bulletin of the Torrey Botanical Club, 53:1-20.
Grime, J.P. 1987. Plant strategies and vegetation processes. Wiley
Interscience, New York NY.
Kabat, P., et al. (eds). Vegetation, Water, Humans and the Climate: A New
Perspective on an Interactive System. Heidelberg: Springer-Verlag
2004.
Macarthur, R.H. and E.O. Wilson. The theory of Island Biogeography.
Princeton: Princeton University Press. 1967
Mueller-Dombois, D., and H. Ellenberg. Aims and Methods of Vegetation
Ecology. The Blackburn Press, 2003.
Van Der Maarel, E. Vegetation Ecology. Oxford: Blackwell Publishers,
2004. .
Vankat, J. L. The Natural Vegetation of North America. Krieger Publishing
Co., 1992.
3
Impact of Physical Environment on
Plant Growth

This chapter describes the physical environment (soil, light,


temperature, humidity, wind) about plants and how the
physical environment affects the physiological status plants
and how plants affect their physical environment.
SOIL AND PLANTS

Soils physically support plants, and act as reser oirs for the
water and nutrients needed by plants. Soils are complex
mixtures of mineral particles of various shapes and sizes;
living and dead organic materials including micro-
organisms, roots, and plant and animal residues; air; and
water. In the soil, physical chemical, and biological
reactions occur constantly and are closely interrelated. The
physical form of the soil plays a large role in influencing
the nature of biological and chemical reactions. Optimum
plant growth depends as much on a favorable physical
environment as it does on what we call soil fertility.
The discussion of soil physical characteristics begins
with the sizes (texture) and arrangements (structure) of
individual soil particles. These two characteristics
intimately affect the pore space between the particles. The
pore space is important as the conveyor of water, dissolved
mineral nutrients, and air, as well as for providing space
in which rots can grow. Soil color is discussed because it
38 Plant Ecology

often provides information about the chemical makeup or


status of drainage in the soil. Finally, it is important to
consider the whole soil mass, and how it changes with
depth below the surface.

Soil Texture
Soil texture is a term which describes the mixture of
different sizes of mineral particles. The mineral particles,
originally from solid rock, assumed their present form
because of physical and chemical processes called
weathering. At some stage in the weathering process,
mineral particles became a favorable medium for plant
growth, that is, they were able to provide storage of water,
air, and mineral nutrients, as well as space in which roots
could grow. Organic matter then accumulated near the soil
surface due to the decomposition of plant residues.
Generally, organic matter further improved the properties
of the soil as an environment for plant growth.
Soil texture relates primarily to particles smaller than 2
millimeters in diameter-sand, silt, and clay-since these are
the particles most active in soil processes which support
plant growth. Coarser particles, gravel and stones, are
either inert or detrimental to plant cultivation.
Sand, the coarsest of the active particles, feels gritty
when rubbed. Sandy soils usually have rapid water
infiltration and good aeration but low water holding and
nutrient storage capacity. However, there is a considerable
range in these properties within the sand fraction.
Silt, the intermediate size, feels smooth when dry, and
slippery but not sticky when moist. Because the smaller
particle size promotes smaller pore spaces between
particles, silty soils have a slower water intake rate but a
higher water holding capacity than sandy soils. A few soils
are very high in silt. These are difficult for storage because
they often lack aggregation. This results in high density and
Impact of Physical Environment on Plant Growth 39

a pore size too small for suitable water percolation and


aeration. Nevertheless, silt is an essential component of the
medium textured, versatile soil called loam.
Clay, the finest size fraction, gives the soils a sticky or
plastic feel. Clay exhibits some unusual properties,
unexpected if it were merely composed of smaller particles
or the same minerals that make up sand and silt. Clay is
largely composed of a different set of minerals, called
secondary minerals. These are weathering products of the
primary minerals-quartz, feldspar, and mica-of which sand
and silt are largely composed.
One unusual property of clay is its attraction (called
adsorption) for positive ions, such as calcium, magnesium,
potassium, ammonium, and others. Because of this
adsorption, the clay in as quantities of the plantions. On the
other, negative plant nutrient ions such as nitrate,
phosphate, and sulfate are repelled by clay particles, and
can only be stored for plant use to the extent that they occur
dissolved in the water held in soil pores.
Clay has a very high affinity for water, partly because
of its small particle size and partly because the
aforementioned positive ions associated with clay also
attract water. Montmorillonite clay, the type found in many
soils, swells greatly when wetted, and shrinks-leaving wide
cracks when dry. While soils high in day are difficult to
manage because of their great strength and sticky nature,
an intermediate amount of clay in a soil improves its
capacity to hold water and plant nutrient ions. The swelling
and shrinking of clay also helps form favorable structure in
medium textured soils.
One useful and often used grouping of soil texhires
includes the following three categories:
Coarse-textured soils-Sands, loamy sands, and some
sandy loams.
Medium-textured soils-Loams, sandy loams, silt loams,
and some sandy day loams and clay loams.
40 Plant Ecology

Fine-textured soils ---Clays, sandy clays, silty clays, and


some sandy clay loams, silty clay loams, and clay
loams.
With experience, the texture of a soil can be felt and
determined fairly simply by rubbing moist soil between
thumb and forefinger, and noticing its characteristics-how
it ribbons or is pushed out into a thin strip-how it hangs
together, and how sticky, smooth, or gritty it is.

Soil Structure
Soil structure refers to the arrangement of soil particles.
Sand, silt, and clay seldom occur as separate units in the
soil; rather, they combine into aggregates held together by
small binding forces of clay and organic matter. The size
and form of aggregation is known as the structure of the
soil. Soil structure is one of the more important physical
characteristics of soil, yet perhaps the least understood.
Plant growth is strongly influenced by soil structure. Soil
structure affects movement of water, air, and roots through
the soil.
Soil structural aggregate may vary from a fraction of an
inch to several inches in diameter; may be approximately
spherical, elongated, or platelike; and may be held together
strongly or weakly.
A granular structure provides an ideal environment for
plant roots, and is particularly helpful for establishing
plants from seeds or transplants. The larger pores between
the granular aggregates are continuous, and roots may
penetrate them with ease.
Water drains readily through this soil, yet moisture is
held back sufficiently in the aggregates to supply root
needs. Granular structure occurs in loam soils and in some
clay soils near the surface. One of the good things about
clay is its promotion of granular structure in medium
textured soils. A greater organic matter content also results
Impact of Physical Environment on Plant Growth 41

in better granular structure of a soil. Sandy soils are low


in both organic matter and clay, and aggregation is very
weak to nonexistent. The structure is called single-grained;
such a soil drains well but doe not retain much moisture.
Single-grained soils require more frequent irrigation and
fertilisation for plant roots to thrive.
Prismatic and blocky structures most often occur as the
result of shrinking and cracking of clay loams and clay soil
layers (called horizons) upon drying. The large cracks that
are visible at the surface of dry clay soils may occasionally
extend to three feet or more in depth. The elongated chunks
of soil between these vertical cracks are called prisms.
The lower portions of the prisms often have horizontal
cracks intersecting the vertical ones so that more or less
equidimensional blocky structure results. Prismatic or
blocky aggregates may vary considerably in size but are
always coarser than those of granular structure. The
aggregates swell when wet and fit together so tightly that
water drains through them rather slowly. Plant roots may
follow cracks downward but do not usually penetrate to the
centers of prismatic or blocky aggregates. Thus, the roots
may not have access to a significant portion of the water
and nutrients in these soils.
Platy structure refers to the occurrence of thin layers of
soil stacked on top of one another. These most often occur
when silty soil materials are deposited in thin layers by
stream overflow. The discontinuities caused by this minute
layering may interrupt the movement of water, air, and
roots into the soil. Artificial platy structure may be caused
by repeated compression of soils in faim roadways.
Many medium textured soils do not have well defined
structural aggregates. This is true because of a much lower
organic matter content than most midwestern soils. If
particles are weakly bound together in the whole soil mass,
soils are said to have a massive structure. If open and
porous, these soils may still provide a favorable root
42 Plant Ecology

environment. Many massive soils, however, are dense and


nonporous providing only slow water and air movement.
Compact, massive layers occur naturally in the subsoils of
some old terrace soils, but farming activity has caused
similar compaction near the surface of many cultivated
soils that originally had granular or single-grained
structure.
Intensive cultivation usually results in some
breakdown of the natural soil structure. Forces holding soil
particles together in aggregates may not be strong enough
to resist the crushing effect of heavy equipment, or the
shearing effect resulting from working the soil at too high
a moisture content. Excessive traffic over the land results
in a compact soil mass in which large pores have collapsed
due to crushing of the granules.
In the absence of large pores, water penetration
becomes very slow. The small pores, still present, may fill
slowly with water after irrigation, and drain even more
slowly because water is held strongly by particle surfaces.
This has two serious effects. Water movement to lower
depths is very slow; and little or no airspace is left in the
compacted soil. Feeder roots of most crops will die if
deprived of air for only a few hours.
The more dense layers resulting from man-made soil
compaction usually show up within the surface foot of soil.
However, compression by tractor wheels and tillage
equipment may cause some compaction as deep as two feet
below the soil surface. Regardless of soil permeability
beneath the compact layer, water cannot percolate or
infiltrate faster than the limiting rate set by the compacted
layer. Compaction can develop in almost all soils, although
some soils seem more susceptible than others.

Preventing Soil Structure Breakdown


Although some breakdown of structure within the surface
font may be inevitable where land is intensively cultivated,
Impact of Physical Environment on Plant Growth 43

and understanding of soil texture and structure enables the


cultivator to apply solid cultural practices with a minimum
of structural breakdown. Structural breakdown is easier to
prevent than to cure. The following recommendations will
help prevent structural breakdown.
Plow and cultivate soil at an intermediate moishire
content-not to wet, not too dry.
It is especially important to avoid recompaction of
freshly plowed or loosened soil. The less tillage after
loosening the better.
Make tractor and implement tracks on the smallest
amount of land possible and use the same tracks for all
operations.
Harvest and spray when the soil is as dry as possible,
within the limitations of weather and timely schedule
of operation.

Rejuvenating Good Soil Structure


If compaction is severe, there is some possibility of
rejuvenating structure. The method used for such
rejuvenation will depend on the crop and the soil. The
factors favoring formation of granular structure are:
Wetting and drying of soils ca use swelling and
shrinking, resulting in improved aggregation.
Bacterial decomposition of plant residues produces
gums that help band soil particles together.
Planting fibrous rooted cover crops, particularly
grasses, helps to push soil particles together and makes
aggregates with continuous pore spaces between them.
The effect of swelling and shrinking on improving
granulation is particularly noticeable with medium and fine
textured soils in fall plowed fields left rough through
winter. To be most effective, the compacted layers should
be brought to the surface by deep plowing. If compacted
44 Plant Ecology

layers come up in large chunks or slabs, they will be able


to undergo swelling and shrinking in three dimensions due
to alternating wetting and drying. By spring the soil should
be in much better physical condition.
Incorporating crop residues should be included as a
management ,practice whenever possible in field or
vegetable crop production. Although it is difficult to build
up the percentage of soil organic matter because of rapid
decomposition in the hot regions, regular additions of crop
wastes can only have a beneficial effect on maintaining or
improving soil structure.
Cover crops or permanent sod in orchards and
vineyards can provide some structural improvement is
these plant roots can penetrate the compacted layer. Often,
however, the root penetration of the cover or sod crop itself
is restricted by the compaction, so it may be advisable to
break up the compaction mechanically before planting the
cover or sod.
Soil Color
Soil color is obvious and easily determine is one of the most
useful characteristics in class~fication and identification.
Determination of soil color with a Munsell color chart
provides a standard method of describing solid. Although
color has no direct influence on the functioning or
productivity of the soil, a great deal may be inferred about
a soil from its color. A few broad generalisations may be
made about soils of different colors.
Gray and brown soils form the largest group of soils.
They are moderately low in organic matter but include
some of the most productive alluvial soils. Gray soils of the
eastside, formed from granite alluvium, tend to be coarse
to medium textured. The brown soils of the westside
formed from sedimentary alluvium, tend to be medium to
fine textured. In all areas within each group, there is a wide
variation in productivity and other characteristics.
Impact of Physical Environment on Plant Growth 45

Black soils are relatively high in organic matter but the


amount may vary from less than 5 percent (mineral soils)
to more than 50 percent (peats and mucks). Black soils
formed under poorly drained conditions and are either
peaty or clayey in texture, but may with good management,
be highly productive for field and vegetable crops. In
upland or coastal areas, black soils with strong granular
structure have formed under native grassland, on fine
textured parent materials, and cool climates.
Red soils are generally older soils that have undergone
intensive weathering. In valleys, red soils occur on terraces
or bench lands much older than the soils of the recent
alluvial fans. These older soils often have restrictive clay
pans or hardpans in the subsoil. In the mountains, red soils
occur in the lower timber zone where a combination of high
winter rainfall and warm summer temperatures prevail.
Red soils are often deficient in phosphorous, zinc, and
sulfur, in addition to nitrogen.
White or light gray soils are usually sandy or calcareous
(contain lime). In sandy soils, look for possible
waterholding and nutrient problems; in calcareous soils,
iron deficiency may be a problem to some crops and
ornamental plants, but particularly to orchard crops. Blue
of blue-gray layers are usually found in poorly aerated
subsoils where organic matter is decomposing
anaerobically (without air). Often, such soils have a sewer-
like odor. These soils contain gases and dissolved materials
toxic to plant roots. Extensive aeration is necessary to
restore these soils to a condition suitable for plant growth.

Soil Depth
Soil depth is important to the management of plant growth.
The deeper the soil, the greater the totai water and nutrient
storage capacity available to plants. Soil depth can be
observed in roadcuts, stream banks, or by digging holes. A
soil auger is useful where exposed cuts are unavailable.
Plant Ecology

Holes are normally dug at least 5 feet deep unless hard rock
or hardpan is encountered. In making soil surveys, the soil
is investigated to a depth of 5 to 6 feet. ,In special cases,
investigation to a greater depth, possibly 10 to 20 feet, may
be desirable, particularly where salty layers or a fluctuating
water table may damage deep rooted crops.
Root and water penetration through a soil are altered
by layers having a distinctly different texture from the
layers above or below. If a sub-soil layer has a noticeable
increase in clay, water may accumulate above this layer,
and roots may be injured because of poor aeration. This
condition is often called waterlogging.
Very sandy or gravely layers can also interrupt the
normal downward penetration of roots or percolation of
water. For example, water does not drain freely from a
loamy layer into a sandy or gravely layer until the loamy
layer becomes saturated for some depth above the coarser
layer. When drainage has ceased, a saturated layer
remaining just above the textural change will have an
adverse effect on roots. The lingering saturated zone
remains because particle-to-particle flow of water is poor
from the loamy layer into sand or gravel.
Very dense, unfractured rocklike layers (hardpan)
sometimes occur in older alluvial soils on relatively flat
terraces. These cemented hardpans are impervious to both
water and roots. Winter rainfall accumulates above the
hardpan but cannot soak through it. Unless the hardpan is
shattered and drainage is improved, native grasses or crops
grow very poorly on the shallow root zone left as water
slowly evaporates from saturated soil.
Many of the soils in the uplands rest on hard rock. The
density, as well as the degree of fracture of the rock, is quite
variable. As a rule, the rock under the soil is more dense
in the lower foothills than in the mountainous areas. The
density and degree of fracture of the rock are important to
moisture storage, drainage, and runoff. A dense,
Impact of Physical Environment on Plant Growth 47

nonfractured, hard rock does not allow water to drain


readily from the soil above, nor does the rock store water.
A highly fractured rock stores water and allows soil
drainiige. In a soil underlain with fractured rocks, forest tree
roots may extract water to a depth of more than 20 feet.
The term effective root depth has been used to describe
that portion of the soil favorable for roots. In an alluvial soil,
with no noticeable stratification, effective root depth may
be more than five feet; in a claypan soil it may be as little
as 12 inches, or the depth of soil above the clay layer. Thus,
to determine soil depth, it is necessary to determine which
layers in the soil will be restrictive to root and water
penetration.
PLANT -WATER RELATIONSHIPS

Water is essential in the plant environment for a number


of reasons. Water transports minerals through the soil to the
roots where they are absorbed by the plant. Water is also
the principal medium for the chemical and biochemical
processes that support plant metabolism. Under pressure
within plant cells, water provides physical support for
plants. It also acts as a solvent for dissolved sugars and
minerals transported throughout the plant. In addition,
evaporation within intercellular spaces provides the
cooling mechanism that allows plants to maintain the
favorable temperatures necessary for metabolic processes.
Water is transported throughout plants almost
continuously. There is a constant movement of water from
the soil to the roots, from the roots into the various parts
of the plant, then into the leaves where it is released into
the ahnosphere as water vapor through the stomata (small
openings in the leaf surfaces). This process is called
transpiration. Combined with evaporation from the soil
and wet plant surfaces the total water loss to the
ahnosphere is called evapotranspiration.
48 Plant Ecology

One of the openings (stoma) is shown on the leaf cross


section in Figure 1.

~\
ls:v
VII..

Figure 1. Leaf cross section

Guard cells which are found on both sides of the stoma


control its opening and closing (Figure 2). Stomata can be
found on one (typically underside) or both sides of a leaf
depending Qll plant species.

Figure 2. Stoma and guard cells

\ Well-watered plants maintain their shape due to the


internal pressure in plant cells {turgor pressure}. This
pressure is also necessary for plant cell expansion and
Impact of Physical Environment on Plant Growth 49

consequently for plant growth. Loss of this pressure due to


insufficient water supply can be noticed as plant wilting.
The schematic effects of water stress on plant growth
are presented in Figure 3 . The major economic consequence
of insufficient water in agricultural crops is yield reduction.
When too little water is available in the root zone, the plant
will reduce the amount of water lost through transpiration
by partial or total stomatal closure. This results in decreased
photosynthesis since the CO2 required for this process
enters the plant through the stomata. Decreased
photosynthesis reduces biomass production and results in
decreased yields.

WA'TIIR DIFlCrr

LJJS.&~
10_=,_ ••

Figure 3. Schematic effects of water stress on plant growth

The role of soil in the soil-plant-etmosphere continuum is


unique. It has been demonstrated that soil is not essential
so Plant Ecology

for plant growth and indeed plants can be grown


hydroponically (in a liquid culture). However, usually
plants are grown in the soil and soil properties directly
affect the availability of water and nutrients to plants. Soil
water affects plant growth directly through its controlling
effect on plant water status and indirectly through its effect
on aeration, temperature, and nutrient transport, uptake
and transformation. The understanding of these properties
is helpful in good irrigation design and management.
The soil system is composed of three major components:
solid particles (minerals and organic matter), water with
various dissolved chemicals, and air. The percentage of
these components varies greatly with soil texture and
structure. An active root system requires a delicate balance
between the three soil components; but the balance
between the liquid and gas phases is most critical, since it
regulates root activity and plant growth process.
The amount of soil water is usually measured in terms
of water content as percentage by volume or mass, or as soil
water potential. Water content does not necessarily
describe the availability of the water to the plants, nor
indicates, how the water moves within the soil profile. The
only information provided by water content is the relative
amount of water in the soil.
Soil water potential, which is defined as the energy
required to remove water from the soil, does not directly
give the amount of water present in the root zone either.
Therefore, soil water content and soil water potential
should both be considered when dealing with plant growth
and irrigation. The soil water content and soil water
potential are related to each other, and the soil water
characteristic curve provides a graphical representation of
this r~lationship (Figure 4).
The nature of the soil characteristic curve depends on
the physical properties of the soil namely, texture and
structure. Soil texture refers to the distribution of the soil
Impact of Physical Environment on Plant Growth S1

particle sizes. The mineral particles of soil have a wide


range of sizes classified as sand, silt, and clay. The
proportion of each of these particles in the soil determines
its texture.

....
Figure 4. Graphical representation of soil water content-soil water potential
relationship

All mineral soils are classified depending on their texture.


Every soil can be placed in a particular soil group using a
soil textural triangle presented in Figure 5 . For example a
soil with 60% sand and 10% clay separates is classified as
a Sandy loam.
In addition almost all soils contain some organic
material, particularly in the top layer. This organic material,
together with the fine soil particles, contributes to aggregate
formation which results in the improvement of the soil
structure. Soil structure refers to the arrangement of soil
particles into certain patterns. The structural pattern, the
extent of aggregation, and the amount and nature of the
pore space describe the structure of the particular soil. No
S2 Plant Ecology

structure is usually present in sandy soils, however the


presence of the organic matter can improve tho Tu"'ter
holding capacity of the soil.
The size, shape, and arrangement of the soil particles
and the associated voids (pores) determine the ability of a
soil to retain water. It is important to realize that large pores
in the soil can conduct more water more rapidly than fine
pores. In addition, removing water from large pores is
easier and requires less energy than removing water from
smaller pores.

Figure 5. Soil textural triangle

Sandy soils consist mainly of large mineral particles with


very small percentages of clay, silt, and organic matter. In
sandy soils there are many more large pores than in clayey
soils. In addition the total ,,:olume of pores in sandy soils
Impact of Physical Environment on Plant Growth 53

is significantly smaller than in clayey soils (30 to 40% for


sandy soils as compared to 40 to 60% for clayey soils). As
a result, much less water can be stored in sandy soil than
in the clayey soil. It is also important to realize that a
significant number of the pores in sandy soils are large
enough to drain within the first 24 hours due to gravity and
this portion of water is lost from the system before plants
can use it.
To study soil-water-plant relationships it is convenient
to subdivide soil water into water available to the plant and
water unavailable to the plant. After the soil has been
saturated with water one can observe a vertical, downward
movement of water due to gravity. In Florida soils, this
drainage process happens quickly. Usually 24 hours is
sufficient to remove most of the gravitational water in
sandy soils. The exact time depends on the soil type; the
drainage of the gravitational water generally takes a little
longer for clayey soils. Most gravitational water moves out
of the root zone too rapidly to be used by the plants.
The remaining water is stored under tension in the
various size pores. The smaller the pore the greater the
tension and the more energy required to remove its water.
As a result plants have the ability to remove water only
from the certain size pores. The removal of water from very
small pores requires too much energy and consequently,
this water is not available to the plant. There is also some
water which is very closely bound to soil particles. This
water is called hygroscopic water. It is also very difficult to
remove, and is not available to the plants.
The range of water available to plants is between field
capacity (FC) and the permanent wilting point (PWP). The
soil is at field capacity when all the gravitational water has
been drained and a vertical movement of water due to
gravity is negligible. Further water removal for most of the
soils will require at least 7 kPa (7 cbars) tension. The
permanent wilting point is defined as the point where there
54 Plant Ecology

is no more water available to the plant. The permanent


wilting poi,ht depends on plant variety, but is usually
around 1500 kPa (15 bars). This means that
in order for plants to remove water from the soil, it must
exert a tension of more than 1500 kPa (IS,bars). This is the
limit for most plants and beyond .. t1lis
they experience
permanent wilting. It is easy to see that soils which hold
significant amounts of water at tension in the range plants
are able to exert (up to 1500 kPa (15 bars) of tension) will
provide better water supply for plant growth (Figure 6).

DMlIIAGII
WMIIR

DIW.....
W.,..

UJllLI.AIIY
WA1'D
IAVAILAIII..,

OAPlIJ.AIIIY
WAla
CAVAILAII....
NYOA08COPIC
WATER pwp
HYUOSCOPtC
WATUt

CLAY SAND

Figure 6. Water supply for plant growth

The pores in sandy soils are generally large and a


significant percentage drain under the force of gravity in
the first few hours after a rain. This water is lost from the
root zone to deep percolation. What remains is used very
quickly and the state of PWP can be reached in only a few
days.
Impact of Physical Environment on Plant Growth 55

PLANTS AND LIGHT

Plants have three basic responses or reactions to light. They


are photosynthesis, phototropism and photoperiodism.
Photosynthesis is, of course, the process on which all life
on earth depends. Radiant energy from the sun is converted
into chemical energy. The energy is stored in chemical
bonds in sugars like glucose and fructose.
Phototropism is the plant's movement in response to
light. All of us have seen the houseplant that leans toward
the window. That is phototropism. Growth hormones are
produced which cause the stem cells on the side away from
the light to multiply causing the stem to tilt. The leaves are
then closer to the light source and aligned to intercept the
most light.
The most interesting response is photoperiodism. This
is the plant's reaction to dark and is controlled by the
phytochrome pigment in the leaves. The pigment shifts
between two forms based on whether it'receives more red
or far red light. The reaction controls several different plant
reactions including seed germination, stem elongation,
dormancy, and blooming in day length sensitive plants.
Some seeds are also light sensitive. Germination is
controlled by the reaction in the phytochrome pigment.
Many lettuce varieties must have light to germinate.
Lettuce is packaged and distributed in foil packets to
prevent sprouting before planting. Most weed seeds are in
this category. Have you noticed how every time you till the
soil more weeds shoot up? Weed seeds lie dormant in the
soil for years waiting for you to stir up the soil so they get
enough light to germinate.
Phytochrome also controls lengthening or elongation of
stems. Leggy plants in low light are one example. The light
reaction in phytochrome also guides the germinating
seedling stem through the soil toward light. The last
photoperiod response is stimulation of dormancy. Several
56 Plant Ecology

things trigger dormancy, but a major one is the shortening


day length. This is critical when we move plants out of the
area where they evolved. For example, a sugar maple
grown in the north but from southern seed will not become
dormant early enough to escape winter cold injury.
Therefore it is important to buy perennial plants from seed
sources at similar latitudes to our own.

Photosynthesis
Photosynthesis is the conversion of light energy into
chemical energy by living organisms. The raw materials are
carbon dioxide and water; the energy source is sunlight;
and the end-products are oxygen and (energy rich)
carbohydrates, for example sucrose, glucose and starch.
This process is arguably the most important biochemical
pathway, since nearly all life on Earth either directly or
indirectly depends on it. It is a complex process occurring
in higher plants, phytoplankton, algae, as well as bacteria
such as cyanobacteria. Photosynthetic organisms are also
referred to as photoautotrophs.
Photosynthesis uses light energy and carbon dioxide to
make triose phospates (G3P). G3P is generally considered
the prime end-product of photosynthesis. It can be used as
an immediate food nutrient, or combined and rearranged
to form disaccharide sugars, such as sucrose, which can be
transported to other cells, or stored as insoluble
polysaccharides such as starch.
Photosynthesis occurs in two stages. In the first phase,
light-dependent reactions or photosynthetic reactions (also
called the Light reactions) capture the energy of light and
use it to make high-energy molecules. During the second
phase, the light-independent reactions (also called the
Calvin-Benson Cycle, and formerly. known as the Dark
Reactions) use the high-energy molecules to capture carbon
dioxide (C02 ) and make the precursors of carbohydrates.
Ini.pact of Physical Environment on Plant Growth 57

In the light reactions, one molecule of the pigment


chlorophyll absorbs one photon and loses one electron. This
electron is passed to a modified form of chlorophyll called
pheophytin, which passes the electron to a quinone
molecule, allowing the start of a flow of electrons down an
electron transport chain that leads to the ultimate reduction
of NADP into NADPH.
In addition, it serves to create a proton gradient across
the chloroplast membrane; its dissipation is used by ATP
Synthase for the concomitant synthesis 6f ATP. The
chlorophyll molecule regains the lost electron by taking one
from a water molecule through a process called photolysis,
that releases oxygen gas.
In the Light-independent or dark reactions the enzyme
RuBisCO captures CO2 from the atmosphere and in a
process that requires the newly-formed NADPH, called the
Calvin-Benson cycle releases three-carbon sugars, which
are later combined to form sucrose and starch.
Photosynthesis may simply be defined as the
conversion of light energy into chemical energy by living
organisms. It is affected by its surroundings and the rate of
photosynthesis is affected by the concentration of carbon
dioxide, the intensity of light, and the temperature.
Most plants are photoautotrophs, which means that
they are able to synthesize food directly from inorganic
compounds using light energy - for example from the sun,
instead of eating other organisms or relying on nutrients
derived from them. This is distinct from chemoautotrophs
that do not depend on light energy, but use energy from
inorganic compounds.

The energy for photosynthesis ultimately comes from


absorbed photons and involves a reducing agent, which is
water in the case of plants, releasing oxygen as a waste
58 Plant Ecology

,produc~, The light energy is converted to chemical energy


(knowr..;;\ts light-dependent reactions), in the form of ATP
and N~.DPH, -which are used for synthetic reactions in
\'
pho~Qa\ltotrophs. The overall equation for the light-
dependent reactions under the conditions of non-cydic
electron flow in green plants is:

2~0 + 2NADP+ + 2ADP + 2Pi + light ~ 2NADPH + 2H+


+ 2ATP + 02

Most notably, plants use the chemical energy to fix carbon


dioxide into carbohydrates ann other organic compounds
through light-independent reactions. The overall equation
for carbon fixation (sometimes referred to as carbon
reduction) in green plants is:

3C02 + 9ATP + 6NADPH + 6 H+ ~ C3HP3-phosphate -t


9ADP + BPi + 6NADP+ + 3 Hp

To be more specific, carbon fixation produces an


intermediate product, which is then converted to the final
carbohydrate products. The carbon skeletons produced by
photosynthesis are then variously used to form other
organic compounds, such as the building material cellulose,
as precursors for lipid and amino acid biosynthesis, or as
a fuel in cellular respiration. The latter occurs not only in
plants but also in animals when the energy from plants gets
passed through a food chain.
Organisms dependent on photosynthetic and
chemosynthetic organisms are called heterotrophs. In
general outline, cellular respiration is the opposite of
photosynthesis: Glucose and other compounds are
oxidized to produce carbon dioxide, water, and chemical
energy. However, the two processes take place through a
different sequence of chemical reactions and in different
cellular compartments.
Impact of Physical Environment on Plant Growth S9

Plants absorb light primarily using the pigment


chlorophyll, which is the reason that most plants have a
green color. The function of chlorophyll is often supported
by other accessory pigments such as carotenes and
xanthophylls. Both chlorophyll and accessory pigments are
contained in organelles (compartments within the cell)
called chloroplasts. Although all cells in the green parts of
a plant have chloroplasts, most of the energy is captured
in the leaves.
The cells in the interior tissues of a leaf, called the
mesophyll, can contain between 450,000 and 800,000
chloroplasts for every square millimeter of leaf. The surface
of the leaf is uniformly coated with a water-resistant waxy
cuticle that protects the leaf from excessive evaporation of
water and decreases the absorption of ultraviolet or blue
light to reduce heating. The transparent epidermis layer
allows light to pass through to the palisade mesophyll cells
where most of the photosynthesis takes place.
Plants convert light into chemical energy with a
maximum photosynthetic efficiency of approximately 6%.
By comparison solar panels convert light into electric
energy at a photosynthetic efficiency of approximately 10-
20%. Actual plant's photosynthetic efficiency varies with
the frequency of the light being converted, light intensity,
temperature and proportion of CO2 in atmosphere.
Algae come in multiple forms from multicellular
organisms like kelp, to microscopic, single-cell organisms.
Although they are not as complex as land plants, the
biochemical process of photosynthesis is the same. Very
much like plants, algae have chloroplasts and chlorophyll,
but various accessory pigments are present in some algae
such as phycocyanin, carotenes, and xanthophylls in green
algae and phycoerythrin in red algae (rhodophytes),
resulting in a wide variety of colors.
Brown algae and diatoms contain fucoxanthol as their
primary pigment. All algae produce oxygen, and many are
60 Plant Ecology

autotrophic. However, some are heterotrophic, relying on


materials produced by other organisms. For example, in
coral reefs, there is a mutualistic relationship between
zooxanthellae and the coral polyps.
Photosynthetic bacteria do not have chloroplasts (or
any membrane-bound organelles). Instead, photosynthesis
takes place directly within the cell. Cyanobacteria contain
thylakoid membranes very similar to those in chloroplasts
and are the only prokaryotes that perform oxygen-
generating photosynthesis.
In fact, chloroplasts are now considered to have evolved
from an endosymbiotic bacterium, which was also an
ancestor of and later gave rise to cyanobacterium. The other
photosynthetic bacteria have a variety of different
pigments, called bacteriochlorophylls, and do not produce
oxygen. Some bacteria, such as Chromatium, oxidize
hydrogen sulfide instead of water for photosynthesis,
producing sulfur as waste.

Evolution of Photosynthetic Systems


The ability to convert light energy to chemical energy
confers a significant evolutionary advantage to living
organisms. Early photosynthetic systems, such as those
from green and purple sulfur and green and purple non-
sulfur bacteria, are thought to have been anoxygenic, using
various molecules as electron donors.
Green and purple sulfur bacteria are thought to have
used hydrogen and sulfur as an electron donor. Green
nonsulfur bacteria used various amino and other organic
acids. Purple nonsulfur bacteria used a variety of non-
specific organic molecules. The use of these molecules is
consistent with the geological evidence that the atmosphere
was highly reduced at that time.
Oxygen in the atmosphere exists due to the evolution
of oxygenic photosynthesis, sometimes referred to as the
Impact of Physical Environment on Plant Growth 61

oxygen catastrophe. Geological evidence suggests that


oxygenic photosynthesis, such as that in cyanobacteria,
became important during the Paleoproterozoic era around
2 billion years ago. Modem photosynthesis in plants and
most photosynthetic prokaryotes is oxygenic. Oxygenic
photosynthesis uses water as an electron donor which is
oxidized into molecular oxygen by the absorption of a
photon by the photosynthetic reaction center.
In plants the process of photosynthesis occurs in
organelles called chloroplasts. Chloroplasts have many
similarities with photosynthetic bacteria including a
circular chromosome, prokaryotic-type ribosomes, and
similar proteins in the photosynthetic reaction center.
The endosymbiotic theory suggests that photosynthetic
bacteria were acquired (by endocytosis or gene fusion) by
early eukaryotic cells to form the first plant cells. In other
words, chloroplasts may simply be primitive
photosynthetic bacteria adapted to life inside plant cells,
whereas plants themselves have not actually evolved
photosynthetic processes on their own.
Another example of this can be found in complex plants
and animals, including humans, whose cells depend upon
mitochondria as their energy source; mitochondria are
thought to have evolved from endosymbiotic bacteria,
related to modem Rickettsia bacteria. Both chloroplasts and
mitochondria actually have their own DNA, separate from
the nuclear DNA of their animal or plant host cells.
This ~ontention is supported by the finding that the
marine molluscs Elysia viridis and Elysia chlorotica seem
to maintain a symbiotic relationship with chloroplasts frl'JIn
algae with similar RDA structures that they encounter.
However, they do not transfer these chloroplasts to the next
generations.
The biochemical capacity to use water as the source for
electrons in photosynthesis evolved once, in a coIIU1l.on
62 Plant Ecology

ancestor of extant cyanobacteria. The geological record


indicates that this transforming event took place early in
our planet's history, at least 2450-2320 million years ago
(Ma), and possibly much earlier.
Geobiological interpretation of Archean (>2500 Ma)
sedimentary rocks remains a challenge; available evidence
indicates that life existed 3500 Ma, hut the question of when
oxygenic photosynthesis evolved continues to engender
debate and research. A clear paleontological window on
cyanobacterial evolution opened about 2000 Ma, revealing
an already-diverse biota of blue-greens.
Cyanobacteria remained principal primary producers
throughout the Proterozoic Eon (2500-543 Ma), in part
because the redox structure of the oceans favored
photautotrophs capable of nitrogen fixation. Green algae
joined blue-greens as major primary producers on.
continental shelves near the end of the Proterozoic, but only
with the Mesozoic (251-65 Ma) radiations of dinoflagellates,
coccolithophorids, and diatoms did primary production in
marine shelf waters take modern form.
Cyanobacteria remain critical to marine ecosystems as
primary producers in oceanic gyres, as agents of biological
nitrogen fixation, and, in modified form, as the plastids of
marine algae.

Carbon Fixation
The fixation or reduction of carbon dioxide is a light-
independent process in which carbon dioxide combines
with a five-carbon sugar, ribulose 1,5-bisphosphate (RuBP),
to yield two molecules of a three-carbon compound,
glycerate 3-phosphate (GP), also known as 3-
phosphoglycerate (PGA). GP, in the presence of ATP and
NADPH from the light-dependent stages, is reduced to
glyceraldehyde 3-phosphate (G3P). This product is also
referred to as 3-phosphoglyceraldehyde (PGAL) or even as
triose phosphate.
Impact of Physical Environment on Plant Growth 63

Triose is a 3-carbon sugar. Most (5 out of 6 molecules)


of the G3P produced is used to regenerate RuBP so the
process can continue. The lout of 6 molecules of the triose
phosphates not "recycled" often condense to form hexose
phosphates, which ultimately yield sucrose, starch .md
cellulose. The sugars produced during carbon metabolif, 11
yield carbon skeletons that can be used for other metaboli\-
reactions like the production of amino acids and lipids.

C4, C3 and CAM


In hot and dry conditions, plants will close their stomata
to prevent loss of water. Under these conditions, oxygen
gas, produced by the light reactions of photosynthesis, will
concentrate in the leaves causing photorespiration to occur.
Some plants have evolved mechanisms to increase the CO2
concentration in the leaves under these conditions.
C4 plants capture carbon dioxide using an enzyme
called PEP Carboxylase that adds carbon dioxide to the
three carbon molecule Phosphoenolpyruvate (PEP)
creating the 4-carbon molecule oxaloacetic acid. Plants
without this enzyme are called C3 plants because the
primary carboxylation reaction produces the three-carbon
sugar 3-phosphoglycerate directly in the Calvin-Benson
Cycle. When oxygen levels rise in the leaf, C4 plants reverse
the reaction to release carbon dioxide thus preventing
photorespiration. By preventing photorespiration, C4 plants
can produce more sugar than C3 plants in conditions of
strong light and high temperature. Many important crop
plants are C4 plants including maize, sorghum, sugarcane,
and millet.
Xerophytes such as cacti and most succulents also can
use PEP Carboxylase to capture carbon dioxide in a process
called Crassulacean acid metabolism (CAM). They store the
CO2 in different molecules than the C4 plants (mostly they
store it in the form of malic acid via carboxylation of
phosphoenolpyruvate to oxaloacetate, which is then
64 Plant Ecology

reduced to malate). Nevertheless, C4 plants capture the CO2


in one type of cell tissue (mesophyll) and then transfer it
to another type of tissue (bundle sheath cells) so that carbon
fixation may occur via the Calvin cycle. They also have a
different leaf anatomy than C4 plants. They grab the CO2
at night, when their stomata are open, and they release it
into the leaves during the day to increase their
photosynthetic rate. C4 metabolism physically separates
CO2 fixation from the Calvin cycle, while CAM metabolism
temporally separates CO2 fixation from the Calvin cycle.

Phototropism
Phototropism is directional growth in which the direction
of growth is determined by the direction of the light source.
Phototropism is most often observed in plants, but can also
occur in other organisms such as fungi. Phototropism is one
of the many plant tropiSms or movements which respond
to external stimuli.
Growth towards a light source is a positive
phototropism, while growth away from light is called
negative phototropism. Most plant shoots exhibit positive
phototropism, while roots usually exhibit negative
phototropism, although gravitropism may playa larger role
in root behavior and growth. Some vine shoot tips exhibit
negative phototropism, which allows them to grow
towards dark, solid objects and climb them.
Phototropism in plants such as Arabidopsis thaliana is
regulated by blue light receptors called phototropins. Other
photosensitive receptors in plants include phytochromes
that sense red light and cryptochromes that sense blue light.
Different orgOans of the plant may exhibit different
phototropic reactions to different wavelengths of light.
Stem tips exhibit positive phototropic reactions to blue
light, while root tips exhibit negative phototropic reactions
to blue light. Both root tips and most stem tips exhibit
positive 'phototropism to red light.
Impact of Physical Environment on Plant Growth 6S

Phototropism is enabled by auxins. Auxins are plant


hormones that have many functions. In this respect, auxins
are responsible for expelling H+ ions (creating proton
pumps) which decreases pH in the cells on the dark side
of the plant. This acidification of the cell wall region
activates enzymes known as expansins which break bonds
in the cell wall structure, making the cell walls less rigid.
In addition, the acidic environment causes disruption of
hydrogen bonds in the cellulose that makes up the cell wall.
The decrease in cell wall strength causes cells to swell,
exerting the mechanical pressure that drives phototropic
movement. Phototropism relates to photosynthesis.

Photoperiodism
Photoperiodicity is the physiological reaction of organisms
to the length of day or night. It occurs in plants and animals.
Many flowering plants use a photoreceptor protein, such as
phytochrome or cryptochrome, to sense seasonal changes
in day length, which they take as signals to flower.
Broadly, flowering plants can be classified as long day
plants, short day plants, or day neutral plants. Long day
plants are plants that flower when the day is longer than
a critical length (i.e. the night is shorter than a critical
length). These plants generally flower in the spring or early
summer, as days are getting longer.
Short day plants are plants that flower when the day is
shorter than a critical length, or the night is longer than a
critical length. These plants generally flower in late summer
or fall, as days are getting shorter.
It is actually the night length rather than day length that
controls flowering, so flowering in a long day plant is
triggered by a short night (which of course will mean it also
sees a long day). Conversely, short day plants will flower
when nights get longer than a critical length. This is known
by using night break experiments. Fcx example, a short day
66 Plant Ecology

plant (long night) will not flower It a pwse (say 5 minutes)


of artificial light is shone on the plant during the middle
of the night. This generally does not occur from natural
light such as moonlight, lightning, fire flies, etc, since the
light from these sources is not sufficiently strong to trigger
the response.
Day- neutral plants do not initiate flowering based on
photoperiodism i.e. they can flower regardless of the night
length; some may use temperature (vernalization) instead.
Quantitative long day or short day plants will have their
flowering advanced or retarded by short or long days, but
will eventually flower in sub-optimal day lengths. Again,
temperature is likely to also influence flowering time in
these plants.
Modem biologists believe that it is the coincidence of
the active forms of phytochrome or cryptochrome, created
by light during the daytime, with the rhythms of the
circadian clock that allows plants to measure the length of
the night. Other instances of photoperiodism in plants
include the growth of stems or roots during certain seasons,
or the loss of leaves.
TEMPERATURE EFFECI'S ON PLANT

Sometimes temperatures are used in connection with day


length to manipulate the flowering of plants.
Chrysanthemums will flower for a longer period of time if
daylight temperatures are 50°F. The Christmas cactus forms
flowers as a result of short days and low temperatures.
Temperatures alone also influence flowering. Daffodils
are forced to flower by putting bulbs in cold storage in
October at 35 to 40°F. The cold temperature allows the bulb
to mature. The bulbs are transferred to the greenhouse in
midwinter where growth begins. The flowers are then
ready for cutting in 3 to 4 weeks.
Thermoperiod refers to daily temperature change.
Plants produce maximum growth when exposed to a day
Impact of Physical Environment on Plant Growth 67

temperature that is about 10 to 15°F higher than the night


temperature. This allows the plant to photosynthesize
(build -up) and respire (break down) during an optimum
daytime temperature, and to curtail the rate of respiration
during a cooler night. High temperatures cause increased
respiration, sometimes above the rate of photosynthesis.
This means that the products of photosynthesis are being
used more rapidly than they are being produced. For
growth to occur, photosynthesis must be greater than
respiration.
Low temperatures can result in poor growth.
Photosynthesis is slowed down at low temperatures. Since
photosynthesis is slowed, growth is slowed, and this results
in lower yields. Not all plants grow best in the same
temperature range. For example, snapdragons grow best
when night time temperatures are 55°F, while the poinsettia
grows best at 62°F. Florist cyclamen ,does well under very
cool conditions, while many bedding plants grow best at
a higher temperature.
Buds of many plants require exposure to a certain
number of days below a critical temperature (chilling
hours) before they will resume growth in the spring.
Peaches are a prime example; most cultivars require 700 to
1,000 hours below 45°F and above 32°F before they break
their rest period and begin growth. This time period varies
for different plants. The flower buds of forsythia require a
relatively short rest period and will grow at the first sign
of warm wea!her. During dormancy, buds can withstand
very low temperatures, but after the rest period is satisfied,
buds become more susceptible to weather conditions, and
can be damaged easily by cold temperatures or frost.
PLANTS AND HUMIDITY

Humidity is a measure of the amount of water that air will


hold. The water is usually in the form of invisible droplets.
At 100 percent humidity the air cannot hold any more
68 Plant Ecology

water. The highest humidity often occurs on hot days,


creating a "muggy" feeling. Fog occurs when the air is
saturated and the invisible water now becomes visible.
Humidity is measured relative to temperature and is
called relative humidity (RH). The measurement is taken
this way because humidity and temperature are directly
related: the warmer the air, the more water it can hold.
Humidity in the Home If warm air holds more moisture
than cold, then why is it so dry inside the house in winter?
Remember that the furnace is taking dry outside air and
warming it. If no water is added to this outside air, then
it will still be dry. You can increase the humidity inside to
,a certain extent by adding water to the air. Warm, moist air
is always being lost from the house, and cold dry air is
always being brought in, so high humidity in the entire
house is not possible.
When warm, moist air comes in contact with a cold, dry
surface, the water in the air condenses. This is very common
on windows, and is an indication that the humidity inside
the house is higher than outside. If the inside walls of the
house are cooler than the air inside the house, water can
condense on the walls, and can cause wallpaper to come
unstuck, but don't rely on this as a means of stripping
wallpaper.
HUmidity is important to plants because it partly
controls the moisture loss from the plant. The leaves of
plants· have tiny pores in them called stomata. Carbon
dioxide enters the plants through these pores; oxygen and
water leave through them.
The humidity inside a plant is close to 100%. A plant
growing in a dry room will lose moisture because water
always moves from high to low humidity. When the
difference in humidity is large, the loss of moisture from the
plant is rapid and severe.
Impact of Physical Environment on Plant Growth 69

Most houseplants prefer a humidity of about 60%. Cacti,


succulents and plants native to desert environments
tolerate much lower humidity (30-35%), but prefer not to
drop below 20%. House plants that are native to tropical
rain forests require much higher humidity, 90% for
example, and thus pose problems for most home owners.
Plants that require a very high humidity are best grown in
terrariums or closed containers where it is possible to
regulate the humidity.
Under very humid environments, fungal diseases can
spread. This seldom happens during winter, but can be a
problem in fall when the temperature is cool and the
humidity is high. Mildew on plant leaves is an indication
of excess humidity and lack of ventilation.
Plants that prefer a more humid environment, but that
are forced to grow in a dry environment will commonly
suffer damage to younger leaves and to leaf tips. New
leaves and leaf tips are the area of the plant with the most
actively growing cells, and these cells are the most
susceptible to dry air. Older leaves that are fully formed
may be shed as a result of lack of humidity, but they will
not be deformed or damaged by the dry air. Plants stressed
in this way very frequently shed flower buds, or flowers die
soon after opening.
There are a number of ways that a home owner can
increase the humidity in the room or around the plants.
Humidifiers-Using a humidifier is by far the most
effective way to increase humity. Humidifiers that
attach directly to the furnace will increase the humidity
throughout the house. Portable humidifiers can be
used to increase the humidity in one or more rooms.
Changing locations-Bathrooms and kitchens, if they are
sunny, often have a higher humidity than other areas
of the home, and may be more suitable for house plants
requiring extra humidity.
70 Plant Ecology

Double potting-Take a small potted plant or a number


of small plants, and put them in a larger pot. Fill the
area underneath and around the small pots with peat
moss. Keep the peat moss constantly moist. As water
evaporates from the peat moss, it increases the
humidity around the plants. Make sure the large pot
has a tray underneath to catch excess moisture from the
peat moss. A similar approach is to place a house plant
in a basket lined with moist peat moss.
Pebble trays-Fill a large plant saucer with pebbles or
stones. Place a number of small pots (or a large pot) on
top of the stones. To assure that the pots do not contact
the water, you may wish to place them on saucers
which sit on the pebbles. Now fill the larg~ plant saucer
with water up to the level of the pebbles. Make sure
the saucer with pebbles is large enough to be effective
- the larger the surface area of pebbles, the more
effective the method will be.
Totems for climbing plants-Take chicken wire and roll
it into a totem (tube). Fill the tub with peat moss.
Anchor the tube in the plant pot and then wind the
climbing plant around the tube. Keep the peat moss
inside the tub moist. Do not worry if the plant forms
roots into the totem, but if this occurs make sure the
totem is kept evenly moist.
Grouping-Moisture loss from one plant can benefit the
plant next to it. Try and group plants with similar
watering requirements together, and keep them close
to each other. The closer together they are, the more
effective the method.
Misting-This is the least effective but often the most
used method. Misting plants with tepid water will
result in a layer of water on the leaves, which will
reduce the transpiration of water from the leaves.
However, soon after misting, the water will evaporate,
and once this occurs, the air is once again low in
Impact of Physical Environment on Plant Growth 71

humidity. If plants are misted too often or too much,


however, fungal growth and tissue rotting may result.
Plants with hairy leaves cannot be misted, for leaf
spotting wi1llikely occur as a result.
EFFECT OF WIND ON PLANT GROWTH

One dictionary defined wind as "air moving hOrizontally."


It is not quite correct because some winds blow up and
some down. Such a definition gives no idea of the diversity
of winds; wind is moving air; that may carry water, dust,
ice, sand and chemicals. Above all wind has energy that can
be transferred to sailing ships and windmills and can be
used by plants in ways essential to their growth and
development. Different is the "wet monsoon" wind of India
from the dry hot "Mistral" blowing into France from the
Sahara, from the "foehn" blowing from the European Alps
or from the "Squamish" winds blowing into the fiords of the
Be coast from the Interior Plateau.
To the ancients, wind was mysterious. Where did it
come from and where did it go? In Scandinavian
mythology, Thor was the god of the north wind and the god
of battle and tempest. Aeolius was the ancient Greek god
of the four winds. Today wind has lost much of what the
ancients found so intangible but we still feel the "poetic" or
"archaic" wind as somewhat different from the winds of
science, the winds of air mass analysis, the jet stream or the
solar wind.
When unicellular plants first established on land, wind
may have played a role in their dispersal. Once they
evolved as multicellular and erect, they have affected and
been affected by wind speed and direction. Fossil evidence
suggests that by the Silurian, 500 to 700 million years ago,
leaf-like structures with stomata and semi-permeability
had evolved and that the sporophyte generation of the seed
plants was large (and the gametophyte generation
diminished) and a conducting system with root and stem
72 Plant Ecology

had evolved. Why the bryophytes with a relatively large


leafy gametophyte and a small sporophyte with only
rhizoids and a poorly developed conducting system should
remain short statured, hugging the ground as the mosses
and liverworts do today is not dear from the fossil record.
Wind played a role in the evolution of the root as a
holdfast and uptake organ for water and soil nutrients, and
in evolution of the stem to support leaf and flower. The
vascular system provided effective transport between root
and shoot to meet the needs of the expanding
photosynthetic surfaces. Does wind continue to playa role
as one of many factors in the evolu-tion of plants? The
internal and external architecture of leaves seem to affect
the type of damage caused by high wind. Often the
parallelveined monocot leaves are torn to strips reticulate-
veined dicot leaves are bent at the petiole or broken at the
stem insertion.
Wind, Flex Plants and Brittle Plants
Broken cottonwoods silhouetted against a Fraser Valley sky
attest to the brittleness of old branches in wind and ice
storms while nearby the closely related Lombardy poplars
gracefully sway unbroken in the wind. The American,
Robert Frost in his famous poem, "Birches" recalls from
boyhood swinging up and down on young and bending
birch trees.
Older birch trees are brittle. Zea mays (Indian corn)
plants in the windy areas produce brace roots, whereas
other cereals, grasses and many forbs, such as Epilobium
angustifolium (fireweed) flatten and lodge in windstorms,
but when young become erect again. Grasses have special
structures termed pulvini on stem (culm) just below nodes
that assist lodged grasses to become erect again. Young
arborescent species, such as Alnus sinuata (slide alder) that
have been flattened by wind, avalanche or the weight of
Impact of Physical Environment on Plant Growth 73

snowpacks can often become erect again. Astonishingly the


wind helps to untangle and separate members and to assist
in the process of re-erection. There is a special charm in the
flexible shoots of weeping willows or birches. Landscapers
often place them on the margins of lakes and streams
providing a curtain of swaying limbs that create a dynamic
view across the water.

Wind and Aquatic Plants


Most aquatic plants are flexible. The energy balances
between physical support and transport in land plants that
are provided by lignin and cellulose are changed because
the water provides much of the physical support for stem
and leaf. Some aquatics are submerged and largely out of
the direct impact of the buffeting effects of wind and wave
but others such as water lilies and pond weeds produce
foliage on the water surface or near the surface.
The surface foliage of aquatics may serve to calm small
wind-driven small wave motion. In the shallow bays of lake
margins and in ponds the long petioles and "split" fronds
of water lilies or the ribbon-like leaves of the pondweeds
absorb the energy of the waves. Other emergents such as
the bulrush (Scirpus spp.) rely on the round or triangular
cross-section of their rigid stems to absorb most of the
buffeting of wind and waves.
Flexibility is only one of several factors important in the
ecology of emergent aquatics; for example, the timing of the
emerging first floating leaves of Zizania palustris (annual
wild rice) and the water regime and depth appears to
confine its geographical range to the shallow waters of the
Great Lakes region. Once past the critical flex-leaf stage,
stiff culms arise above the water surface. The timing of the
leaf emergence must match the rise in the spring water
levels; if the levels are too high, the first floating leaves die
and the emergent wave resistant culms are not produced.
Mankind has, from ancient times used natural fencing of
74 Plant Ecology

trees and shrubs to reduce the impact of wind on dwellings


and on gardens. The use of flexible whips (branches) by
coppicing, pleaching and inosculation deserves mention-
c,oppicing to grow and harvest whips, pleaching to
interweave whips to make a fence, which when dry
becomes strong, and ino~ulating to link living whips from
one tree or shrub to another. In various ways and to varying
degrees and times almost all plants flex in wind.

Windfall and Breakage


Violent wind can cut into a forest like a scythe cutting grass.
The ecological effects of blowdown, windthrow and
breakage of trees by wind differ somewhat. Blowdown
leaves patches of tangled trees that are often discernable at
a distance. Windthrow of individual trees allows light
penetration of the forest canopy and the sun flecks serve to
enhance understorey vegetation and seedling growth on
decaying "nurse 10gs".In addition, windthrow mixes litter
with mineral soil.
Windfall and breakage leave stumps and sloughing
bark accumulating at the base of boles. Insects and fungi
provide cavities for nesting birds and foocl for woodpeckers
as they decay attack the stumps. Decayed stumps leave
mounds of organic matter on the forest floor cr~ating
microhabitat that may favour western hemlock seedlings
while nearby mineral soil favours Douglas fir seedlings.
Wind, rain, frost, snow and ice all contribute to the
'throughfall', of organic materials shed by woody plants.
This rain builds the duff and litter (the mor and the mull)
on the mineral soil surface. Throughfall occurs in all
seasons but varies in amount. The rain of organic matter
includes pollen, bud scales, diseased needles and leaves,
twigs, insect frass, bark dust, flower parts and whole
catkins, and debris from epiphytes (lichens, mistletoe, etc).
Wind serves to clean trees and shrubs of old unproductive
or diseased materials, removing shade and increasing
Impact of Physical Environment on Plant Growth 7S

exposure to the sun's radiant energy. The throughfall from


winter storms, including twigs and branches from the high
crowns falls on a snowy surface and provides food for
wintering deer, elk, moose and even native sheep.
The twigs are nutrient rich but may also be toxic. Much
has been written about the decomposition of throughfall
and litter, but very little about its function in nature or
about the roles of wind. As if defying gales and age, trees
and shrubs standing on promontories, rocky shores and
sand dunes are often dwarfed and have scraggy limbs and
twisted boles. Trained by wind their foliage is flagged and
tattered. Bonsai art from Japan and China reflects these
forms of woody plants. So do the trees in the paintings of
the Canadian "Group of Seven" and the photos by Ansel
Adams. Bonsai artists in practice pinch leaf and twig to
dwarf a tree or shrub and achieve a wind-tattered form.
In nature, cold or dry wind, sometimes does the
pinching, the pinches remove and kill new twigs and
leaves, greatly reducing the opportunities to build
carbohydrate food reserves. Sometimes loss of leaf and twig
occurs as a result of abrasion by sand or ice or snow crystals
and on marine shores by salty spray. Limited mineral
nutrient and water supplies can only come from root
penetration of cracks in the rocky cliffs or deep root
penetration of sand dunes. Desiccation is often a reason for
dwarfing but it is likely not that alone.

Throughfall and Summer Drought


Dry hot wind and cold soil contribute to water stress in
plants. Water stresses evidenced in summer leaf fall often
escape the attention of photographers, naturalists and the
urbanized public. Plants may pay a high energetic price by
losing green foliage in summer. Aided by wind, leaf
abscission quickly reduces transpiration and water loss.
Older and less photosynthetically active leaves drop first.
The impact of needle and leaf losses on flower and fruit
76 Plant Ecology

production and on carbohydrate reserve food in twig, bole


and root may continue for several years following summer
drought. Leaf fall in summer often goes unnoticed because
there is no associated colour and nutrient withdrawal as is
characteristic of autumn leaf fall.
Hot dry summer winds contribute to other responses to
water stress. Even pubescent plants such as Antennaria
parvifolia (pussytoes) behave as natural hygrometers when
leaves curl during the day reducing transpiration and
uncurl by night or during a humid breeze as they regain
turgor. Other plants, such as the Balsamorrhiza sagittata
(balsam root), draw their crowns into the soil, as the leaves
curl, dry, become brittle and are soon shredded by wind
and the trampling of animals.

Surface, Roughness and Wind Breaks


Even low-growing rosettes and mats reduce wind velocity
and create turbulence. Tall trees influence wind velocity in
their lee for up to one hundred times their height. In dense
forest, wind velocity on the ground may be greatly reduced
while the forest canopy is buffeted. The boles of spaced
trees create eddies manifested in winter by the
development of snow cirques which in turn result in
variations in snow depth, snow melt and soil moisture.
Snow cirques provide a distinctive micro habitat for plants
and animals. In spring snow patches may linger and
summer soil moisture vary over relatively short distances
as a result of variations caused by uneven surfaces, drifting
snow, and living snow fences that influence the winds of
winter.
Land plants evolved in the presence of fires and
lightning. Fire is nearly universal. It is unique in that it
generates its own wind in complex ways with some special
ecological responses. Its destructive, sanitizing or renewing
actions make it both an ecological enemy and friend.
Impact of Physical Environment on Plant Growth 77

Wind and Dispersal of Plants


The ancients and many aboriginal peoples had some
understanding of the roles of wind pollination and pollen
dispersal. In the western world appreciation of wind
function was limited un~il in the 1700's when several
studies on pollination were publiShed. By the mid 1800's,
Victorians such as Charles Darwin had established that
pollination was a noteworthy service in nature.
Th.g fossil record of wind and insect pollination dates
back to the mid-Paleozoic time. Today many more plant
species are wind pollin-ated than insect pollinated,
although public attention is directed more towards
pollination by insects and other animals and to the
colourful co-evolution of flowers and-insects. Many plant
species take advantage of both wind and animal
pollination.
Wind is reliable and wind is everywhere but pollen
must be produced in prodigious quantities ana. demands
the' allocation of much of a plant's energy. The direct
transfer of pollen from flower to flower by insect or other
animal agent is less energy demanding, but insect
populations may be limited by the temperature at which
insects fly and by their distribution.
Air-dispersed pollen falls by gravity to receptive
surfaces in a seemingly haphazard process, but plants have
evolved some efficiencies to aid the basic process. Rate of
fall varies, as does its buoyancy. Some pollen is winged or
variously patterned on the surface. Plant grouping and the
timing of foliage production both can modify and even
direct pollen fall. As sufferers from pollen allergies know,
the release of pollen is temperature and humidity
controlled; as a result, pollen release is effi-cient and timely.
Receptive surfaces and structures, such as the cones of
conifers or the foliage of jojoba, facilitate the sequestering
of airborne pollen.

78 Plant Ecology

Feathery stigmas like those of grasses "pluck" pollen


from passing zephyrs; other stigmas are sticky and hold
grains until germination. The modern literature on
pollination received new impetus and direction in the 1960s
with the work of Faegri and van der Pijl . Spores from
bryophytes and pteridophytes as well as those from fungi
and lichens deserve direct conSideration as do bacteria, but
reviews of this very large literature do not apparently exist,
although the case studies are scattered through the journals
of microbiology, plant pathology and other disciplines.
Wind may not be as effective a dispersal agent of seeds,
fruits and associated plant parts as\it is with the generally
smaller propagules of some fungi, lichens and bryophytes,
some of which may be distributed worldwide. Dispersal is
governed to some extent by the nature of major
reproductive structures such as pods or inflorescence: for
example, in Sporobdus cryptandrus (sand dropseed grass),
where the' whole mature inflorescence with seed (fruits) is
abscised in a leafy sheath, the seed is shaken loose as wind
blows the severed structure over grass and dune.
Whole tumbleweed plants, such as Sisymbrium
atissimum (tumbling mustard), mature, break at soil level
and tumble in the wind across grass prairie and drop seed
from siliques either intermittently as they tumble or when
at a fence or other barrier. Many conifers, elms, maples and
many other plants have winged propagules, which, borne
by wind, may be delivered several hundred metres from
their parent plant.
Wind may assist establishment of seed that falls into
soil, cracks and crannies, because the wings or other
appendages are hygroscopic and respond to changes in
wind humidity and temperature to lengthen and contract
thus driving the seed further into the subsoil. Watch the
behaviour of the awns of Stipa spp. (needlegrass) in a light
'breeze. Wihd can carry parachute "seeds" like those of
Tragopogon spp. (yellow salsify) to elevations of several
Impact of Physical Environment on Plant Growth 79

hundred metres and over low mountains. Balloon fruits


such as those of Physalis peruviana (ground cherry), a
common garden weed, may not travel so far in the wind
but can effectively move from urban lot to urban lot.

Effect of Wind on Leaf Surfaces


The thin layE9 of air immediately adjacent to leaf surfaces
does not behave in quite the same way as ambient air; it
is usefully termed the "boundary layer". Boundary layer air
is not quite motionless because it is held close to the leaf
surface. It may be only a millimetre or two thick, a distance
that varies from species to species and with the nature of
the leaf surface. Leaf form (flat or needle-like), pubescence,
venation, location next to upper or lower leaf surface, and
the number and placement of stomata all affect the
boundary layer. Gas exchange occurs at and through the
semi-permeable protective cuticle and the boundary layer.
This is also where leaf aroma and fragrance are generated,
and where palatability and the flavour of foliage to all
herbivores and omnivores are perceived.
The wind distributes the greenhouse gases, carbon
dioxide and methane, and water vapour, which contribute
to the haze that develops over tropical rainforests and over
some temperate latitude skies on hot summer days. The
chemistry and physics of the boundary layer are difficult
to study, but its gas composition is very complex and the
number of chemicals in one fragrance may alone exceeded
four hundred. Chemistry and physics aside, the astonishing
diversity of leaf surfaces and variety of leaves is telling a
story of immense importance to life on earth and also of a
long evolutionary history of intimacy of plant and wind.
REFERENCES

Bouma J., R.B. Brown, and P.S.c. Rao. 1982. "Basics of Soil-Water
Relationships -Part I. Soil as a Porous Medium." Soil Science Fact
Sheet SL-37. Florida Cooperative Extension Service. IFAS.
Gainesville, FL.
80 Plant Ecology

de Villiers, M. Windswept: the story of wind and weather. Toronto:


McClelland and Stewart 2006.
Dickenson, C.H. and Pugh, G.J.F. (editors) Biology of Plant Litter
Decomposition. London and New York: Academic Press. 1974.
Paergi, K. and van der Pijl, L. The Principles oj Pollination Ecology. Oxford,
UK: Pergamon Press Ltd. 1966.
Merva G.E. Physioengineering principles. 1975. The AVI Publishing
Company, Inc. Westport, CT,
Phillipson, J. Ecological Energenics. London and Beede!>. William Clowes
and Sons. 1966.
Pielou, E.C. The Energy of Nature. Chicago: The University Of Chicago
Press. 200l.
Ridley, H.N. The dispersal of plants throughout the world. Ashford (UK): L.
Reeve. 1930.
Thomas B. and Vince-Prue D., Photoperiodism in plants, Academic Press,
1997.
4
Ecological Evolution of Plants

Plants are multicellular photosynthetic organisms that are


believed to have evolved from green algae. Both groups
have chlorophylls a and band betacarotene as their
photosynthetic pigments, both store reserve food as starch,
and both have cellulose containing cell walls. Perhaps the
evolution of plants began on land, when algae are left high
and dry between tides. When they stayed ashore, they
adapted to the new open-air lifestyle with great success. To
survive in air, they thickened their walls, thus staying wet
on the inside while the air dried them on the outside. Fungi,
a kingdom including molds, mushrooms, and yeasts, also
now appear on shores, where they practice the ability to
digest organic food with excreted enzymes before
consuming it. Some fungi cooperate with algae to form a
single-organism partnership called lichens.
Plant cells are more complex than those of either
animals or fungi because they have both mitochontria and
chloroplasts. The first real plants to evolve are mosses, often
found in close association with fungi. From the time they
invent a cooperative lifestyle as lichens until today, plants
and fungi have a very close association. It may even be that
the first plants are a genetic fusion of the two. Ninety
percent of plants have fungi called mycorrhizal (root
fungus) living in special co-evolved compartments inside
their roots or in the soil, intertwined with roots, making
82 Plant Ecology

food for each other according to specialty and recycling


wastes.
Plants have evolved through a number of grades, from
the earliest algal mats, to bryophytes, lycopods, ferns and
gymnosperms to the complex angiosperms of today. While
the simple plants continue to thrive, especially in the
environments in which they evolved, each new grade of
organisation has eventually become more "successful" than
its predecessors by most measures. Further, most cladistic
analyses suggest that each more complex group arose from
the most complex group at the time.
Evidence suggests that an algal scum formed on the
land 1200 million years ago, but it was not until the
Ordovician period, around 500 million years ago, that land
plants appeared. These begun to diversify in the late
Silurian period, around 420 million years ago, and the fruits
of their diversification are displayed in re.markable detail
in an early Devonian fossil assemblage known as the
Rhynie chert. This chert preserved early plants in cellular
detail, petrified in volcanic springs.
By the middle of the Devonian period most of the
features recognised in plants today are present, including
roots, leaves and seeds. By the late Devonian, plants had
reached a degree of sophistication that allowed them to
form forests of tall trees. Evolutionary innovation
continued after the Devonian period. Most plant groups
were relatively unscathed by the Permo-Triassic extinction
event, although the structures of communities changed.
This may have set the scene for the evolution of flowering
plants in the Triassic, which exploded the Cretaceous and
Tertiary. The latest major group of plants to evolve were the
grasses, which became important in the mid Tertiary, from
arolmd 40 million years ago. The grasses, as well as many
other groups, evolved new mechanisms of metabolism to
survive the low CO2 and warm, dry conditions of the
tropics over the last 10 million years.
Ecological Evolution of Plants 83

Land plants evolved from chlorophyte algae, perhaps


as early as 510 million years ago; their closest living
relatives are the charophytes, specifically Charales.
Assuming that the Charales' habit has changed little since
the divergen~e of lineages, this means that the land plants
evolved from a branched, filamentous, haplontic alga,
dwelling in shallow, fresh water, perhaps at the edge of
desiccating pools.
Plants weren't the first photosynthesisers on land,
though: consideration of weathering rates suggests that
organisms were already living on the land 1200 million
years ago. These organisms were probably small and
simple, forming little more than an "algal scum". The first
evidence of plants on land comes from trilete spores, from
the mid-Ordovician The microstructure of the earliest
spores resembles that of modern liverwort spores,
suggesting they share an equivalent grade of organisation.
Trilete spores are the progeny of spore tetrads. These
consist of four identical, connected spores, produced when
a single cell undergoes meiosis. Spore tetrads are borne by
all land plants, and some algae.
Depending exactly when the tetrad splits, each of the
four spores may bear a "trilete mark", a Y-shape, reflecting
the points at which each cell was squashed up against its
neighbours. However, in order for this to happen, the spore
walls must be sturdy and resistant at an early stage. This
resistance is closely associated with having a desiccation-
resistant outer wall-a trait only of use when spores have
to survive out of water. Indeed, even those embryophytes
that have returned to the water lack a resistant wall, thus
don't bear trilete marks. A close examination of algal spores
shows that none have trilete spores, either because their
walls are not resistant enough, or in those rare cases where
it is, the spores disperse before they are squashed enough
to develop the mark, or don't fit into a tetrahedral tetrad.
84 Plant Ecology

The earliest megafossils of land plants were thalloid


organisms, which dwelt in fluvial wetlands and are found
to have covered most of an early Silurian flood plain. They
could only survive when the land was waterlogged, edit
Once plants had reached the land, th~re were two
approaches to desiccation, The bryophytes avoid it or give
in to it, restricting their ranges to moist settings, or drying
out and putting their metabolism "on hold" until more
water arrives. Tracheophytes resist desiccation. They all
bear a waterproof outer cuticle layer wherever they are
exposed to air, to reduce water loss-but since a total
covering would cut them off from CO2 in the atmosphere,
they rapidly evolved stomata-small openings to allow gas
exchange. Tracheophytes also developed vascular tissue to
aid in the movement of water within the organisms, and
moved away from a gametophyte dominated life cycle.
The establishment of a land-based fauna permitted the
accumulation of oxygen in the atmosphere as never before,
as the new hoardes of land plants pumped it out as a waste
product. When this concentration rose above 13%, it
permitted the possibility of wildfire. This is first recorded
in the early Silurian fossil record by charcoalified plant
fossils. Apart from a controversial gap in the Late
Devonian, charcoal is present ever since.
Charcoalification is an important taphonomic mode.
Wildfire drives off the volatile compounds, leaving only a
shell of pure carbon. This is not a viable food source for
herbivores or detritovores, so is prone to preservation; it is
also robust, so can withstand pressure and display
exquisite, sometimes sub-cellular, detail.
CHANGING LIFE CYCLES

All multicellular plants have a life cycle comprising two


phases (often confusingly referred to as "generations"). One
is termed the gametophyte, has a single set of chromosomes
(denoted In), and produces gametes (sperm and eggs). The
Ecological Evolution of Plants 85

other is termed the sporophyte, has paired chromosomes


(denoted 2n), and produces spores. The two phases may be
identical, or phenomenally different.
The overwhelming pattern in plant evolution is for a
reduction of the gametophytic phase, and the increase in
sporophyte dominance. The algal ancestors to land plants
were almost certainly haplobiontic, being haploid for all
their life cycles, with a unicellular zygote providing the 2n
stage. All land plants (i.e. embryophytes) are diplobiontic-
that. is, both the haploid and diploid stages are
multicellular. There are two competing theories to explain
the appearance of a diplobiontic lifecyc1e.
The interpolation theory (also known as the antithetic
or intercalary theory) holds that the sporophyte phase was
a fundamentally new invention, caused by the mitotic
division of a freshly germinated zygote, continuing until
meiosis produces spores. This theory implies that the first
sporophytes would bear a very different morphology to the
gametophyte, on which they would have been dependant.
This seems to fit well with what we know of the
bryophytes, in which a vegetative thalloid gametophyte is
parasitised by simple sporophytes, which often comprise
no more than a sporangium on a stalk.
Increasing complexity of the ancestrally simple
sporophyte, including the eventual acquisition of
photosynthetic cells, would free it from its dependence on
a sporophyte, as we see in some hornworts, and eventually
result in the sporophyte developing organs and vascular
tissue, and becoming the dominant phase, as in the
tracheophytes. This theory may be supported by
observations that smaller Cooksonia individuals must have
been supported by a gametophyte generation. The
observed appearance of larger axial sizes, with room for
photosynthetic tissue and thus self-sustainability, provides
a possible route for the development of a self-sufficient
sporophyte phase.
86 Plant Ecology

The alternative hypothesis is termed the transformation


theory (or homologous theory). This posits that the
sporophyte appeared suddenly by a delay in the occurrence
of meiosis after the zygote germinated. Since the same
genetic material would be employed, the haploid and
diploid phases would look the same. This explains the
behaviour of some algae, which produce alternating phases
of identical sporophytes and gametophytes. Subsequent
.adaption to the desiccating land environment, which makes
sexual reproduction difficult, would result in the
simplification of the sexually active gametophyte. and
elaboration of the sporophyte phase to better disperse the
waterproof spores. The tissue of sporophytes and
gametophytes preserved in the Rhynie chert is of similar
complexity, which is taken to support this hypothesis.
WATER TRANSPORT

In order to photosynthesise, plants must uptake CO2 from


the atmosphere. However, this comes at a price: while
stomata are open to allow CO 2 to enter, water can
evaporate: Water is lost much faster than CO2 is absorbed,
so plants need to replace it, and have developed systems
to transport water from the moist soil to the site of
photosynthesis. Early plants sucked water between the
walls of their cells, then evolved the ability to control water
loss (and CO2 aquisition) through the use of stomata.
Specialised water transport tissues soon evolved in the
form of hydroids, tracheids, then secondary xylem,
followed by an endodermis and ultimately vessels.
The high CO2 levels of the Silu-Devonian, when early
plants were colonising land, meant that the need for water
was relatively low in the early days; as CO2 was withdrawn
from the atmosphere by plants, and stored in coal, so more
water was lost in its .capture, and more elegant transport
mechanisms evolved. As water transport mechanisms, and
waterproof cuticles, evolved, plants could survive without
Ecological Evolution of Plants 87

being continually covered by a film of water. This transition


from poikilohydry to homoiohydry opened up new
potential for colonisation. Plants were then faced with a
balance, between transporting water as efficiently as
possible and preventing transporting vessels to implode
and cavitate.
During the Silurian, CO2 was readily available, so little
water needed expending to acquire it. By the end of the
Carboniferous, when COo levels had lowered to something
approaching today's, around 17 times more water was lost
per unit of CO2 uptake. However, even in these "easy" early
days, water was at a premium, and had to be transported
to parts of the plant from the wet soil to avoid desiccation.
This early water transport took advantage of the cohesion-
tension mechanism inherent in water.
Water has a tendency to diffuse to areas that are drier,
and this process is exacberated when water can be wicked
along a fabric with small spaces. In small passages, such as
that between the plant cell walls (or in tracheids), a column
of water behaves like steel -when molecules evaporate
from one end, they literally pull the molecules behind them
along the channels. Therefore transpiration alone provided
the driving force for water transport in early plants.
However, without dedicated transport vessels, the
cohesion-tension mechanism cannot transport water more
than about 2cm, severely limiting the size of the earliest
plants. This process demands a steady supply of water from
one end, to maintain the chains; to avoid exhausing it,
plants developed a waterproof cuticle; early cuticle may not
have had pores but did not cover the entire plant surface,
so that gas exchange could continue. However,
dehydration at times was inevitable; early plants cope with
this by having a lot of water stored between their cell walls,
and when it comes to it sticking out the tough times by
putting life "on hold" until more water is supplied.
88 Plant Fcolll~)

In order to be free from the constraints of small size and


constant moisture that the parenchymatiC' transport system
inflicted, plants needed a more efficient water transport
system. During the early {'ilurian, they developed
specialized cells, which were\{ignified (or bore similar
chemical compounds) to avoid implosion; this process
coincided with cell death, allowing their innards to be
emptied and water to be passed through them. These wider,
dead, empty cells were a million times more conductive
than the inter-cell method, giving the potential for transport
over longer distances, and higher CO2 diffusion rates.
The first macrofossils to bear water-transport tubes in
situ are the early Devonian pretracheophytes Aglaophyton
and Horneophyton, which have structures very similar to
the hydroids of modern mosses. Plants continued to
innovate new ways of reducing the resistance to flow
within their cells, thereby increasing the efficiency of their
water transport.
Bands on the walls of tubes, in fact apparent from the
early Silurian onwards, are an early improvisation to aid
the easy flow of water. Banded tubes, as well as tubes with
pitted ornamentation on their walls, were lignified and,
when they form single celled conduits, are considered to be
tracheids. These, the "next generation" of transport cell
design, have a more rigid structure than hydroids, allowing
them to cope with higher levels of water pressure,
Tracheids may have a single evolutionary origin, possibly
within the hornworts, uniting all tracheophytes.
Water transport requires regulation, and dynamic
control is provided by stomata. By adjusting the amount of
gas exchange, they can restrict the amount of water lost
through transpiration. This is an important role where
water supply is not constant, and indeed stomata appear to
have evolved before tracheids, being present in the non-
vascular hornworts.
Ecological Evolution of Plants 89

An endodermis probably evolved during the Silu-


Devonian, but the first fossil evidence for such a structure
is Carboniferous. This structure in the roots covers the
water transport tissue and regulates ion exchange (and
prevents unwanted pathogens etc. from entering the water
transport system). The endodermis can also provide an
upwards pressure, forcing water out of the roots when
transpiration is not enough of a driver.
Once plants had evolved this level of controlled water
transport, they were truly homoiohydric, able to extract
water from their environment through root-like organs
rather than relying on a film of surface moisture, enabling
them to grow to much greater size. As a result of their
independence from their surroundings, they lost their
ability to survive desiccation-a costly trait to retain.
During the Devonian, maximum xylem diameter
increased with ~ime, with the minimum diameter
remaining pretty constant. By the middle Devonian, the
tracheid diameter of some plant lineages had plateaued.
Wider tracheids allow water to be transported faster, but
the overall transport rate depends also on the overall cross-
sectional area of the xylem bundle itself. The increase in
vascular bundle thickness further seems to correlate with
the width of plant axes, and plant height; it is also closely
related to the appearance of leaves and increased stomatal
density, both of which would increase the demand for
water.
While wider tracheids with robust walls make it
possible to achieve higher water transport pressures, this
increases the problem of cavitation. Cavitation occurs when
a bubble of air forms within a vessel, breaking the bonds
between chains of water molecules and preventing them
from pulling more water up with their cohesive tension. A
tracheid, once cavitated, cannot have its embolism removed
and return to service. Therefore it is well worth plants'
while to avoid cavitation occurring. For this reason, pits in
90 Plant Ecology

tracheid walls have very small diameters, to prevent air


entering and allowing bubbles to nucleate. Freeze-thaw
cycles are a major cause of cavitation. Damage to a
tracheid's wall almost inevitably leads to air leaking in and
cavitation, hence the importance of many tracheids
working in parallel.
Cavitation is hard to avoid, but once it has occurred
plants have a range of mechanisms to contain the damage.
Small pits link adjacent conduits to al10w fluid to flow
between them, but not air--although ironically these pits,
which prevent the spread of embolisms, are also a major
cause of them. These pitted surfaces further reduce the flow
of water through the xylem by as much as 30%. Conifers,
by the Jurassic, developed an ingenious improvement,
using valve-like structures to isolate cavitated elements.
These torus-margo structures have a blob floating in the
middle of a donut; when one side depressurises the blob
is sucked into the torus and blocks further flow.
Growing to height also employed another trait of
tracheids-the support offered by their lignified walls.
Defunct tracheids were retained to form a strong, woody
stem, produced in most instances by a secondary xylem.
However, in early plants, tracheids were too mechanically
vulnerable, and retained a central position, with a layer of
tough sclerenchyma on the outer rim of the stems. Even
when tracheids do take a structural role, they are supported
by sclerenchymatic tissue,
Tracheids end with walls, which impose a great deal of
resistance on flow; vessel members have perforated end
walls, and are arranged in series to operate as if they were
one continuous vessel. The function of end walls, which
were the default state in the Devonian, was probably to
avoid embolisms. An embolism is where an air bubble is
created in a tracheid. This may happen as a result of
freezing, or by gases dissolving out of solution. Once an
embolism is formed, it usually cannot be removed; the
Ecological Evolution of Plants 91

affected cell cannot pull water up, and is rendered useless.


End walls excluded, the tracheids of prevascular plants
were able to operate under the same hydraulic conductivity
as those of the first vascular plant, Cooksonia.The size of
tracheids is limited as they comprise a single cell; this limits
their length, which in turn limits their maximum useful
diameter to 80 pm. Conductivity grows with the fourth
power of diameter, so increased diameter has huge
rewards; vessel elements, consisting of a number of cells,
joined at their ends, overcame this limit and allowed larger
tubes to form, reaching diameters of up to 500 pm, and
lengths of up to 10 m.
Vessels first evolved during the dry, low CO2 periods
of the late Permian, in the horsetails, ferns and
Selaginellales independently, and later appeared in the mid
Cretaceous in angiosperms and gnetophytes. Vessels allow
the same cross-sectional area of wood to transport around
a hundred times more water than tracheids! This allowed
plants to fill more of their stems with structural fibres, and
also opened a new niche to vines, which could transport
water without being as thick as the tree they grew on.
Despite these advantages, tracheid-based wood is a lot
lighter, thus cheaper to make, as vessels need to be much
more reinforced to avoid cavitation.
EVOLUTION OF LEAVES

Leaves today are, in almost all instances, an adaptation to


increase the amount of sunlight that can be captured for
photosynthesis. Leaves certainly evolved more than once,
and probably originated as spiny outgrowths to protect
early plants from herbivory.
The rhyniophytes of the Rhynie chert comprised
nothing more than slender, unornamented axes. The early
to middle Devonian trimerophytes, therefore, are the first
evidence we have of anything that could be considered
leafy. This group of vascular plants are recognisable by
92 Plant Ecology

their masses of terminal sporangia, which adorn the ends


of axes which may bifurcate or trifurcate. Some organisms,
such as Psilophyton bore enations. These are small, spiny
outgrowths of the stem, lacking their own vascular supply.
Around the same time, the zosterophyllophytes were
becoming important. This group is recognisable by their
kidney-shaped sporangia, which grew on short lateral
branches close to the main axes. They sometimes branched
in a distinctive H-shape. The majority of this group bore
pronounced spines on their axes. However, none of these
had a vascular trace, and the first evidence of vascularised
enations occurs in the Rhynie genus Asteroxylon. The
spines of Asteroxylon had a primitive vasuclar supply-at
the very least, leaf traces could be seen departing from the
central protostele towards each individual "leaf".
A fossil known as Baragwanathia appears in the fossil
record slightly earlier, in the late Silurian. In this organism,
these leaf traces continue into the leaf to form their mid-
vein. One theory, the "enation theory", holds that the leaves
developed by outgrowths of the protostele connecting with
existing enations, but it is also possible that microphylls
evolved by a branching axis forming "webbing".
Asteroxylon and Baragwanathia are widely regarded as
primitive lycopods. The lycopods are still extant today,
familiar as the quillwort Isoetes and the club mosses.
Lycopods bear distinctive microphylls-Ieaves with a
single vascular trace. Microphylls could grow to some
size-the Lepidodendrales boasted microphylls over a
meter in length-but almost all just bear the one vascular
bundle. The branching pattern of megaphyll veins may
belie their origin as webbed, dichotomising branches.
The more familiar leaves, megaphylls, are thought to
have separate origins-indeed, they appeared four times
independently, in the ferns, horsetails, progymnosperms,
and seed plants. They appear to have originated from
dichotomising branches, which first overlapped one
Ecological Evolution of Plants 93

another, and eventually developed "webbing" and evolved


into gradually more leaf-like structures. So megaphyUs, by
this "teleome theory", are composed of a group of webbed
branches-hence the "leaf gap" left where the leaf's vascular
bundle leaves that of the main branch resembles two axes
splitting. In each of the four groups to evolve megaphylls,
their leaves first evolved during the late Devonian to early
Carboniferous, diversifying rapidly until the designs
settled down in the mid Carboniferous.
The cessation of further diversification can be attributed
to developmental constraints, but why did it take so long
for leaves to evolve in the first place? Plants had been on
the land for at least 50 million years before megaphylls
became significant. However, smal1, rare mesophylls are
known from the early Devonian genus Eophyllophyton-
so development could not have been a barrier to their
appearance.
The best explanation so far incorporates observations
that atmospheric CO2 was declining rapidly during this
time-falling by around 90% during the Devonian. This
corresponded with an increase in stomatal density by 100
times. Stomata allow water to evaporate from leaves, which
causes them to curve. It appears that the low stomatal
density in the early Devonian meant that evaporation was
limited, and leaves would overheat if they grew to any size.
The stomatal density could not increase, as the primitive
steles and limited root systems would not be able to supply
water quickly enough to match the rate of transpiration.
Clearly, leaves are not always beneficial, as illustrated
by the frequent occurrence of secondary loss of leaves,
famously exemplified by cacti and the "whisk fern"
Psilotum.
Secondary evolution can also disguise the true
evolutionary origin of some leaves. Some genera of ferns
display complex leaves which are attached to the
pseudostele by an outgrowth of the vascular bundle,
94 Plant Ecology

leaving no leaf gap. Further, horsetail (Equisetum) leaves


bear only a single vein, and appear for all the world to be
microphyllous; however, in the light of the fossil record and
molecular evidence, we conclude that their forbears bore
leaves with complex venation, and the current state is a
result of secondary simplification.
Deciduous trees deal with another disadvantage to
having leaves. The popular belief that plants shed their
leaves when the days get too short is misguided; evergreens
prospered in the Arctic circle during the most recent
greenhouse earth. The generally accepted reason for
shedding leaves during winter is to cope with the
weather-the force of wind and weight of snow are much
more comfortably weathered without leaves to increase
surface area.
Seasonal leaf loss has evolved independently several
times and is exhibited in the ginkgoales, gymnosperms and
angiosperms. Leaf loss may also have arisen as a response
to pressure from insects; it may have been less costly to lose
leaves entirely during the winter or dry season than to
continue investing resources in their repair.
EVOLUTION OF TREES

The early Devonian landscape was devoid of vegetation


taller than waist height. Without the evolution of a robust
vascular system, taller heights could not be obtained. There
was, however, a constant evolutionary pressure to attain
greater J;1.eight. The most obvious advantage is the
harvestihg of more sunlight for photosynthesis-by
overshadowing competitors-but a further advantage is
present in spore distribution, as spores (and, later, seeds)
can be blown greater distances if they start higher. This may
be demonstrated by Prototaxites, thought to be a late
Silurian fungus reaching eight metres in height.
In order to attain arborescence, early plants needed to
develop woody tissue that would act as both support and
Ecological Evolution of Plants 95

water transport. To understand wood, we must know a


little of vascular behaviour. The stele of plants undergoing
"secondary growth" is surrounded by the vascular
cambium, a ring of cells which produces more xylem (on
the inside) and phloem (on the outside). Since xylem cells
comprise dead, lignified tissue, subsequent rings of xylem
are added to those already present, forming wood.
The first plants to develop this secondary growth, and
a woody habit, were apparently the ferns, and as early as
the middle Devonian one species, Wattieza, had already
reached heights of 8 m and a tree-like habit. Other clades
did not take long to develop a tree-like stature; the late
Devonian Archaeopteris, a precursor to gymnosperms
which evolved from the trimerophytes, reached 30 m in
height. These progymnosperms were the first plants to
develop true wood, grown from a bifacial cambium, of
which the first appearance is in the mid Devonian Rellimia.
True wood is only thought to have evolved once, giving rise
to the concept of a "lignophyte" clade.
These Archaeopteris forests were soon supplemented
by lycopods, in the form of lepidodendrales, which topped
Sam in height and 2m across at the base. These lycopods
rose to dominate late Devonian and Carboniferous coal
deposits.
Lepidodendrales differ from modern trees in exhibiting
determinate growth: after building up a reserve of nutrients
at a low height, the plants would "bolt" to a genetically
determined height, branch at that level, spread their spores
and die. They consisted of "cheap" wood to allow their
rapid growth, with at least half of their stems comprising
a pith-filled cavity. Their wood was also generated by a
unifacial vascular cambium-it did not produce new
phloem, meaning that the trunks could not grow wider
over time.
The horsetail Calamites was next on the scene,
appearing in the Carboniferous. Unlike the modern
96 Plant Ecology

horsetail Equisetum, Calamites had a unifacial vascular


cambium, allowing them to develop wood and grow to
heights in excess of 10 m. They also branched multiple
times. While the form of early trees was similar to that of
todays', the groups containing all modem trees had yet to
evolve.
The dominant groups today are the gymnosperms,
which include the coniferous trees, and the angiosperms,
which contain all fruiting and flowering trees. It was long
thought that the angiosperms arose from within the
gymnosperms, but recent molecular evidence suggests that
their living representatives form two distinct groups. It
must be noted that the molecular data has yet to be fully
reconciled with morphological data, but it is becoming
accepted that the morphological support for paraphyly is
not especially strong. This would lead to the conclusion that
both groups arose from within the pteridosperms, probably
as early as the Permian.
The angiosperms and their ancestors played a very
small role until they diversified during the Cretaceous.
They started out as small, damp-loving organisms in the
understory, and have been diversifying ever since the mid-
Cretaceous, to become the dominant member of non-boreal
forests today.
EVOLUTION OF ROOTS

Roots are important to plants for two main reasons: Firstly,


they provide anchorage to the substrate; more importantly,
they provide a source of water and nutrients from the soil.
Roots allowed plants to grow taller and faster.
The onset of roots also had effects on a global scale. By
disturbing the soil, and promoting its acidification (by
taking up nutrients such as nitrate and phosphate), they
enabled it to weather more deeply, promoting the draw-
down of CO2 with huge implications for climate. These
effects may have been so profound they led to a mass
Ecological Evolution of Plants 97

extinction. But how and when did roots evolve in the first
place? While there are traces of root-like impressions in
fossil soils in the late Silurian, body fossils show the earliest
plants to be devoid of roots. Many had tendrils which
sprawled alonp; or beneath the ground, with upright axes
or thalli dotted here and there, and some even had non-
photosynthetic subterranean branches which lacked
stomata.
The distinction between root and specialised branch is
developmental; true roots follow a different developmental
trajectory to stems. Further, roots differ in their branching
pattern, and in possession of a root cap. So while "Silu-
Devonian plants such as Rhynia and Horneophyton
possessed the physiological equivalent of roots, roots-
defined as organs differentiated from stems-did not arrive
until later. Unfortunately, roots are rarely preserved in the
fossil record, and our understanding of their evolutionary
origin is sparse.
Rhizoids-small structures performing the same role as
roots, usually a a cell in diameter-probably evolved very
early, perhaps even before plants colonised the land; they
are recognised in the Characeae, an algal sister group to
land plants. That said, rhizoids probably evolved more than
once; the rhizines of lichens, for example, perform a similar
role. Even some animals have root-like structures!
More advanced structures are common in the Rhynie
chert, and many other fossils of comparable early Devonian
age bear structures that look like, and acted like, roots. The
rhyniophytes bore fine rhizoids, and the trimerophytes and
herbaceous lycopods of the chert bore root-like structure
penetrating a few centimetres into the soil. However, none
of these fossils display all the features borne by modem
roots. Roots and root-like structures became increasingly
more common and deeper penetrating during the
Devonian period, with lycopod trees forming roots around
20 cm long during the Eifelian and Givetian. These were
98 Plant Ecology

joined by progymnosperms, which rooted up to about a


metre deep, during the ensuing Frasnian stage. True
gymnosperms and zygopterid ferns also formed shallow
rooting systems during the Famennian period.
The rhizomorphs of the lycopods provide a slightly
approach to rooting. They were equivalent to stems, with
organs equivalent to leaves performing the role of rootlets.
A similar construction is observed in the extant lycopod
Isoetes, and this appears to be evidence that roots evolved
independently at least twice, in the lycophytes and other
plants.
A vascular system is indispensable to a rooted plants,
as non-photosynthesising roots need a supply of sugars,
and a vascular system is required to transport water and
nutrients from the roots to the rest of the plant. These plants
are little more advanced than their Silurian forbears,
without a dedicated root system; however, the flat-lying
axes can be clearly seen to have growths similar to the
rhizoids of bryophytes today.
By the mid-to-Iate Devonian, most groups of plants had
independently developed a rooting system of some nature.
As roots became larger, they could support larger trees, and
the soil was weathered to a greater depth. This deeper
weathering had effects not only on the aforementioned
drawdown of CO2, but also opened up new habitats for
colonisation by fungi and animals.
Roots today have developed to the physical limits. They
penetrate many metres of soil to tap the water table. The
narrowest roots are a mere 40 pm in diameter, and could
not physically transport water if they were any narrower.
The earliest fossil roots recovered, by contrast, narrowed
,from 3 mm to under 700 pm in diameter; of course,
taphonomy is the ultimate control of what thickness we can
see.
The efficiency of many plants' roots is increased via a
symbiotic relationship with a fungal partner. The most
Ecological Evolution of Plants 99

common are arbuscular mycorrhizae (AM), literally "tree-


like fungal roots". These comprise fungi which invade some
root cells, filling the cell membrane with their hyphae. They
feed on the plant's sugars, but return nutrients generated
Dr extracted from the soil, which the plant would otherwise
have no access to.
This symbiosis appears to have evolved early in plant
history. AM are found in all plant groups, and 80% of extant
vascular plants, suggesting an early ancestry; a "plant"-
fungus symbiosis may even have been the step that enabled
them to colonise the land, and indeed AM are abundant in
the Rhynie chert; the association occurred even before there
were true roots to colonise, and is has even been suggested
that roots evolved in order to provide a more comfortable
habitat for mycorrhizal fungi.
EVOLUTION OF SEEDS

Early land plants reproduced in the fashion of ferns: spores


germinated into small gametophytes, which produced
sperm. These would swim across moist soils to find the
female organs on the same or another gametophyte, where
they would fuse with an ovule to produce an embryo,
which would germinate into a sporophyte.
This mode of reproduction restricted early plants to
damp environments, moist enough that the sperm could
swim to their destination. Therefore, early land plants were
constrained to the lowlands, near shores and streams. The
development of heterospory freed them from this
constraint. Heterosporic organisms, as their name suggests,
bear spores of two sizes-microspores and megaspores.
These would germinate to form microgametophytes and
megagametophytes, respectively. This system paved the
way for seeds: taken to the extreme, the megasporangia
could bear only a single megaspore tetrad, and to complete
the transition to true seeds, three of the megaspores in the
original tetrad cold be aborted, leaving one megaspore per
megasporangium.
100 Plant Ecology

The transition to seeds continued with this megaspore


being "boxed in" to its sporangium while it germiates. Then,
the megagametophyte is contained within a waterproof
integuement, which forms the bulk of the seed. The
microgametophyte-a pollen grain which has germinated
from a microspore-is employed for dispersal, only
releasing its desiccation-prone sperm when it reaches a
receptive microgametophyte.
Lycopods go a fair way down the path to seeds without
ever crossing the threshold. Fossil lycopod megaspores
reaching 1 cm in diameter, and surrounded by vegitative
tissue, are known-these even germinate into a
megagametophyte in situ. However, they fall short of being
seeds, since the nucellus, an inner spore-covering layer,
does not completely enclose the spore. A very small sJit
remains, meaning that the seed is still exposed to the
atmosphere. This has two consequences-firstly, it means
it is not fully resistant to desiccation, and secondly, sperm
do not have to "burrow" to access the archegonia of the
megaspore.
The first "spermatophytes"-that is, the first plants to
bear true seeds-were progymnosperms called
pteridosperms: literally, "seed ferns". They ranged from
trees to small, rambling shrubs; like most early
progymnosperms, they were woody plants with fern-like
foliage. They all bore ovules, but no cones, fruit or similar.
While it is difficult to track the early evolution of seeds, we
can trace the lineage of the seed ferns from the simple
trimerophytes through homosporous Aneruophytes.
This seed model is shared by basically all gymnosperms
("naked seeds"), most of which encase their seeds in a
woody or fleshy cone, but none of which fully enclose their
seeds. The angiosperms ("vessel seeds") are the only group
to fully enclose the seed, in a carpel.
Fully enclosed seeds opened up a new pathway for
plants to follow: that of seed dormancy. The embryo,
Ecological Evolution of Plants 101

completely isolated from the external atmosphere and


hence protected from desiccation, could survive some years
of draught before germinating.
Gymnosperm seeds from the late Carboniferous have
been found to contain embryos, suggesting a lengthy gap
between fertilisation and germination. This period is
associated with the entry into a greenhouse earth period,
with an associated increase in aridity. This suggests that
dormancy arose as a response to drier climatic conditions,
where it became advantageous to wait for a moist period
before germinating. This evolutionary breakthrough
appears to have opened a floodgate: previously
inhospitable areas, such as dry mountain slopes, could now
be tolerated, as were soon covered by trees.
Seeds offered further advantages to their bearers: they
increased the success rate of fertilised gametophytes, and
because a nutrient store could be "packaged" in with the
embryo, the seeds could germinate rapidly in inhospitable
environments, reaching a size where it could fend for itself
more quickly. For example, without an endosperm,
seedlings growing in arid environments would not have
the reserves to grow roots deep enough to reach the water
table before they expired. Likewise, seeds germinating in a
gloomy understory require an additional reserve of energy
to quickly grow high enough to capture sufficient light for
self-sustenance. A combination of these advantages gave
seed plants the ecological edge over the previously
dominant genus Archaeopteris, this increasing the
biodiversity of early forests.
EVOLUTION OF FLOWERS

Flowers are organs possessed only by the group known as


the angiosperms, a relatively late appearance on the
evolutionary scene. Colourful and/or pungent structures
surround the cones of plants such as cycads and gnetales,
making a strict definition of the term "flower" elusive.
102 Plant Ecology

The flowering plants have long been assumed to have


evolved from within the "gymnosperms"; according to the
traditional morphological view, they are closely allied to
the gnetales. However, as noted above, recent molecular
evidence is at odds to this hypothesis, and further suggests
that gnetales are more closely related to some gymnosperm
groups than angiosperms, and that extant gymnosperms
form a distinct clade to the angiosperms,
The relationship of stem groups to the angiosperms is
of utmost importance in determining the evolution of
flowers; stem groups provide an insight into the state of
earlier "forks" on the path to the current state, If we identify
an unrelated group as a stem group, then we will gain an
incorrect image of the lineages' history, The traditional
view that flowers arose by modification of a structure
similar to that of the gnetales, for example, no longer bears
weight in the light of the molecular data.
Convergence increases our chances of misidentifying
stem groups. Since the protection of the megagametophyte
is evolutionarily desirable, it would be unsurprising if
many separate groups stumbled upon protective
encasements independently. Distinguishing ancestry in
such a situation, especially where we usually only have
fossils to go on, is tricky-to say the least, In flowers, this
protection is offered by the carpel, an organ believed to
represent an adapted leaf, recruited into a protective role,
shielding the ovules. These ovules are further protected by
a double-walled integument.
Penetration of these protective layers needs something
more that a free-floating microgametophyte. Angiosperms
have pollen grains comprising just three cells, One cell is
responsible for drilling down through the integuments, and
creating a conduit for the two sperm cells to flow down. The
megagametophyte has just seven cells; of these, one fuses
with a sperm cell, forming the nucleus of the egg itself, and
another other joins with the other sperm, and dedicates
Ecological Evolution of Plants 103

itself to forming a nutrient-rich endosperm. The other cells


take auxilIary roles. This process of "double fertilisation" is
unique and common to all angiosperms. In the fossil record,
there are three intriguing groups which bore flower-like
structures. The first is the Permian pteridosperm
Glossopteris, which already bore recurved leaves
resembling carpels. The Triassic Caytonia is more flower-
like still, with enclosed ovules-but only a single
integument. Further, details of their pollen and stamens set
them apart from true flowering plants.
The Bennettitales bore remarkably flower-like organs,
protected by whorls of bracts which may have played a
similar role to the petals and sepals of true flowers.
However, no true flowers are found in any groups save
those extant today. Most morphological and molecular
analyses place Amborella, the nymphaeales and
Austrobaileyaceae in a basal clade dubbed "ANA". This
clade appear to have diverged in the early Cretaceous,
around 130 million years ago--around the same time as the
earliest fossil angiosperm, and just after the first
angiosperm-like pollen, 136 million years ago. The
magnoliids diverged soon after, and a rapid radiation had
produced eudicots and monocots by 125 million years ago.
ADVANCES IN METABOLISM

The. most recent major innovation by the plants is the


development of the C4 metabolic pathway. Photosynthesis
is not quite as simple as adding water to CO2 to produce
sugars and oxygen. A complex chemical pathway is
involved in this miraculous reaction, facilitated along the
way by a range of enzymes and co-enzymes. The enzyme
RuBisCO is responsible for "fixing" CO2-that is, it attaches
it to a carbon-based molecule to form a sugar, which can
be used by the plant, releasing an oxygen molecule along
the way. However, the enzyme is notoriously inefficient,
and works just as effectively in the other direction through
104 Plant Ecology

a process known as photorespiration. As well as a hot-off-


the press sugar molecule, this also costs the plant energy
required to "re-set" the RuBisCO molecule,

Concentrating Carbon
To work around this inefficiency, C4 plants developed
"carbon concentrating" mechanisms, These work by
bombarding RuBisCO molecules with CO2, thereby
increasing the amount of time they are performing the
useful task of making sugars. The process of bombarding
the RuBisCO requires more energy than allowing gasses to
come and go where they please, but under the right
conditions-Leo warm temperatures, low CO 2
concentrations, or high oxygen concentrations-pays off in
terms of the decreased loss of sugar through
photorespiration. One, C 4 metabolism, employs a so-called
Kranz anatomy. This transports CO2 through an outer
mesophyll layer, via a range of organic molecules, to a
central bundle sheath cell, where, the CO2 is released. In
this way, CO2 is concentrated near the site of RuBisCO
operation. Because RuBisCO is operating in an
environment with much more CO2 than it otherwise would
be, it performs more efficiently.
A second method, CAM photosynthesis, temporally
separates photosynthesis from the action of RuBisCO.
RuBisCO only operates during the day, when stomata are
sealed and CO2 is provided by the breakdown of the
chemical malate. More CO2 is then harvested from the
atmosphere when stomata open, during the cool, moist
nights, reducing water loss.
These two pathways, with the same effect on RuBisCO,
evolved a number of times independently-indeed, C 4
alone arose 18 times! The C4 construction is most famously
used by a subset of grasses, while CAM is employed by
many succulents and cacti. The trait appears to have
emerged during the Oligocene, around 25 to 32 million
Ecological Evolution of Plants lOS

years ago; however, they did not become ecologically


significant until the Miocene, 6-7 million years ago.
Remarkably, some charcoalified fossils preserve tissue
organised into the Kranz anatomy, with intact bundle
sheath cells, allowing the presence C4 metabolism to be
identified without doubt at this time.
In deducing their distribution and significance, we
resort to the use of isotopic markers. C3 plants preferentially
use the lighter of two isotopes of carbon in the atmosphere,
12C, which is more readily involved in the chemical
pathways involved in its fixation. Because C4 metabolism
involves a further chemical step, this effect is accentuated.
Plant material can be analysed to deduce the ratio of the
heavier 13C to 12(:. C3 plants are on average around 120/00
lighter than the atmospheric ratio, while C4 plants are about
270/00 lighter. It's troublesome procuring original fossil
material in sufficient quantity to analyse the grass itself, but
fortunately we have a good proxy: horses. Horses were
globally widespread in the period of interest, and browsed
almost exclusively on grasses. There's an old phrase in
isotope palreontology, "you are what you eat (plus a little
bit),,-this refers to the fact that organisms reflect the
isotopic composition of whatever they eat, plus a small
adjustment factor. There is a good record of horse teeth
throughout the globe, and their d13C has been measured.
The record shows a sharp negative inflection around 6-7
million years ago, during the Messinian, and this is
interpreted as the rise of C4 plants on a global scale.
While C 4 enhances the efficiency of RuBisCO, the
concentration of carbon is highly energy intensive. This
means that C4 plants only have an advantage over C 3
organisms in certain conditions: namely, high
temperatures, low CO2 and high oxygen concentrations. C4
plants also need high levels of sunlight in order to thrive.
Models suggest that without wildfires removing shade-
casting trees and shrubs, there would be no space for C4
106 Plant Ecology

plants. But wildfires have occurred for 400 million years-


.why did C 4 take so long to arise, and then appear
independently so many times? The Carboniferous period
had notoriously high oxygen levels-almost enough to
allow spontaneous combustion-and very low CO2, but
there is no C4 isotopic signature to be found. And there
doesn't seem to be a sudden trigger for the Miocene rise.
During the Micoene, the atmosphere and climate was
relatively stable. If anything, it increased gradually from 14-
9 million years ago before settling down to concentrations
similar to the Holocene. This suggests that it did not have
a key role in invoking C4 evolution. Grasses themselves had
probably been around for 60 million years or more, so had
had plenty of time to evolve C4, which in any case is present
in a diverse range of groups and thus evolved
independentl y.
There is a strong signal of climate change in South Asia;
increasing aridity-hence increasing fire frequency and
intensity-may have led to an increase in the importance
of grasslands. However, this is difficult to reconcile with the
North American record. It is possible that the signal is
entirely biological, forced by the fire- driven acceleration of
grass evolution-which, both by increasing weathering and
incorporating more carbon into sediments, reduced
atmospheric CO2 levels.
REFERENCES

Raven, J.A.; Edwards, D."Roots: evolutionary ongms and


biogeochemical significance" . Journal of Experimental Botany 52 (90001):
381-401. 2001.
Kenrick, P., Crane P.R., The origin and early diversification of land plants.
A cladistic study. Smithsonian Institution Press, Washington & London.
1997.
Gray, J. "The Microfossil Record of Early Land Plants: Advances in
Understanding of Early Terrestrialization, 1970-1984". Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences
(1934-1990) 309 (1138): 167-195. 1985,
5
Ecology of Fungi

The fungi are heterotrophic organisms possessing a


chitinous cell wall. The majority of species grow as
multicellular filaments called hyphae forming a mycelium;
some fungal species also grow as single cells. Sexual and
asexual reproduction of the fungi is commonly via spores,
often produced on specialized structures or in fruiting
bodies. Some species have lost the ability to form
specialized reproductive structures, and propagate solely
by vegetative growth. Yeasts, molds, and mushrooms are
examples of fungi. The fungi are a monophyletic group that
is phylogenetically clearly distinct from the
morphologically similar slime molds (myxomycetes) and
water molds (oomycetes). The fungi are more closely
related to animals than plants, yet the discipline of biology
devoted to the study of fungi, known as mycology, often
falls under a branch of botany.
Occurring worldwide, most fungi are largely invisible
to the naked eye, living for the most part in soil, dead
matter, and as symbionts of plants, animals, or other fungi.
They perform an essential role in all ecosystems in
decomposing organic matter and are indispensable in
nutrient cycling and exchange. Some fungi become
noticeable when fruiting, either as mushrooms or molds.
Many fungal species have long been used as a direct source
of food, such as mushrooms and truffles and in
108 Plant Ecology

fermentation of various food products, such as wine, beer,


and soy sauce. More recently, fungi are being used as
sources for antibiotics used in medicine and various
enzymes, such as cellulases, pectinases, and proteases,
important for industrial use or as active ingredients of
detergents. Many fungi produce bioactive compounds
called mycotoxins, such as alkaloids and polyketides that
are toxic to animals including humans. Some fungi are used
recreationally or in traditional ceremonies as a source of '
psychotropic compounds. Several species of the fungi are
significant pathogens of humans and other animals, ' and
losses due to diseases of crops (e.g., rice blast disease) or
food spoilage caused by fungi can have a large impact on
human food supply and local economies.

Fruiting Body
'\

Figure 1. Basic structure of Q fungal body

The evolution of multicellular eukaryotes increased the size


and complexity of organisms, allowing them to exploit the
terrestrial habitat. Fungi first evolved in water but made the
transition to land through the development of specialized
structures that prevented their drying out. First classified
as plants, fungi are now considered different enough from
plants to be placed in a separate kingdom, and in fact are
more like animals than plants. . .
Ecology of Fungi 109

Fungi have a worldwide distribution, and grow in a


wide range of habitats, including deserts. Most fungi grow
in terrestrial environments, but several species occur only
in aquatic habitats. Fungi along with bacteria are the
primary decomposers of organic matter in most if not all
terrestrial ecosystems worldwide. Based on observations of
the ratio of the number of fungal species to the number of
plant species in some environments, the fungal kingdom
has been estimated to contain about 1.5 million species.
Around 70,000 fungal species have been formally described
by taxonomists, but the true dimension of fungal diversity
is still unknown.ost fungi grow as thread-like filaments
called hyphae, which form a mycelium, while others grow
as single cells. Until recently many fungal species were
described based mainly on morphological characteristics,
such as the size and shape of spores or fruiting structures,
and biological species concepts; the application of
molecular tools, such as DNA sequencing, to study fungal
diversity has greatly enhanced the resolution and added
robustness to estimates of diversity within various
taxonomic groups.
Human use of fungi for food preparation or
preservation and other purposes is extensive and has a long
history: yeasts are required for fermentation of beer, wine
and bread, some other fungal species are used in the
production of soy sauce and tempeh. Mushroom farming
and mushroom gathering are large industries in many
countries. Many fungi are producers of antibiotics,
including B-Iactam antibiotics such as penicillin and
cephalosporin. Widespread use of these antibiotics for the
treatment of bacterial diseases, such as tuberculosis,
syphilis, leprosy, and many others began in the early 20th
century and continues to playa major part in anti-bacterial
chemotherapy. The study of the historical uses and
sociological impact of fungi is known as ethnomycology.
110 Plant Ecology

ECOLOGY

Although often inconspicuous, fungi occur in every


environment on Earth and play very important roles in
most ecosystems. Along with bacteria, fungi are the major
decomposers in most terrestrial (and some aquatic)
ecosystems, and therefore play a critical role in
biogeochemical cycles and in many food webs. As
decomposers, they play an indispensable role in nutrient
cycling, especially as saprotrophs and symbionts,
degrading organic matter to inorganic molecules, which
can then re-enter anabolic metabolic pathways in plants or
other organisms.

Symbiosis
Many fungi have important symbiotic relationships with
organisms from most if not all Kingdoms. These
interactions can be mutualistic or antagonistic in nature, or
in case of commensal fungi are of no apparent benefit or
detriment to the host.
Mycorrhizal symbiosis between plants and fungi is one
of the most well-known plant-fungus associations and is of
significant importance for plant growth and persistence in
many ecosystems; over 90% of all plant species engage in
some kind of mycorrhizal relationship with fungi and are
dependent upon this relationship for survival. The
mycorrhizal symbiosis is ancient, dating to at least 400
million years ago. It often increases the plant's uptake of
inorganic compounds, such as nitrate and phosphate from
soils having low concentrations of these key plant nutrients.
In some mycorrhizal associations, the fungal partners may
mediate plant-to-plant transfer of carbohydrates and other
nutrients. Such mycorrhizal communities are called
"common mycorrhizal networks".
Lichens are formed by a symbiotic relationship between
algae or cyanobacteria and fungi, in which individual
Ecology of Fungi 111

photobiont cells are embedded in a tissue formed by the


fungus. As in mycorrhizas, the photobiont provides sugars
and other carbohydrates, while the fungus provides
minerals and water. The functions of both symbiotic
organisms are so closely intertwined that they function
almost as a single organism.
Many insects also engage in mutualistic relationships
with various types of fungi. Several groups of ants cultivate
fungi in the order Agaricales as their primary food source,
while ambrosia beetles cultivate various species of fungi in
the bark of trees that they infest. Termites on the African
Savannah are also known to cultivate fungi.

Fungi as Pathogens and Parasites


However, many fungi are parasites on plants, animals
(including humans), and other fungi. Serious fungal
pathogens of many cultivated plants causing extensive
damage and losses to agriculture and forestry include the
rice blast fungus Magnaporthe oryzae, tree pathogens such
as Ophiostoma ulmi and Ophiostoma novo-ulmi causing
Dutch elm disease, and Cryphonectria parasitica
responsible for chestnut blight, and plant-pathogenic fungi
in the genera Fusarium, Ustilago, Alternaria, and
Cochliobolus; fungi with the potential to cause serious
human diseases, especially in persons with immuno-
deficiencies, are in the genera Aspergillus, Candida,
Cryptoccocus, Histoplasma, and Pneumocystis. Several
pathogenic fungi are also responsible for relatively minor
human diseases, such as athlete's foot and ringworm. Some
fungi are predators of nematodes, which they capture using
an array of specialized structures, such as constricting rings
or adhesive nets.
Nutrition and Autotrophy
Growth of fungi as hyphae on or in solid substrates or
single cells in aquatic enviTonments is adapted to efficient
112 Plant Ecology

extraction of nutrients from these environments, because


these growth forms have high surface area to volume ratios.
These adaptations in morphology are complemented by
hydrolytic enzymes secreted into the environment for
digestion of large organic molecules, such as
polysaccharides, proteins, lipids, and other organic
substrates into smaller molecules. These molecules are then
absorbed as nutrients into the fungal cells.
Traditionally, the fungi are considered heterotrophs,
organisms that rely solely on carbon fixed by other
organisms for metabolism. Fungi have evolved a
remarkable metabolic versatility that allows many of them
to use a large variety of organic substrates for growth,
including simple compounds as nitrate, ammonia, acetate,
or ethanol. Recent research raises the possibility that some
fungi utilize the pigment melanin to extract energy from
ionizing radiation, such as gamma radiation for
"radiotrophic" growth. It has been proposed that this
process might bear some similarity to photosynthesis in
plants, but detailed biochemical data supporting the
existence of this hypothetical pa4-hway are presently
lacking.
MORPHOLOGY

Microscopic Structures
Though fungi are part of the opisthokont clade, all phyla
except for the chytrids have lost their posterior flagella.
Fungi are unusual among the eukaryotes in having a cell
wall that, besides glucans and other typical components,
contains the biopolymer chitin. Many fungi grow as thread-
like filamentous microscopic structures called hyphae, and
an assemblage of intertwined and interconnected hyphae
is called a mycelium. Hyphae can be septate, i.e., divided
into hyphal compartments separated by a septum, each
compartment containing one or more nuclei or can be
coenocytic, i.e., lacking hyphal compartmentalization.
Ecology of Fungi 113

However, septa have pores, such as the doliporus in the


basidiomycetes that allow cytoplasm, organelles, and
sometimes nuclei to pass through. Coenocytic hyphae are
essentially multinucleate supercells. In some cases, fungi
have developed specialized structures for nutrient uptake
from living hosts; examples include haustoria in plant-
parasitic fungi of nearly all divisions, and arbuscules of
several mycorrhizal fungi, which penetrate into the host
cells for nutrient uptake by the fungus ..

Macroscopic Structures
Fungal mycelia can become visible macroscopically, for
example, as concentric rings on various surfaces, such as
damp walls, and on other substrates, such as spoilt food,
and are commonly and generically called mould; fungal
mycelia grown on solid agar media in laboratory petri
dishes are usually referred to as colonies, with many species
exhibiting characteristic macroscopic growth morphologies
and colours, due to spores or pigmentation.

PUeulI-_~"

YOUNG STAGE

Figure 2. Structure of a Toadstool Fungus

Specialized fungal structures important in sexual


reproduction are the apothecia, perithecia, and c1eistotheda
114 Plant Ecology

in the ascomycetes, and the fruiting bodies of the


basidiomycetes, and a few ascomycetes. These reproductive
structures can sometimes grow very large, and are well
known as mushrooms.

Structures for Substrate Penetration


Fungal hyphae are specifically adapted to growth on solid
surfaces and within substrates, and can exert astoundingly
large penetrative mechanical forces. The plant pathogen,
Magnaporthe grise a, forms a structure called an
appressorium specifically designed for penetration of plant
tissues, and the pressure generated by the appressorium,
which is directed against the plant epidermis can exceed 8
MPa. The generation of these mechanical pressures is the
result of an interplay between physiological processes to
increase intracellular turgor by production of osmolytes
such as glycerol, and the morphology of the appressorium.
REPRODUCI10N

Reproduction of fungi is complex, reflecting the


heterogeneity in lifestyles and genetic make up within this
group of organisms. Many fungi reproduce either sexually
or asexually, depending on conditions in the environment.
These conditions trigger genetically determined
developmental programs leading to the expression of
specialized structures for sexual or asexual reproduction.
These structures aid both reproduction and efficient
dissemination of spores or spore-containing propagules.

Asexual Reproduction
Asexual reproduction via vegetative spores or through
mycelial fragmentation is common in many fungal species
and allows more rapid dispersal than sexual reproduction.
In the case of the "Fungi imperfecti" or Deuteromycota,
which lack a sexual cycle, it is the only means of
propagation. Asexual spores, upon germination, may
Ecology of Fungi 115

found a population that is clonal to the population from


which the spore originated, and thus colonize new
environments.

Sexual Reproduction
Sexual reproduction with meiosis exists in all fungal phyla,
except the Deuteromycota. It differs in many aspects from
sexual reproduction in animals or plants. Many differences
also exist between fungal groups and have been used to
discrimina te fungal clades and species based on
morphological differences in sexual structures and
reproductive strategies. Experimental crosses between
fungal isolates can also be used to identify species based on
biological species concepts.
The major fungal clades have initially been delineated
based on the morphology of their sexual structures and
spores; for example, the spore-containing structures, asci
and basidia, can be used in the identification of ascomycetes
and basidiomycetes, respectively. Many fungal species
have elaborate vegetative incompatibility systems that
allow mating only between individuals of opposite mating
type, while others can mate and sexually reproduce with
any other individual or itself. Species of the former mating
system are called heterothallic, and of the latter
homothallic.
Most fungi have both a haploid and diploid stage in
their life cycles. In all sexually reproducing fungi,
compatible individuals combine by cell fusion of vegetative
hyphae by anastomosis, required for the initiation of the
sexual cycle. Ascomycetes and basidiomycetes go through
a dikaryotic stage, in which the nuclei inherited from the
two parents do not fuse immediately after cell fusion, but
remain separate in the hyphal cells.
In ascomycetes, dikaryotic hyphae of the hymenium
form a characteristic hook at the hyphal septum. During cell
116 Plant Ecology

division formation of the hook ensures proper distribution


of the newly divided nuclei into the apical and basal hyphal
compartments. An ascus (plural asci) is then formed, in
which karyogamy (nuclear fusion) occurs. These asci are
embedded in an ascocarp, or fruiting body, of the fungus.
Karyogamy in the asci is followed immediately by meiosis
and the production of ascospores. The ascospores are
disseminated and germinate and may form a new haploid
mycelium.
Sexual reproduction in basidiomycetes is similar to that
of the ascomycetes. Compatible haploid hyphae fuse to
produce a dikaryotic mycelium. However, the dikaryotic
phase is more extensive in the basidiomycetes, in many
cases also present in the vegetatively growing mycelium. A
specialized anatomical structure, called a clamp connection,
is formed at each hyphal septum.
As with the structurally similar hook in the
ascomycetes, formation of the clamp connection in the
baSidiomycetes is required for controlled transfer of nuclei
during cell division, to maintain the dikaryotic stage with
two genetically different nuclei in each hypha I
compartment. A basidiocarp is formed in which club-like
structures known as basidia generate haploid basidiospores
after karyogamy and meiosis. The most commonly known
basidiocarps are mushrooms, but they may also take many
other forms.
In zygomycetes, haploid hyphae of two individuals
fuse, fOrmlng a zygote, which develops into a zygospore.
When the zygospore germinates, it quickly undergoes
meiosis, generating new haploid hyphae, which in tum
may form asexual sporangiospores. These sporangi6spores
are means of rapid dispersal of the fungus and germinate
into new genetically identical haploid fungal colonies, able
to mate and undergo another sexual cycle followed by the
generation of new zygospores, thus completing the
lifecycle.
Ecology of Fungi 117

Dispersal of Spores
Both asexual and sexual spores or sporangiospores of many
fungal species are actively dispersed by forcible ejection
from their reproductive structures. This ejection ensures
exit of the spores from the reproductive structures as well
as travelling through the air over long distances. Many
fungi thereby possess specialized mechanical and
physiological mechanisms as well as spore-surface
structures, such as hydrophobins, for spore ejection. These
mechanisms include, for example, forcible discharge of
ascospores enabled by the structure of the ascus and
accumulation of osmolytes in the fluids of the ascus that
lead to explosive discharge of the ascospores into the air.
The forcible discharge of single spores term~d
ballistospores involves formation of a small drop of water
(Buller's drop), which upon contact with the spore leads to
its projectile release with an initial acceleration of more than
10,000 g. Other fungi rely on alternative mechanisms for
spore release, such as external mechanical forces,
exemplified by puffballs. Attracting insects, such as flies, to
fruiting structures, by virtue of their having lively colours
and a putrid odour, for dispersal of fungal spores is yet
another strategy, most prominently used by the stinkhorns.
Besides regular sexual reproduction with meiosis, some
fungal species may exchange genetic material via
parasexual processes, initiated by anastomosis between
hyphae and plasmogamy of fungal cells. The frequency and
relative importance of parasexual events is unclear and may
be lower than other sexual processes. However, it is known
to playa role in intraspecific hybridization and is also likely
required for hybridization between fungal species, which
has been associated with major events in fungal evolution.
PHYLOGENY AND CLASSIFICATION

For a long time taxonomists considered fungi to be


118 Plant Ecology

members of the Plant Kingdom. This early classification


was based mainly on similarities in lifestyle: both fungi and
plant are mainly sessile, have similarities in general
morphology and growth habitat. Moreover, both groups
possess a cell wall, which is absent in the Animal Kingdom.
However, the fungi are now considered a separate
kingdom, distinct from both plants and animals, from
which they appear to have diverged approximately one
billion years ago. Many studies have identified several
distinct morphological, biochemical, and genetic features in
the Fungi, clearly delineating this group from the other
kingdoms. For these reasons, the fungi are placed in their
own kingdom.
Similar to animals and unlike most plants, fungi lack the
capacity to synthesize organic carbon by chlorophyll-based
photosynthesis; whereas plants store the reduced carbon as
starch, fungi, like animals and some bacteria, use glycogen
for storage of carbohydrates.
A major component of the cell wall in many fungal
species is the nitrogen-containing carbohydrate, chitin, also
present in some animals, such as the insects and
crustaceans, while the plant cell wall consists chiefly of the
carbohydrate cellulose. The defining and unique
characteristics of fungal cells include growth as hyphae,
which are microscopic filaments of between 2-10 microns
in diameter and up to several centimetres In length, and
which combined form the fungal mycelium. Some fungi,
such as yeasts, grow as single ovoid cells, similar to
unicellular algae and the protists.
Unlike many plants, most fungi lack an efficient
vascular system, such as xylem or phloem for long-distance
transport of water and nutrients; as an example for
convergent evolution, some fungi, such as Armillaria, form
rhizomorphs or mycelial cords, resembling and
functionally related to, but morphologically distinct from,
plant roots.
Ecology of Fungi 119

Some characteristics shared between plants and. fungi


include the presence of vacuoles in the cell, and a similar
pathway in the biosynthesis of terpenes using mevalonic
acid and pyrophosphate as biochemical precursors; plants
however use an additional terpene biosynthesis pathway in
the chloroplasts that is apparently absent in fungi.
Ancestral traits shared among members of' the fungi
include chitinous cell walls and heterotrophy by
absorption. A further characteristic of the fungi that is
absent from other eukaryotes, and shared only with some
bacteria, is the biosynthesis of the amino acid, L~lysine, via
the a-aminoadipate pathway.
Similar to plants, fungi produce a plethora of secondary
metabolites functioning as defensive compounds or for
niche adaptation; however, biochemical pathways for the
synthesis of similar. or even identical compounds often
differ markedly between fungi and plants.
EVOLUTIONARY HISTORY

The first organisms having features typical of fungi date to


1,200 million years ago, the Proterozoic. However, fungal
fossils do not become common and uncontroversial until
the early Devonian, when they are abundant in the Rhynie
chert. Even though traditionally included in many botany
. curricula and textbooks, fungi are now thought to be more
closely related to animals than to plants and are placed with
the animals in the monophyletic group of opisthokonts. For
much of the Paleozoic Era, the fungi appear to have been
aquatic, and consisted of organisms similar to the extant
Chytrids in having flagellum-bearing spores. The early
fossil record of the fungi is fragmentary, to say the least. The
fungi probably colonized the land during the Cambrian,
long before land plants.
For some time after the Permian-Triassic extinction
event, a fungal spike, originally thought to be an
extraordinary abundance of fungal spores in sediments
120 Plant Ecology

formed shortly after this event, sugf', ~sted that they were
the dominant life form during this period-nearly 100% of
the fossil record available from this period. However, the
relative proportion of fungal spores relative to spores
formed by algal species is difficult to assess, the spike did
not appear world-wide, and in many places it did not fall
on the Permian-Triassic boundary.
Analyses using molecular phylogenetics support a
monophyletic origin of the Fungi. The taxonomy of the
Fungi is in a state of constant flux, especially due to recent
research based on DNA comparisons. These current
phylogenetic analyses often overturn classifications based
on older and sometimes less discriminative methods based
on morphological features and biological species concepts
obtained from experimental matings.
There is no unique generally accepted system at the
higher taxonomic levels and there are constant name
changes at every level, from species upwards. However,
efforts among fungal researchers are now underway to
establish and encourage usage of a unified and more
consistent nomenclature. Fungal species can also have
multiple scientific names depending on its life cycle and
mode (sexual or asexual) of reproduction. Web sites such
as Index Fungorum and ITIS define preferred up-to-date
names, but do not always agree with each other.

Taxonomic Groups
The major divisions (phyla) of fungi have been classified
based mainly on their sexual reproductive structures.
Currently, seven fungal divisions are proposed:

Chytridiomycota
The Chytridiomycota are commonly known as chytrids.
These fungi are ubiquitous with a worldwide distribution;
chytrids produce zoospores that are capable of active
Ecology of Fungi 121

movement through aqueous phases with a single flagellum.


Consequently, some taxonomists had earlier classified
them as protists on the basis of the flagellum. Molecular
phylogenies, inferred from the rRNA-operon sequences
representing the 185, 285, and 5.85 ribosomal subunits,
suggest that the Chytrids are a basal fungal group divergent
from the other fungal divisions, consisting of four major
clades with some evidence for paraphyly or possibly
polyphyly.

Blastocladiomycota
The Blastocladiomycota were previously considered a
taxonomic clade within the Chytridiomycota. Recent
molecular data and ultrastructural characteristics, however,
place the Blastocladiomycota as a sister clade to the
Zygomycota, Glomeromycota, and Dikarya (Ascomycota
and Basiomycota). The blastodadiomycetes are fungi that
are saprotrophs and parasites of all eukaryotic groups and
undergo sporic meiosis unlike their close relatives, the
chytrids, which mostly exhibit zygotic meiosis.

Neocallimastigomycota
The Neocallimastigomycota were earlier placed in the
phylum Chytridomycota. Members of this small phylum
are anaerobic organisms, living in the digestive system of
larger herbivorous mammals and possibly in other
terrestrial and aquatic environments. They lack
mitochondria but contain hydrogenosomes of
mitochondrial origin.

Zygomycota
The Zygomycota contain the taxa, Zygomycetes and
Trichomycetes, and reproduce sexually with meiospores
called zygospores and asexually with sporangiospores.
Black bread mold is a common species that belongs to this
122 Plant Ecology

group; another is Pilobolus, which is capable of ejecting


spores several meters through the air, Medically relevant
genera include Mucor, Rhizomucor, and Rhizopus.
Molecular phylogenetic investigation has shown the
Zygomycota to be a polyphyletic phylum with evidence of
paraphyly within this taxonomic group.

Glomeromycota
Members of the Glomeromycota are fungi forming
arbuscular mycorrhizae with higher plants, Only one
species has been observed forming zygospores; all other
species solely reproduce asexually. The symbiotic
association between the Glomeromycota and plants is
ancient, with evidence dating to 400 million years ago,

Ascomycota
The Ascomycota, commonly known as sac fungi or
ascomycetes, constitute the largest taxonomic group within
the Eumycota. These fungi form meiotic spores called
ascospores, which are enclosed in a special sac-like
structure called an ascus. This division includes morels, a
few mushrooms and truffles, single-celled yeasts (e.g., of
the genera Saccharomyces, Kluyveromyces, Pichia, and
Candida), and many filamentous fungi living as
saprotrophs, parasites, and mutualistic symbionts.
Prominent and important genera of filamentous
ascomycetes include Aspergillus, Penicillium, Fusarium,
and Claviceps. Many ascomycetes species have only been
observed undergoing asexual reproduction (called
anamorphic species), but molecular data has often been
able to identify their closest teleomorphs in the
Ascomycota. Because the products of meiosis are retained
within the sac-like ascus, several ascomyctes have been
used for elucidating prine ·-,les of genetics and heredity (e.g.
:'\Jeurospora crassa).
Ecology of Fungi 123

Basidiomycota
Members of the Basidiomycota, commonly known as the
club fungi or basidiomycetes, produce meiospores called
basidiospores on club-like stalks called basidia. Most
common mushrooms belong to this group, as well as rust
(fungus) and smut fungi, which are major pathogens of
grains. Other important Basidiomyces include the maize
pathogen,Ustilago maydis, human commensal species of
the genus Malassezia, and the opportunistic human
pathogen, Cryptococcus neoformans.
PHYLOGENETIC RELATIONSHIPS

Because of some similarities in morphology and lifestyle,


the slime molds (myxomycetes) and water molds
(oomycetes) were formerly classified in the kingdom Fungi.
Unlike true fungi, however, the cell walls of these
organisms contain cellulose and lack chitin. Slime molds
are unikonts like fungi, but are grouped in the Amoebozoa.
Water molds are diploid bikonts, grouped in the
Chromalveolate kingdom. Neither water molds nor slime
molds are closely related to the true fungi, and, therefore,
taxonomists no longer group them in the kingdom Fungi.
Nonetheless, studies of the oomycetes and myxomycetes
are still often included in mycology textbooks and primary
research literature.
REFERENCES

Alexopoulos, c.J., Charles W. Mims, M. Blackwell et al., Introductory


Mycology, 4th ed. John Wiley and Sons, Hoboken NJ, 2004.
Arora, David. (1986). "Mushrooms Demystified: A Comprehensive
Guide to the Fleshy Fungi". 2nd ed. Ten Speed Press.
Barea JM, Pozo MJ, Azc6n R, Azc6n-Aguilar C. "Microbial co-operation
in the rhizosphere". J. Exp. Bot. 56: 1761-1778. 2005.
Perotto S, Bonfante P. "Bacterial associations with mycorrhizal fungi:
close and distant friends in the rhizosphere.". Trends Microbial. 5:
496-501. 1997
124 Plant Ecology

Deacon JW. "Fungal Biology" Malden, MA: Blackwell Publishers. 2005.


Deshpande MV. "Mycopesticide production by fermentation: potential
and challenges.". Crit Rev Microbial. 25: 229-243. 1999.
Perotto S, Bonfante P. "Bacterial associations with mycorrhizal fungi:
close and distant friends in the rhizosphere.". Trends Microbial. 5:
496-501. 1997
Thomas MB, Read AF. "Can fungal biopesticides control malaria?". Nat
Rev Microbial. 5: 377-383. 2007.
6
Ecology of Nonvascular Plants

Plants are divide into two groups: plants lacking lignin-


impregnated conducting cells (the nonvascular plants) and
those containing lignin-impregnated conducting cells (the
vascular plants). Living groups of nonvascular plants
include the bryophytes: liverworts, hornworts, and mosses.
Vascular plants are the more common plants like pines,
ferns, com, and oaks.
Fossil and biochemical evidence indicates plants are
descended from multicellular green algae. Various green
algal groups have been proposed for this ancestral type,
with the Charophytes often being prominently mentioned.
Cladistic studies support the inclusion of the Charophytes
as sister taxa to the land plants. Algae dominated the oceans
of the precambrian time over 700 million years ago.
Between 500 and 400 million years ago, some algae made
the transition to land, becoming plants by developing a
series of adaptations to help them survive out of the water.
Vascular plants appeared by 350 million years ago, with
forests soon following by 300 million years ago. Seed plants
next evolved, with flowering plants appearing around 140
million years ago.
LIFE CYCLE OF PLANT

Plants have an alternation of generations: the diploid spore-


producing plant (sporophyte) alternates with the haploid
126 Plant Ecology

gamete-producing plant (gametophyte). Animal life cycles


have meiosis followed immediately by gametogenesis.
Gametes are produced directly by meiosis. Male gametes
are sperm. Female gametes are eggs or ova.
The plant life cycle has mitosis occurring in spores,
produced by meiosis, that germinate into the gametophyte
phase. Gametophyte size ranges from three cells (in pollen)
to several million. Alternation of generations occurs in
plants, where the sporophyte phase is succeeded by the
gametophyte phase. The sporophyte phase produces spores
by meiosis within a sporangium. The gametophyte phase
produces gametes by mitosis within an antheridium
(producing sperm) and/or archegonium (producing eggs).
These different stages of the flowering plant life cycle are
shown in Figure 1.

Figure 1. Plant life cycle

Within the plant kingdom the dominance of phases varies.


Nonvascular plants, the mosses and liverworts, have the
gametophyte phase dominant. Vascular plants show a
progression of increasing sporophyte dominance from the
ferns and "fern allies" to angiosperms.
HOMOSPORY AND HETEROSPORY

Plants have two further variations on their life cycles. Plants


that produce bisexual gametophytes have those
gametophyte&..germinate from isospores that are about all
Ecology of Nonvascular Plants 127

the same size. This state is reterred to as homospory. A


generalised homosporous plant life cycle is shown in Figure
2. Homosporous plants produce bisexual gametophytes.
Ferns are a classic example of a h01;nosporous plant.

Figure 2. A typical homosporous life cycle

. Plants that produce separate male and female


gametophytes have those gametophytes germinate from (or
within in the case of the more advanced plants) spores of
.different sizes. The male gametophyte produces sperm, and
is associated with smaller or microspores. The female
gametophyte is associated with the larger or megaspores.
Heterospory is considered by botanists as a significan~ step
toward the development of the seed. A generaHsed
heterosporous life cycle is shown in Figure 3.

Figure 3. Typical heterosporous life cycle


128 Plant Ecology

ADAYfATIONS TO LIFE ON LAND

Organisms in water do not face many of the challenges that


terrestrial creatures do. Water supports the organisr::., the
moist surface of the creature is a superb surface for gas
exchange, etc. For organisms to exist on land, a variety of
challenges must be met.
Drying out. Once removed from water and exposed to
air, organisms must deal with the need to conserve
water. A number of approaches have developed, such
as the development of waterproof skin (in animals),
living in very moist environments (amphibians,
bryophytes), and production of a waterproof surface
(the cuticle in plants, cork layers and bark in woody
trees)."
Gas exchange. Organisms that live in water are often
able to exchange carbon dioxide and oxygen gases
through their surfaces. These exchange surfaces are
moist, thin layers across which diffusion can occur.
Organismal response to the challenge of drying out
tends to make these surfaces thicker, waterproof, and
to retard gas exchange. Consequently, another method
of gas exchange must be modified or developed. Many
fish already had gills and swim bladders, so when
some of them began moving between ponds, the swim
bladder began to act as a gas exchange surface,
ultimately evolving into the terrestrial lung. Many
arthropods had gills or other internal respiratory
surfaces that were modified to facilitate gas exchange
on land. Plants are thought to share common ancestry
with algae. The plant solution to gas exchange is a new
structure, the guard cells that flank openings (stomata)
in the above ground parts of the plant. By opening
these guard cells the plant is abl~ to allow gas exchange
by diffusion through the open stomata.
Support. Organisms living in water are supported by
the dense liquid they live in. Once on land, the
Ecology of Nonvascular Plants 129

organisms had to deal with the less dense air, which


could not support their weight. Adaptations to this
include animal skeletons and specialised plant cells/
tissues that support the plant.
Conduction. Single celled organisms only have tyo
move materials in, out, and within their cells. A
multicellular creature must do this at each cell in the
body, plus move material in, out, and within the
organism. Adaptations to this include the circulatory
systems of animals, and the specialised conducting
tissues xylem and phloem in plants. Some multicellular
algae and bryophytes also have specialised conducting
cells.
Reproduction. Organisms in water can release their
gametes into the water, where the gametes will swim
by flagella until they ecounter each other and
fertilisation happens. On land, such a scenario is not
possible. Land animals have had to develop specialised
reproductive systems involving fertilisation when they
return to water (amphibians), or internal fertilisation
and an amniotic egg (reptiles, birds, and mammals).
Insects developed similar mechanisms. Plants have
also had to deal with this, either by living in moist
environments like the ferns and bryophytes do, or by
developing specialised delivery systems like pollen
tubes to get the sperm cells to the egg.

Bryophytes
Plant scientists recognize two kinds of land plants, namely,
bryophytes, or nonvascular land plants and
tracheophytes,or vascular land plants. Bryophytes are
small, herbaceous plants that grow closely packed together
in mats or cushions on rocks, soil, or as epiphytes on the
trunks and leaves of forest trees. Bryophytes are
distinguished from tracheophytes by two important
characters.
130 Plant Ecology

First, in all bryophytes the ecologically persistent,


photosynthetic phase of the life cycle is the haploid,
gametophyte generation rather than the diploid
sporophyte; bryophyte sporophytes are very short-lived,
are attached to and nutritionally dependent on their
gametophytes and consist of only an unbranched stalk, or
seta, and a single, terminal sporangium.
Second, bryophytes never form xylem tissue, the special
lignin- containing, water-conducting tissue that is found in
the sporophytes of all vascular plants. At one time,
bryophytes were placed in a single phylum, intermediate
in position between algae and vascular plants.
Modern studies of ~ell ultrastructure and molecular
biology, however,confirm that bryophytes comprise three
separate evolutionary lineages, which are today recognized
as mosses (phylum Bryophyta), liverworts (phylum
Marchantiophyta) and hornworts (phylum Antho-
cerotophyta). Following a detailed analysis of land plant
relationships, Kenrick and Crane proposed that the three
groups of bryophytes represent a grade or structural level
in plant evolution, identified by their "monosporangiate,r
life cycle. Within this the geologically oldest group, sharing
a fossil record with the oldest vascular plants in the
Devonian era.
Of the three phyla of bryophytes, greatest species
diversity is found in the mosses, with up to 15,000 species
recognized. A moss begins its life cycle when haploid
spores, which are produced in the sporophyte capsule,land
on a moist substrate and begin to germinate. From the one-
celled spore, a highly branched system of filaments, called
the protonema, develops. .
Cell specialization occurs within the protonema to form
a horizontal system of reddish-brown, anchoring filaments,
called caulonemal filaments and upright, green filaments,
called chloronemal filaments. Each protonema, which
superficially resembles a filamentous alga, can spread over
Ecology of Nonvascular Plants 131

several centimeters to form a fuzzy green film over its


substrate. As the protonema grows, some cells of the
caulonemal filaments specialize to form leafy buds that will
ultimately form the adult gametophyte shoots. Numerous
shoots typically develop from each protonema so that, in
fact, a single spore can give rise to a whole clump of moss
plants. Each leafy shoot continues to grow apically,
producing leaves in spiral arrangement on an elongating
stem.
In many mosses the stem is differentiated into a central
strand of thin-walled water-conducting cells, called
hydroids, surrounded by a parenchymatous cortex and a
thick-walled epidermis. The leaves taper from a broad base
to a pointed apex and have lamina that are only one-cell
layer thick. A hydroid-containing midvein often extends
from the stem into the leaf. Near the base of the shoot,
reddish-brown, multicellular rhizoids emerge from the
stem to anchor the moss to its substrate. Water and mineral
nutrients required for the moss to grow are absorbed, not
by the rhizoids,but rather by the thin leaves of the plant as
rain water washes through the moss cushion.
As is typical of bryophytes, mosses produce large,
multicellular sex organs for reproduction. Many
bryophytes are unisexual, or sexually dioicous. In mosses
male sex organs, called antheridia, are produced in dusters
at the tips of shoots or branches on the male plants and
female sex organs, the archegonia, are produced in similar
fashion on female plants. Numerous motile sperm are
produced by mitosis inside the brightly colored, c1ub-
shaped antheridia while a single egg develops in the base
of each vase-shaped archegonium. As the sperm mature,
the antheridium swells and bursts open.
Drops of rain water falling into the cluster of open
antheridia splash the sperm to near-by females. Beating
their two whiplash £lagellae, the sperm are able to move
short distances in the water film that covers the plants to
132 Plant Ecology

Sporangium -

Sp""'phylc -
(2,,) ;

Figure 4. The moss life cycle. The haploid gametophyte phase is free-living
and photosynthetic. The diploid sporophyte grows from and is nourished by
the gametophyte .

the open necks of the archegonia. Slimey mucilage


secretions in the archegonial neck help pull the sp~
downward to the egg. The closely packed arrangement of
the individual moss plants greatly facilitates fertilization.
Rain forest bryophytes that hang in long festoons from the
trees rely on torrential winds with the rain to transport their
sperm from tree to tree, while the small pygmy mosses of
exposed, ephemeral habitats depend on the drops of
morning dew to move their sperm. Regardless of where
Ecology of Nonvascular Plants 133

they grow, all bryophytes require water for sperm dispersal


and subsequent fertilization.
Embryonic growth of the sporophyte begins within the
archegonium soon after fertilization. At its base, or foot, the
growing embryo forms a nutrient transfer zone, or placenta,
with the gametophyte. Both organic nutrients and water
move from the gametophyte into the sporophyte as it
continues to grow. In mosses the sporophyte stalk, or seta,
tears the archegonial enclosure early in development,
leaving only the foot and the very base of the seta
embedded in the gametophyte. The upper part of the
archegonium remains over the tip of the sporophyte as a
cap-like calyptra.
Sporophyte growth ends with the formation of a
sporangium or capsule at the tip of the seta. Within the
capsule, water-resistant spores are formed by meiosis. As _
the mature capsule swells, the calyptra falls away. This
allows the capsule to dry and break open at its tip. Special
membranous structures, called peristome teeth, that are
folded down into the spore mass,now bend outward,
flinging the spores into the drying winds. Moss spores can
travel great distances on the winds, even moving between
continents on tne-le! streams. Their walls are highly
protective, allowing some spores to remain viable for up to
40 years. Of course, if the spore lands in a suitable, moist
habitat, germination will begin the cycle all over, again.
Liverworts and hornworts are like mosses in the
fundamental features of their life cycle, but differ greatly in
organization of their mature gametophytes and
sporophytes. Liverwort gametophytes can be either leafy
shoots or flattened thalli. In the leafy forms, the leaves are
arranged on the stem in one ventral and two lateral rows
or ranks, rather than in spirals like the mosses. The leaves
are one cell layer thick throughout, never have a midvein
and are usually divided into two or more parts called lobes.
The ventral leaves, which actually lie against the substrate,
134 Plant Ecology

are usually much smaller than the lateral leaves and are
; hidden by the stem.
Anchoring rhizoids, which arise near the ventral leaves,
are colorless and unicellular. The flattened ribbon-like to
leaf-like thallus of the thallose liverworts can be either
simple or structurally differentiated into a system of dorsal
air chambers and ventral storage tissues. In the latter type,
the dorsal epidermis of the thallus is punctuated with
scattered pores that open into the air chambers. Liverworts
synthesize a vast array of volatile oils, which they store in
unique organelles called oil bodies. These compounds
impart an often spicy aroma to the plants and seem to
discourage animals from feeding on them. Many of these
compounds have potential as antimicrobial or anticancer
pharmecuhcals.
Liverwort sporophytes develop completely enclosed
within gametophyte tissues until their capsules are ready
to open. The seta, which is initially very short,consists of
small, thin-walled, hyaline cells. Just prior to capsule
opening, the seta cells lengthen, thereby increasing the
length of the seta upto 20 times its original dimensions. This
rapid elongation pushes the darkly pigmented capsule and
upper part of the whitish seta out of the gametophytic
tissues. With drying, the capsule opens by splitting into
four segments, or valves. The spores are dispersed into the
winds by the twisting motions of numerous intermixed
sterile cells, called elaters. In contrast to mosses, which
disperse their spores over several days, liverworts disperse
the entire spore mass of a single capsule in just a few
minutes.
Hornworts resemble some liverworts in having simple,
unspecialized thalloid gametophytes, but they differ in
many other characters. For example, colonies of the
symbiotic _yanobacterium Nostoc fill small cavities that are
scattered throughout the ventral part of the hornwort
thallus. When the thallus is viewed from above, these
Ecology of Nonvascular Plants 135

colonies appear as scattered blue-green dots. The


cyanobacterium converts nitrogen gas from the air into
ammonium, which the hornwort requires in its metabolism
and the hornwort secretes carbohydrate- containing
mucilage which supports the growth of the
cyanobacterium.
Hornworts also differ from all other land plants in
having only one large, algal-like chloroplast in each thallus
cell. Hornworts get their name from their long, horn-
shaped sporophytes. As in other bryophytes, the
sporophyte is anchored in the gametophyte by a foot
through which nutrient transfer from gametophyte to
sporophyte occurs. The rest of the sporophyte, however, is
actually an elongate sporangium in which meiosis and
spore development take place.
At the base of the sporangium, just above the foot, is a
mitotically active meristeII1,which adds new cells to the
spore-producing zone throughout the life span of the
sporophyte. In fact, the sporangium can be releasing spores
at its apex, at the same time that new spores are being
produced by meiosis at its base. Spore release in hornworts
takes place gradually over a long period of time, and the
spores are mostly dispersed by water movements rather
than by wind
Mosses, liverworts and hornworts are found
throughout the world in a variety of habitats. They flourish
particularly well in moist, humid forests like the fog forests
of the Pacific northwest or the montane rain forests of the
southern hemisphere. Their ecological roles are many.They
provide seed beds for the larger plants of the community,
they capture and recycle nutrients that are washed with
rainwater from the canopy and they bind the soil to keep
it from eroding.
In the northern hemisphere peatlands, wetlands often
dominated by the moss Sphagnum, are particularly
136 Plant Ecology

important bryophyte communities. This moss has


exceptional water-holding capacity, and when dried and
compressed, forms a coal-like fuel. Throughout northern
Europe, Asia and North America, peat has been harvested
for centuries for both fuel consumption and horticultural
uses and today peatlands are managed as a sustainable
resource.
VASCULAR PLANTS GROUPS

The vascular plants have specialised transporting cells


xylem and phloem. When we think of plants we invariably
picture vascular plants. Vascular plants tend to be larger
and more complex than bryophytes, and have a life cycle
where the sporophyte is more prominent than the
gametophyte. Vascular plants also demonstrate increased
levels of organisation by having organs and organ systems.
Vascular plants first developed during the Silurian
Period, about 400 million years ago. The earliest vascular
plants had no roots, leaves, fruits, or flowers, and
reproduced by producing spores. Cooksonia, shown in
Figure 5, is a typical early vascular ,plant.

Figure 5. Cooksonia fossil specimen (L) and reconstruction (R)


Ecology of Nonvascular Plants 137

It was less than 15 em tall, with stems that


dichotomously branched. Dichotomous branching appears
a primitive or ancestral trait in vascular plants. Some
branches terminated in sporangia that produced a single
size of spore.
Many scientists now consider "Cooksonia" an
evolutionary grade rather than a true monophyletic taxon.
Their main argument is that not all stems of Cooksonia-type
plants have vascular tissue. The evolutionary situation of
a grade would have some members of the group having the
trait, others not. The shapes of sporangia on various
specimens of Cooksonia also vary considerably.
Rhynia, shown in Figure 6, is another early vascular
plant. Like Cooksonia, it lacked leaves and roots. One of the
species formerly assigned to this genus, R. major, has since
been reclassified as Aglaophyton major.

Figure 6. Rhynia gwynne-vaughanii (L) stem cross section from the Rhynie
Chert in Scotland

Aglaophyton major (Figure 7) a bryophyte, however, it


does have a separate free-living sporophyte that is more
prominent than the sporophyte, but appears to lack
lignified conducting cells. The remaining species, R.
gwynne-vaughanii is an undoubted vascular plant.
138 Plant Ecology

Figure 7. Reconstruction of Aglaophyton major (A-C) and Lyonaphyton


rhyniensis

Devonian plant lines included the trimerophytes and


zosterophyllophytes, which have been interpreted as
related to fems and lycophytes.

Psilophytes
The Psilotales aretl:te least complex of all terrestrial vascular
plants, and were once believed to be remnants of an
otherwise extinct Devonian flora. This is primarily because
psilophytes are the only living vascular plants to lack both
roots and leaves. Though they have been considered
"primitive," recent developmental and molecular evidence
suggests that the group may actually be reduced from fern-
like ancestors. There is not universal agreement on this, but
we here treat them with the ferns for that reason. Despite
the uncertainty of their relationships, psilophytes do
structurally resemble certain early vascular plants, and are
used as a model for understanding the ecology of these
plants.
This is ·a small group with only two genera, Psilotum,
shown above left, and Tmesipteris,. above right, neither
with many species. Both genera grow in tropical or
Ecology of Nonvascular Plants 139

subtropical regions, where they occur on rich soil or as


epiphytes. Psilotum occurs in North America in the
Caribbean, and along the Gulf and Atlantic Coasts to as far
north as North Carolina, and has been reported from one
locality in Arizona. It may also be found in tropical Asia
and on Pacific islands. Tmesipteris grows in New
Caledonia and nearby areas of the South Pacific, including
Australia and New Zealand.

Figure 8. Psilotum nudum

In addition to its natural distribution, Psiloturn is also


found as a common weed in greenhouses, and sometimes
escapes cultivation in regions with mild climate. It
occasionally becomes a nuisance, but is still very popular
for its unusual growth form. In Japan, more than 100
unusual breeds have been produced, some of them highly
prized by cultivators.

Morphology
The psilophyte stern lacks roots; it is anchored instead by
a horizontally creeping stern called a rhizome. The erect
140 Plant EcOlogy

portion of the stem bears paired enations, outgrowths


which look like miniature leaves, but unlike true leaves, the
enations have no vascular tissue. These paired outgrowths
lie immediately below the spore-producing synangia,
which produce the spores. The synangia appear to be the
product of three sporangia which became fused ~ver the
course of evolution, and are borne on the tip of a short
lateral branch. This is another feature in which the
psilophytes differ from other living vascular plants; all
other such plants produce their sporangia on their leaves.
You can click on the picture of the synangia of Psilotum at
right, for a better look
, .
at these structures.
When the synangia mature, they open to release yellow
to whitish spores, from which the gametophyte plants will
later emerge, like the one shown at left. The gametophytes
are very small, usually less than two millimeters long. They
are subterranean and saprophytic, getting their nutrition by
absorbing substances dissolVed in the environment. This is
often aided by the presence of fungi which grow into the
tissues of the gametophyte and through the surrounding .
soil.
Eventually, the gametophyte reaches sexual maturity,
producing both egg and sperm cells. The multiflagellate
sperm swim to the egg cells, where they unite to begin the
sporophyte generation. Psilophyte gametophytes may even
self-fertilize to produce a sporophyte plant. The resulting
sporophyte begins its life as a dependent on its parent
gametophyte, as in other seedless plants. But unlike the
"bryophytes," the sporophyte eventually gains
independence from its parent, and establishes itself in the
environment.
The mature sporophyte of Psilotum will often grow to
30 cm tall, and may grow even taller. It has no true leaves,
and instead the stem is green and photosynthetic, being
covered with stomates to,allow g"s exchange. As the cross-
section at right shows, the stem has a central core of
Ecology of Nonvascular Plants 141

vascular tissue (protostele) which is usually lobed. The


thick-walled cells in the center oflhis core are sometimes
considered to be pith, in which case the v~scular
arrangement would actually be a siphonostele.
Surrounding the vascular tissue is a layer called the
endodermis, which has specially packed cells to regulate
flow of water and nutrients.
Tmesipteris has similar reproductive structures and life
history to that of Psilotum, but by contrast it has broad leaf-
like extensions of its stem, each with a single vascular
bundle. These extensions may lie to either side of the stem,
forming a flat growth, or they may be radially arranged. In
any case, they are not considered leaves by most botanists,
though this interpretation has been challenged by some
workers.

Lycophytes
The next group, the Division Lycophyta, have their
sporangia organised into strobili (singular: strobilus). A
strobilus is a series of sporangia and modified leaves closely
grouped on a stem tip. The leaves in strobili are soft and
fleshy as opposed to the hard, modified leaves in cones.

Figure 9. Steps in the evolution of the microphyll leaf


Leaves that contained vascular tissue are another major
advance for this group. The leaves in lycophytes, both
U2 Plant Ecology

living and fossil forms, are known ~ microphylls. This term


does not imply any size constrain't, but rather refers to the
absence of a leaf gap in the vascular supply of the stem at
the point where the leaf vascular trace departs. Ferns and
other plants have megaphylls, leaves that produce this leaf
gap.
Today there are fewer genera of lycophytes than during
the group's heyday, the Paleozoic Era. Major living
lycophytes include Lycopodium, Isoetes, and Selaginella
(the so-called resurrection plant). LycopodiUm produces
isospores that germinate in the soil and produce a bisexual
gametophyte. These spores are all approximately the same
size.
Selaginella and Isoetes are heterosporous, and thus
produce two sizes of spores: small spores (termed
microspores) that germinate to produce the male
gametophyte; and larger spores (megaspores) that
germinate to produce the female gametophyte. The
production of two sizes of spores, and also making separate
unisexual gametophytes, is thought an important step
toward the seed. Modern lycophytes are small, herbaceous
plants. Many of the prominent fossil members of this group
produced large amounts of wood and were significant trees
in the Carboniferous-aged coal swamps.
Selaginella is a heterosporous member of the
lycophytes. Some species' of this genus are able to withstand
drying out by going dormant until they are rehydrated. For
this reason these forms of the genus are commonly called
resurrection plants.
Fossil Lycophytes: Baragwanathia and Drepanophycus
Baragwanathia is an undoubted lycophyte from the middle
Silurian deposits of Australia. It has microphyllous leaves
spirally attached to the stem, and sporangia clustered in
some areas of the plant, although not in terminal strobili as
in modern lycophytes.
Ecology of Nonvascular Plants 143

Figure 10. (a) Baragwanathia; (b) Drepanophycus

Drepanophycus ·is a middle Devonian lycophyte from the


Northern Hemisphere. Its features are very similar to
modern lycophytes.

Lepidodendron and Sigillaria


The Lycophytes became significant elements of the world's
flora during the Carboniferous time (the Mississippian and
Pennsylvanian are terms used for this time span in the
United States). These non-seed plants evolved into trees
placed in the fossil genera Lepidodendron and Sigillaria,
with heights reaching up to 40 meters and 20-30 meters
respectively. Lepidodendron stems are composed of less
wood (secondary xylem) that usually is found in
gymnosperm and angiosperm trees.
We know much about the anatomy of these coal-age
lycopods because of an odd type of preservation known as
a coal ball. Coal balls can be peeled and the plants that are
anatomically preserved within theIr laboriously studied to
learn the details of cell structure of these coal age plants.
Additionally, we have some exceptional petrifactions and
144 Plant Ecology

compressions that reveal different layers of the plants'


structure. Estimates place the bulk, up tc 70%, of coal
material as being derived from lycophytes.
Lepidodendron, was a heterosporous lycophyte tree
common in coal swamps of the Carboniferous time. As with
many large plant fossils, one rarely if ever finds the entire
tree preserved intact. Consequently there are a number of
fossil plant genera that are "organ taxa" and represent only
the leaves (such as Lepidophylloides), reproductive
structures (Lepidostrobus), stem (Lepidodendron), spores
(Lycospora), and roots (Stigmaria). Lepidodendron had
leaves borne spirally on branches that dichotomously
forked, with roots also arising spirally from the stigmarian
axes, and both small (microspores) and large (megaspores)
formed in strobili (a loose type of soft cone).

Lep1dophy1101des
(leeves) .

~ Lep1dostrobus
(cones)

Ulodendron
(brench seers)

1) Aculeatum
2) Obavatum
(outer berl<- mId trunk)

Figure 11. Lepidodendron


Ecology of Nonvascular Plants 145

Lepidodendron may have attained heigths of nearly 40


meters, with trunks nearly 2 meters in diameter. The trees
branched extensively and produced a large number of
leaves. When these leaves fell from the branches, they left
behind them the leaf scars characteristic of the genus.
Sigillaria was another arborescent lycopod, and is also
common in coal-age deposits. In contrast to the spirally
borne leaves of Lepidodendron, Sigillaria had leaved
arranged in vertical rows along the stem.

Sphenophyta
The division Sphenophyta contains once dominant plants
(both arborescent as well as herbaceous) in Paleozoic
forests, equisetophytes are today relegated to minor roles
as herbaceous plants. Today only a single genus,
Equisetum, survives. The group is defined by their jointed
stems, with many leaves being produced at a node,
production of isospores in cones borne at the tips of stems,
and spores bearing elaters (devices to aid in spore
dispersal). The gametophyte is small, bisexual,
photosynthetic, and free-living. Silica concentrated in the
stems give this group one of their common names: scouring
rushes. These plants were reportedly used by American
pioneers to scour the pots and pans. The fossil members of
this group are often encountered in coal deposits of
Carboniferous age in North America and Europe.

The Ferns
A fern is anyone of a group of about 20,000 species of plants
classified in the phylum or division Pteridophyta, also
known as Filicophyta. The group is also referred to as
Polypodiophyta, or Polypodiopsida when treated as a
subdivision of tracheophyta (vascular plants). The study of
ferns and other pteridophytes is called pteridology, and one
who studies ferns and other pteridophytes is called a
pteridologist. The term "pteridophyte" has traditionally
146 Plant Ecology

been used to describe all seedless vascular plants, making


it synonymous with "ferns and fern allies". This can be
confusing since members of the fern phylum Pteridophyta
are also sometimes referred to as pteridophytes.

Life cycle

Ferns are vascular plants differing from the more primitive


lycophytes by having true leaves (megaphylls), and they
differ from seed plants in their mode of reproduction -
lacking flowers and seeds. Like all other vascular plants,
they have a life cycle referred to as alternation of
generations, characterized by a diploid sporophytic and a
haploid gametophytic phase. Unlike the gymnosperms and
angiosperms, the ferns' gametophyte is a free-living
organism.

gametophyte
(prothaJlt.m)

Figure 12. Fern life cycle


Ecology of Nonvascular Plants 147

The life cycle of a typical fern is as follows:


A sporophyte (diploid) phase produces haploid spores
by meiosis;
A spore grows by cell division into a gametophyte,
which typically consists of a photosynthetic prothallus
The gametophyte produces gametes (often both sperm
and eggs on the same prothallus) by mitosis
A mobile, flagellate sperm fertilizes an egg that
remains attached to the prothallus
The fertilized egg is now a diploid zygote and grows
by mitosis into a sporophyte (the typical "fern" plant).
Ecology
The stereotypic image of ferns growing in moist shady
woodland nooks is far from being a complete picture of the
habitats where ferns can be found growing. Fern species
live in a wide variety of habitats, from remote mountain
elevations, to dry desert rock faces, to bodies of water or
in open fields.
Ferns in general may. be thought of as largely being
specialists in marginal habitats, often succeeding in places
where various environmental factors limit the success of
flowering plants. Some ferns are among the world's most
serious weed species, including the bracken fern growing
in the British highlands, or the mosquito fern (Azolla)
growing in tropical lakes, both species form large
aggressively spreading colonies.
There are four particular types of habitats that ferns are
found in: moist, shady forests; crevices in rock faces,
especially when sheltered from the full sun; acid wetlands
including bogs and swamps; and tropical trees, where
many species are epiphytes. Many ferns depend on
associations with mycorrhizal fungi. Many ferns only grow
within specific pH ranges; for instance, the climbing fern
(Lygodium) of eastern North America will only grow in
148 Plant Ecology

moist, intensely acid soils, while the bulblet bladder fern


with an overlapping range, is only ever found on limestone.
Structure
Like the sporophytes of seed plants, those of ferns consist
of:
Stems: Most often an underground creeping rhizome,
but sometimes an above-ground creeping stolon, or an
above-ground erect semi-woody trunk reaching up to
20 m in a few species.
Leaf. The green, photosynthetic part of the plant. In
ferns, it is often referred to as a frond, but this is
because of the historical division between people who
study ferns and people who study seed plants, rather
than because of differences in structure. New leaves
typically expand by the unrolling of a tight spiral called
a crozier or fiddlehead. This uncurling of the leaf is
termed circinate vemation. Leaves are divided into
three types:
Trophophyll: A leaf that does not produce spores,
instead only producing sugars by photosynthesis.
Analogous to the typical green leaves of seed
plants.
Sporophyll: A leaf that produces spores. These
leaves are analogous to the scales of pine cones or
to stamens and pistil in gymnosperms and
angiosperms, respectively. Unlike the seed plants,
however, the sporophylls of ferns are typically not
very specialized, looking similar to trophophylls
and producing sugars by photosynthesis as the
trophophylls do.
Brophophyll: A leaf that produces abnormally large
amounts of spores. There leaves are also larger than:
. the other leaves but bare a resemblance to
trophopylls.
"',"V&"I§Y 01 Nonvascular Plants 149

Roots:The underground non-photosynthetic structures


that take up water and nutrients from soil. They are
always fibrous and are structurally very similar to the
roots of seed plants.
The gametophytes of ferns, however, are very different
from those of seed plants. They typically consist of:
Prothallus: A green, photosynthetic structure that is one
cell thick, usually heart or kidney shaped, 3-10 mm
long and 2-8 mm broad. The prothallus produces
gametes by means of:
Antheridia: Small spherical structures that produce
flagellate sperm. -
Archegonia: A flask-shaped structure that
produces a single egg at the bottom, reached by the
sperm by swimming down the neck.
- Rhizoids: Root-like structures that consist of single
greatly-elongated cells, water and mineral salts are
absorbed over the whole structure. Rhizoids anchor the
prothallus to the soil.
One interesting difference between sporophytes and
gametophytes might be summed up by the saying that
"Nothing eats ferns, but everything eats gametophytes."
This is an over-simplification, but it is true that
gametophytes are often difficult to find in the field because
they are far more likely to be food than are the sporophytes.
Evolution and Oassification
Ferns first appear in the fossil record in the early-
Carboniferous period. By the Triassic, the first evidence of
ferns related to several modem families appeared. The
"great fern radiation" occurred in the late-Cretaceous,
when many modem families of ferns first appeared.
Ferns have traditionally been grouped in the Class
Filices, but modem classifications ~ssign them their own
division in the plant kingdom, called Pteridophyta.
150 Plant Ecology

Traditionally, three discrete groups of plants have been


considered ferns: two groups of eusporangiate fems-
families Ophioglossaceae and Marattiaceae-and the
leptosporangiate ferns. The Marattiaceae are a primitive
group of tropical ferns with a large, fleshy rhizome, and are
now thought to be a sibling taxon to the main group of
ferns, the leptosporangiate ferns. Several other groups of
plants were considered "fern allies": the clubmosses,
spikemosses, and quillworts in the Lycopodiophyta, whisk
ferns in Psilotaceae, and horsetails in the Equisetaceae.
More recent genetic studies have shown that the
Lycopodiophyta are only distantly related to any other
vascular plants, having radiated evolutionarily at the base
of the vascular plant clade, while both the whisk ferns and
horsetails are as much "true" ferns as are the
Ophioglossoids and Marattiaceae. In fact, the whisk ferns
and Ophioglossoids are demonstrably a clade, and the
horsetails and Marattiaceae are arguably another clade.
Molecular data - which remain poorly constrained for many
parts of the plants' phylogeny - have been supplemented
by recent morphological observations supporting the
inclusion of Equisetaceae within the ferns, notably relating
to the construction of their sperm, and peculiarities of their
roots. One possible means of treating this situation is to
consider only the leptosporangiate ferns as "true" ferns,
while considering the other three groups as "fern allies". In
practice, numerous classification schemes have been
proposed for ferns and fern allies, and there has been little
consensus among them. A new classification by Smith et a1.
is based on recent molecular systematic studies, in addition
to morphological data. This classification divides ferns into
four classes:
Psilotopsida,
Equisetopsida,
Marattiopsida,
Polypodiopsida.
Ecology of Nonvascular Plants 151

The last group includes most plants familiarly known


as ferns. Modern research supports older ideas based on
morphology that the Osmundaceae diverged early in the
evolutionary history of the leptosporangiate ferns; in
certain ways this family is intermediate between the
eusporangiate ferns and the leptosporangiate ferns.
Ferns are not as important economically as seed plants
but have considerable importance. Some ferns are used for
food, including the fiddleheads of bracken, Pteridium
aquilinum, ostrich fern, Matteuccia struthiopteris, and
cinnamon fern, Osmunda cinnamomea. Diplazium
esculentum is also used by some tropical peoples as food.
Ferns of the genus Azolla are very small, floating plants
that do not look like ferns. Called mosquito fern, they are
used as a biological fertilizer in the rice paddies of southeast
Asia, taking advantage of their ability to fix nitrogen from
the air into compounds that can then be used by other
plants.
A great many ferns are grown in horticulture as
landscape plants, for cut foliage and as houseplants,
especially the Boston fern. The Bird's Nest Fern, Asplenium
nidus, is also popular, and the staghorn ferns, genus
Platycerium, have a considerable following. Several ferns
are noxious weeds or invasive species, including Japanese
climbing fern (Lygodium japonicum), mosquito fern and
sensitive fern (Onoclea sensibilis). Giant water fern
(Salvinia molesta) is one of the world's worst aquatic
weeds. The important fossil fuel coal consists of the remains
of primitive plants, including ferns.
REFERENCES

Glime, Janice M., Bryophyte Ecology, Volume 1. PhysiolOgical Ecology.


Ebook sponsored by Michigan Technological University and the
International Association of Bryologists. 2007.
Moran, Robbin C. A Natural History of Ferns. Portland, OR: Timber Press.
2004.
152 Plant Ecology

Lord, Thomas R. Ferns and Fern Allies of Pennsylvania. Indiana, P A;


Pinelands Press. 2006.
Raven, Peter H., Evert, Ray F., & Eichhorn, Susan E. Biology of Plants. New
York: W. H. Freeman and Company. 2005.
Schofield, W. B. Introduction to Bryology New York: Macmillan. 1985.
Watson, E.V. The Structure and Life of Bryophytes. London: Hutchinson
University library. 1971.
7
Ecology of Seed Plants

The spermatophyte!>, which means "seed plants", are some


of the most important organisms on Earth. Life on land as
we know it is shaped largely by the activities of seed plants.
Soils, forests, and food are three of the most apparent
products of this group.
Seed-producing plants are probably the most familiar
plants to most people, unlike mosses, liverworts, horsetails,
and most other seedless plants which are overlooked
because of their size or inconspicuous appearance. Many
seedplants are large or showy. Conifers are seed plants;
they include pines, firs, yew, redwood, and many other
large trees. The other major group of seed-plants are the
flowering plants, including plants whose flowers are
showy, but also many plants with reduced flowers, such as
the oaks, grasses, and palms.
Today, the seed plants are some of the most important
organisms on earth. Life on land as we know it is shaped
largely by the activities of seed plants. This large and
important group appeared early in the evolution of
vascular plants, and throughout the Late Paleozoic shared
dominance of the land flora with ferns, lycophytes, and
sphenopsids. Since the beginning of the Mesozoic,
however, most trees and forests have consisted of seed
plants.
154 Plant Ecology

HISTORY AND EVOLUTION

The seed plants are often divided arbitrarily into two


groups: the gymnosperms and the angiosperms. The basis
for this distinction is that angiosperms produce flowers,
while the gymnosperms do not. This is poor form, since it
defines the gymnosperms by the absence of a character, and
not by any features that the organisms actually share. The
gymnosperms do share a number of features, but, as should
be obvious from the above cladogram, they are not more
closely related to each other than to the angiosperms. The
features shared by gymnosperms were likely present in the
early ancestors of the flowering plants as well. It should
also be noted that the "progymnosperms" are represented
by a box of a different color, in order to make it clear that
they are not actually seed plants, but rather are included
here because they are believed to be the closest relatives of
the seed plants.
The earliest seeds appear in the Late Devonian. The
oldest known seed plant is Elkinsia polymorpha, a "seed
fern" from Late Devonian (Famennian) of West Virginia.
Though the fossils consist only of small seed-bearing
shoots, these fragments are quite well-preserved. This has
allowed us to learn details about the evolutionary
development of the seed. Another such fossil from about
this time is Archaeosperma, also known only from
fragments.
The earliest seed plants produced their seeds along their
branches without specialized structures, such as cones or
flowers, unlike most living seed plants. The seeds were
produced singly or in pairs, and were surrounded by a
loose cupule. This small cup-like structure was lobed in the
earliest seeds, producing a somewhat sheltered chamber at
one end of the seed. Within this cupule, the seed was
enclosed by a more tighly appressed tissue called the
integument. The integument is a layer of tissue found in all
seeds; it is produced by the parent plant, and develops into
Ecology of Seed Plants lSS

the seed coat. As the integument evolved to enclose the


seed more tightly, an opening was left at one end, called
the micropyle, which permitted pollen to enter and provide
sperm to fertilize the egg cell. Both the integuments and
cupule are believed to be the result of reduced and fused
branches or leaves.
In later seed plants, a small pollen chamber appears just
inside the micropyle. In modem cycads and conifers, this
chamber exudes sticky fluids to aid in pollen capture, and
as the fluid dries, it pulls the pollen inside the micropyle.
This structure is preserved in detail in a number of recently
discovered permineralized Devonian seeds. Besides
preserving the pollen drop, minerals replaced the original
tissues gradually, such that fine detail of the cell walls can
be studied - a few Permian seeds even have preserved
embryos.
Seed plants diversified and spread in the Late Paleozoic.
By the end of the Devonian, a variety of early seed plants
collectively known as "lyginopterids" appeared. These
include Sphenopteris, a plant with fern-like leaves, but
which bore seeds and cupules.
The Carboniferous saw an increase in the number and
kinds of seed plants. In the coal swamps of North America
grew pteridosperms like Medullosa, a seed plant that
resembles modern tree-ferns, but which bore seeds.
Cordaites also grew in these swamps, and in a number of
other habitats including ocean-edge environments similar
to that of the modem mangrove. However, the cordaites are
believed to be closer relatives of modem conifers. Both the
medullosans and cordaites were small trees when
compared to the great scale-trees which dominated these
Late Paleozoic coal swamps. Seed plants were thus
overshadowed in their early evolution by plants which did
not produce seeds. By the Westphalian, the Voltziales first
show up. These are believed to be the closest relatives of
modem conifers, and in fact some paleobotanists classify
156 Plant Ecology

them as conife~s. By the Permian, the seed plants were


beginning to produce large trees, and by the Triassic, all
major groups of seedplants had appeared, except for the
flowering plants.
LIFE CYCLE OF SEED PLANTS

Seed plants are heterosporous- they have 2 different spore


sizes: megaspores and microspores. The generalized life
cycle of plants has been modified to illustrate plants which
have separate male and female gametophytes
(megagametophyte and microgametophyte) produced by
different sized spores (megaspores and microspores).

/ sporophyte ~

seed ~ ~ 2N
- fertiliz8\" -- -- - -- - - - - -- -\ riOSr t---~ .
sperm eggs megaspores mlcrospores

\
"---= \... megagametoPhyt!)
microgametophyte

Figure 1. Life cycle of seed plants

The evolutionary trend from nonvascular plants to seedless


vascular plants to seed plants has been a reduction in the
size of the gametophyte. In seed plants, the gametophyte
is usually microscopic and is retained within the tissues of
the sporophyte.
The megasporangium is surrounded by layers of
sporophyte tissue called the integument. The integument
and structures within (megasporangium, megaspore) are
the ovule.
Microspores germinate within the sporophyte tissue
and become pollen grains. The microgametophyte is
Ecology of Seed Plants 157

contained within the tough, protective coat of the pollen


grain. The entire microgametophyte (pollen grain) is
transferred to the vicinity of the megagametophyte by a
process of pollination. Wind or animals usually accomplish
this transfer.
When pollen reaches the female gametophyte, it
produces an elongate structure (pollen tube) that grows to
the egg cell. Sperm are transferred directly through this
tube to the egg. The advantage of this process is that sperm
do not have to swim long distances as they do in seedless
plants.
MORPHOLOGY OF SEED PLANTS

The seed includes three primary regions: the embryo,


nutritive tissue, and seed coat. The embryo is the young
sporophyte plant. This is what will grow into the new tree,
shrub, vine, etc. The embryo is usually surrounded by some
sort of nutritive tissue which will feed it during its early
growth, until it can establish its own root system and leaves
to support itself.
The origin of this nutritive tissue varies from group to
group of seed plants. Nutrients in the tissue are absorbed
into the developing embryo by specially modified leaves
called cotyledons. In some plants, the cotyledons may
absorb all the nutrients before the seed is even dispersed,
storing the food inside themselves.
Around the whole seed is a layer called the seed coat.
This layer may be thick or thin, depending on the species,
but it often contains light-sensitive chemicals. When
conditions are right - there is appropriate light and water
- the seed coat may trigger the germination of the seed.
Many plants use this to break a period of dormancy, when
the embryo remains inactive. This dormancy can be very
important for plants in seasonal habitats, or any
environment where the water or light vary greatly over
time.
158 . Plant Ecology

pUnu!e--"'1II! (seed coat)

eplcotyl

embryo
~---mlcropyle

radicle
dlcot seed (bean)

Figure 2. Structure of a typical seed

The seed does not develop from just any part of the plant,
but from special structures called ovules. The ovule is an
immature seed, which does not yet contain a viable embryo.
It is only when the egg cell inside the ovule is fertilized by
sperm that the ovule is called a seed. The ovule is
surrounded by integument tissues which produce the seed
coat, and in the earliest seed plants another layer called the
cupule enclosed the entire ovule/seed.
While all plants may grow larger by primary growth
from their branch tips, not all plants are capable of
secondary growth. Secondary growth is the increase in
diameter of existing tissues and organs, and this process
results in secondary tissues. In seed plants, two kinds of
secondary tissues are produced: wood and periderm.
Wood is produced by the vascular cambium, a layer of
cells whose job it is to divide off cells for new conducting
tissues. The vascular cambium is a cylindrical region
running through the entire stem of the plant, and branching
into every twig and limb.
When the cells of this cHmbium divide, they may
produce new cells toward the outside of the cylinder, or
Ecology of Seed Plants 159

toward the inside. Tho~ which split off to the outside of


the cylinder become new phloem tissue, which transports
the sap of the tree and thus moves food manufactured in
the leaves down to the roots. Those cells which split off
toward the inside of the cylinder become new xylem tissue,
which transports water and minerals up from the roots, and
also provides for movement of materials between the
exterior and interior of the plant. When a great deal of
xylem accumulates, it is called wood; plants with wood
may be trees, shrubs, or stout vines.
Periderm is the other product of secondary growth; it
is produced by the cork cambium, a cylindrical layer of cells
which develops not far under the outer skin (epidermis) of
the plant. Like the vascular cambium, the cork cambium
divides new cells toward the inside and toward the outside.
Those which are produced toward the outside become the
particularly important tissue called cork. Cork is important
because it replaces the original outer layers of tissue as the
plant grows.
A young plant begins with a smooth intact surface, but
the growth of two cambial layers producing new tissues
strains this outer layer, causing it to rupture. Cork is
produced to replace these lost tissues, and thus protect the
inner tissues. Usually, this new cork will have a different
texture, and may even later be replaced by additional new
cork. Together, the periderm and the phloem (which lies
just to the inside of the periderm) are called the bark.
Secondary growth occurs only in seed plants and a few
of their extinct close relatives, such as the progymno-
sperms. Even so, not all seed plants actually produce such
secondary growth; the monocots are one such group which
lacks this kind of growth. There are a number of other
groups which do develop secondary tissues, such as some
extinct lycophytes and sphenophytes, and their secondary
growth occurs by a similar process, though it differs in a
number of important details. The extinct scale trees, for
160 Plant Ecology

instance, produced woody growth enough to reach heights


of several dozen meters, but their primary support was
their periderm (bark, not wood), and both cambia were
unifacial, that is they divided new cells only toward one
side of the cambium, and not to both sides as in seed plants.
GYMNOSPERMS

Gymnosperms have seeds but not fruits or flowers.


Gymnos means naked, sperm means seed: thus the term
gymnosperm = naked seeds. Gymnosperms developed
during the Paleozoic Era and became the dominant seed
plant group during the early Mesozoic Era. The ancestors
of gymnosperms were some now-extinct type of
heterosporous fern or related group. There are 700 living
species of gymnosperms placed into four divisions:
conifers, cycads, ginkgos, and gnetophytes. Gymnosperms
are undoubtedly the group from which the angiosperms
developed, although, as Charles Darwin noted in Origin of
Species, which group "remains an abominable mystery".
Numerous gymnosperm groups have been proposed as
flowering plant ancestors over the past century.

Cycads
Cycads are placed in the Division Cycadophyta. They retain
several fern-like features, notably pinnate leaves and
circinate vernation. However, they usually produce cones
of nonphotosynthetic reproductive structures, a
distinctively unfemlike feature. Cycads, like all seed plants,
are also heterosporous, unlike the ferns which are all
homosporous. Cycad cones are unisexual, in fact the plants
producing them are dioecious, having separate male and
female plants. Cycads also proouce free-swimming sperm.
Cycads were much more prominent in the forests of the
Mesozoic than they are today. Presently, they are restricted
to the tropics. Zamia floridana is the only cycad occurring
natively in the continental United States. I
Ecology of Seed Plants 161

Figure 3. Cycas revoluta female corn

Figure 4. Cycas revoluta male corn

Several species of Cycas, notably C. revoluta, are commonly


encountered cultivated plants in warm, moist areas. Cycas
revoluta leaves are often used in Palm Sunday services in
some churches, both for .their feathery appearance and ease
of obtaining from local greenhouses.
162 Plant Ecology

Ginkgos
The ginkgoes also were a much more prominent group in
the past than they are today. The sole survivor of this once
robust and diverse group is Ginkgo biloba, the maidenhair
tree shown in Figure 5.

Fanshaped leaves

The Ginkgo tree

Figure 5. Ginkgo biloba

Extensively used as an ornamental plant, Ginkgo was


thought extinct in the wild until it was discovered growing
natively in a remote area of China. Ginkos are dioecious,
with separate male and female plants. The males are more
commonly planted since the females produce seeds that
have a nasty odor. Pollination is by wind. Recently, Ginkgo
has become the current herbal rave, although scientific
studies have debunked the claim that the herbal
supplement made from ginkgoes improves memory.
Precise systematic placement of the ginkgoes has yet to
bet determined. Ginkgoes have motile (swimming) sperm,
a rarity among living seed plants, although the vegetative
I
Ecology of Seed Plants 163

anatomy of ginkgoes is more conifer-like (long shoot and


short shoot morphology discussed below; structure of their
wood). Ginkgoes, like the cycads, are dioecious, and also
have similar seed features to cycads.
Plants possibly allied to the modern ginkgoes have been
found in Permian-aged and later rocks. These plants have
been classified in the leaf-genera Ginkgoites and Baiera,
although recent studies suggest these genera are really
morphological variants and that the modern genus Ginkgo
should be used to include these fossils . During the
Mesozoic ginkgoes were worldwide in their distribution
and important elements in the gymnosperm forests that
dominated the land.

Conifers
The conifers remain a major group of gymnosperms that
include the pines, spruce, fir, bald cypress and Norfolk
Island Pine (Araucaria). The division Pinophyta contains
approximately 550 species of conifers. The conifers are cone
producing trees and shrubs that usually have evergreen
needle-like leaves. Needles have a thick cuticle, sunken
stomates, and a reduced surface area. The conifers, as a
groupr are well adapted to withstand extremes in climate
and occur in nearly all habitats from the equator to the
subpolar regions. The taiga biome consists largely of
various conifer species.

Auracarias
The Araucariaceae are a very ancient family of conifers.
They achieved maximum diversity in the Jurassic and
Cretaceous periods, when they existed almost worldwide.
At the end of the Cretaceous, when dinosaurs became
extinct, so too did the Araucariaceae in the northern
hemisphere. There are three genera with 41 species alive
today, Agathis, Araucaria and Wollemia, all derived from
164 Plant Ecology

the Antarctic flora and distributed largely in the southern


hemisphere. By far the greatest diversity is in New
Caledonia (18 species), with others in southern South
America, New Zealand, Australia and Malesia, where
Agathis ext~nds a short distance into the northern i
hemisphere, reaching 18°N in the Philippines. All are
evergreen trees, typically with a single stout trunk and ve~
regular whorls of branches, giving them a formal
appearance. Several are very popular ornamental trees--in
gardens, in subtropical regions, and some are also very
important timber trees, producing wood of high quality.
Several have edible seeds similar to pine nuts, and others
produce valuable resin and amber. In the forests where they
occur, they are usually dominant trees, often the largest
species in the forest; the largest is Araucaria hunsteinii,
reported to 89 m tall in New Guinea, with several other
species reaching 50-65 m tall.
The petrified wood of the famous Petrified Forest east
of Holbrook, Arizona are fossil Araucariaceae. During the
Upper (Late) Triassic the region was moist and mild. The
trees washed from where they grew in seasonal flooding
and accumulated on sandy delta mudflats, where they were
buried by silt and periodically by layers of volcanic ash
which mineralized the wood. The fossil trees belong
generally to three species of Araucariaceae, the most
common of them being Araucarioxylon arizonicum. Some
of the segments of trunk represent giant trees that are
estimated to have been over 50 meters tall when they were
alive.
Members of this group of conifers have numerous
small, scale-like leaves spiraling around their stems.
Araucaria, a major genus that gives its name to the group,
is a common ornamental because of the symmetry and
beauty of its growth form. The monkey puzzle tree, shown
in Figure 6, is a species of Araucaria.
Ecology of Seed Plants 165

....,.

Figure 6. Araucaria imbricata

The fossil record of Auracarias and similar plants is quite


good. The fossil genus Auracarioxylon that grew in Arizona
during the Triassic Period comprises the largest group of
petrified wood in the Petrified Forest National Park of
Arizona.

Taxodiaceae

Taxodiaceae was at one time regarded as a distinct plant


family comprising the following ten genera of coniferous
trees:
Athrotaxis
Cryptomeria
Cunninghamia
166 Plant Ecology

Glyptostrobus
Metasequoia
Sciadopitys
Sequoia
Sequoiadendron
Taiwania
Taxodium
However, recent research has shown that the Taxodiaceae,
with the single exception of Sciadopitys, should be merged
into the family Cupressaceae. There are no consistent
characters by whi~h they can be separated, and genetic
evidence demonstrates close relationships; this merging is
now becoming widely accepted.
The one exception, the genus Sciadopitys, is genetically
very distinct from all other conifers, and now treated in a
family of its own, Sciadopityaceae.
Members of this group include some of the largest trees,
and have been significant members of the forests of the
world since the Mesozoic. Sequoia, shown in Figure 7, and
Seq~oiadendron are major genera in this group.
S~quoia is a genus in the cypress family Cupressaceae
(formerly treated in Taxodiaceae), containing the single
living ~pecies Sequoia sempervirens. Common names
include Coast Redwood and California Redwood (it is one
of three ~pecies of trees known as redwoods). It is an
evergreen, long-lived, monoecious tree living for up to
2,200 years, and is the tallest tree in the world, reaching up
to 115.5 m (379.1 ft) in height and 8 m (26 ft) diameter at
breast height. The crown is c€>nical, with horizontal to
slightly drooping branches. The bark is very thick, up t~ 30
cm (12 in), and quite soft, fibrous with a bright red-brown
when freshly exposed (hence the name 'redwood'),
weathering darker. The root system is composed of
shallow, wide-spreading lateral roots.
Ecology of Seed Plants 167

Figure 7. Sequoia plant

The leaves are variable, being 15-25 mm long and flat on


young trees and shaded shoots in the lower crown of old
trees, and scale-like, 5-10 mm long on shoots in full sun in
the upper crown of older trees; there is a full range of
transition between the two extremes. They are dark green
above, and with two blue-white stomatal bands below. Leaf
arrangement is spiral, but the larger shade leaves are
twisted at the base to lie in a flat plane for maximum light
capture. The seed cones are ovoid, 15-32 mm long, with 15-
25 spirally arranged scales; pollination is in late winter with
maturation about 8-9 months after. Each cone scale bears
3-7 seeds, each seed 3-4 mm long and 0.5 mm broad, with
two wings 1 mm wide. The seeds are released when the
168 Plant Ecology

cone scales dry out and open at maturity. The pollen cones
are oval, 4-6 mm long. The species is monoecious, with
pollen and seed cones on same plant. Its genetic makeup
is unusual among conifers, being a hexaploid and likely
autoallopolyploid. The mitochondrial genome is paternally
inherited.

Pine Life Cycle


Pines have an interesting life cycle, that takes two years to
complete. Not all seed plants have such a long time span
to complete their life history: some flowering plants
manage to do it in as little as a few weeks.

Figure 8. Pine Life Cycle


Ecology of Seed Plants 169

The sporophyte, as in all other vascular plant groups,


is the dominant, photosynthetic part of the life cycle: when
you are holding pine needles in your hand you are holding
sporophyte parts. Pines have specialised reproductive
structures in whicn~ meiosis occurs: pine cones. Pollen
grains are produced in the male cones, and contain the male
gametophyte (which consists of only a very few cells).
Pollen released from the male cones is carried by wind to
the female cones, where it lands.
The cones close and the next year the pollen grain
germinates to produce a pollen tube that grows into the
female gametophyte. The sperm cell (from the pollen grain)
and egg cell fuse, forming the next generation sporophyte.
The sporophyte develops into an embryo encased
within a seed. The seed is later released to be transported
by the wind to where it lands and germinates. If you have
seen a large pine tree you realise there are hundreds or
more female cones on such a tree. Pine pollen has been
noted to travel great distances from the plant that produced
it, if the wind is strong enough. To aid this transport pine
pollen has two air sacs, and thus is quite distinctive.

Gnetales
The Gnetales, shown in Figure 9, are an odd group: they
have some angiosperm-like features but are not themselves
angiosperms. Cladistic analyses support placement of the
gnetales (or some portion of them) as outgroups for the
flowering plants. Three distinctive genera comprise this
group: Welwitschia, Gnetum, and Ephedra.
Ephedra occurs in the western United States where it
has the common name "Mormon tea". It is a natural source
for the chemical ephedrine, although there is no evidence
the Mormons in Utah (where the plant is extremely
common) ever used it for tea. Welwitschia is limited to
coastal deserts in South Africa, although fossil leaf, cuticle
170 Plant Ecology

and pollen evidence indicates plants of this type were


widespread during the Mesozoic Era. Welwitschia is noted
for-its two long, prominent leaves. Gnetum: has leaves that
look remarkably like those in angiosperms, as well as
vessels in the xylem, generally considered an angiosperm-
characteristic.

Figure 9. Ephedra

Among the gnetalean plants, Ephedra is perhaps the best


known~ One folkloric name for the plant is "Mormon tea".
This is a misnomer as there appears little or no evidence
that members of a religion that bans stimulants such as
caffeine ever brewed a tea from the plant. However, the
plant does produce the drug ephedrine, a stimulant lately
linked to deaths of athletes. Welwistchia is a very bizarre
plant natively growing only in the coastal deserts of South
Africa. The plant produces two long leaves and a crown of
reproductive cones rimming a brown, central body. Pollen
resembling Welwitschia has been found in many parts of
the world, indicating a formerly more widespread
distribution of this enigmatic plant.
Ecology pf Seed Plants 171

ANGIOSPERMS

Flowering plants, the angiosperms, were the last of the seed


plant groups to evolve, appearing during the later part of
the of the Age of Dinosaurs. All flowering plants produce
flowers. Within the female parts of the flower angiosperms
produce a diploid zygote and triploid endosperm.
Fertilisation is accomplished by a variety of pollinators,
including wind, animals, and water. Two sperm are
released into the female gametophyte: one fuses with the
egg to produce the zygote, the other helps form the
nutritive tissue known as endosperm. The angiosperms
(angios = hidden) produce modified leaves grouped into
flowers that in tum develop fruits and seeds. There are
presently 235,000 known living species.

Flowers
Flowers are collections of reproductive and sterile tissue
arranged in a tight whorled array having very short
internodes. Sterile parts of flowers are the sepals and petals.
When these are similar in size and shape, they are termed
tepals. Reproductive parts of the flower are the stamen
(male, collectively termed the androecium) and carpel
(often the carpel is referred to as the pistil, the female parts
collectively termed the gynoecium). Lily flowers
demonstrates these concepts.
Flowers may be complete, where all parts of the flower
are present and functional, or incomplete, where one or
more parts of the flower are absent. Many angiosperms
produce a single flower on the tip of a shoot. Other plants
produce a stalk bearing numerous flowers, termed an
inflorescence, such as is seen in many orchids. Many
flowers show adaptations for insect pollination, bearing
numerous white or yellow petals. Others, like the grasses,
oaks, and elms, are wind pollinated and have their petals
reduced and often inconspicuous.
172 Plant Ecology

Angiosperm Life Cycle


Flowering plants also exhibit the typical plant alternation
of generations, shown in Figure 10.

Figure 10. Angiosperm life cycle

The dominant phase is the sporophyte, with the


gametophyte being much reduced in size and wholly
dependant on the sporophyte for nutrition. The is not a
unique angiosperm condition, but occurs in all seed plants
as well. What makes the angiosperms unique is their
flowers and the "double fertilisation" that occurs.
Technically this is not double fertilisation, but rather a
single egg-sperm fusion (fertilisation proper) plus a fusion
of the second of two sperm cells with two haploid cells in
the female gametophyte to produce triploid
(3n) endosperm, a nutritive tissue for the developing
embryo.
Ecology of Seed Plants 173

Angiosperm Systematics
The flowering plants, the division Magnoliophyta, contain
more than 235,000 species, six times the number of species
of all other plants combined. The flowering plants divide
into two large groups, informally named the monocots and
the dicots. The techjnical names for these groups are the
class Magnoliopsida for dieots and the class Liliopsida for
monocots.
The-Elirotyledons are in the class Magnoliopsida and
have these features: either woody or herbaceous, flower
parts usually in fours and fives, leaves usually net-veined,
vascular bundles arranged in a circle within the stem, and
produce two cotyledons (seed leaves) at germination.
Prominent dieot families include the mustards, maples,
cacti, peas and roses. Several dieot families are noteworthy
because of the illegal drugs derived from them: the
Cannabinaceae (marijuana) and Papaveraceae (poppies
from which opium and heroin are derived). Erythroxylum
coca (in the dieot family Erythroxylaceae) is the plant from
which the illegal drug cocaine is extracted.
Not all dieot plants are misused to produce illegal
drugs. Notable dieot families with legitimate uses include
the pea family, whieh includes the crop plants beans,
clover, and peas as well as many ornamental landscape
plants such as Acacias. Beans are an excellent source of
nonanimal protein as well as fiber. Chocolate and cola are
products of the plant family Sterculiaceae. CoHee is
produced from CoHea arabica, a plant in the family
Rubiaceae, while tea comes from Camelia sinensis
(Theaceae), a plant native to China.
The class Liliopsida has plants that are herbaceous (a
majority are, only palms and bamboo stand out as monocot
trees), flower parts are in threes, leaves are usually parallel-
veined, vascular bundles are scattered within the stem, and
produce one cotyledon (seed leaf) at germination. Monocot
families include lilies, palms, orchids, irises, and grasses.
174 Plant Ecology

The monocot family Poaceae (known previously as the'


Gramineae) includes the grasses such as com, oats, wheat,
rye, and rice that are staple food products as well as
ornamental plants such as crabgrass and tiff grass. The
importance of this plant family to modem civilisation
cannot be overstated, as the first six plants mentioned in the
previous sentence provide 75% of our food, either directly
as food we eat or indirectly as 'food for animals we eat.

Evolution of Flowering Plants


Several evolutionary trends within the/plant kingdom have
been noted. The mODophyletic nateure of this kingdom is
not in dispute, with the first major division being between
~scular and nonvascular plants. Wihin the vascular plants .
we see increasing changes in the relationship between
sporophyte and gametophyte, culminating in flowering
plants.
Developing from green algal ancestors, plants show a
trend for reduction of the complexity, size, and dominance
of the gametophyte generation. In nonvascular plants the
gametophyte is the conspicuous, photosynthetic, free-
living phase of the life cycle. Conversely, the angiosperm
gametophyte is reduced to between three and eight cells
and is dependent on the free-living, photosynthetic
sporophyte for its nutrition. Plants also developed and
refined the root-shoot-Ieafaxis with its specialized
conducting cells of the xylem and phloem. The earliest
vascular plants, such as Cooksonia and Rhynia, were little,
more than naked (unleafed) photosynthetic stems. Some
plants later developed secondary growth that produced
wood. Numerous leaf modifications are known, including
"carnivorous" plants such as the Venus flytrap, as well as
plants that have reduced or lost leaves, such as Psilotum
and the cacti.
A third trend is the development of the seed to promote
the dormancy of the embryo. The seed allows the plant to
Ecology of Seed Plants 175

wait out harsh environmental conditions. With the


development of the seed during the Paleozoic era plants
became less prone to mass extinctions. The fourth trend in
plant evolution is the encasing of a seed within a fruit. The
only plant group that prodl,lces true fruit is the flowering
plants, the angiosperms. Fruits serve to protect the seed, as
well as aid in seed dispersal.
Land plants have existed for about 425 million years.
Early land plants reproduced by spores like their aquatic
counterparts. Marine organisms can easily scatter copies of
themselves to float away and grow elsewhere. Land plants
soon found it advantageous to protect their copies from
drying out and other hazards by enclosing them in a case,
the seed. Early seed bearing plants, like the ginkgo, and
conifers (such as pines and firs), did not produce flowers.
The earliest fossil of an angiosperm, or flowering plant,
Archaefructus liaoningensis, is dated to about 125 million
years BP. Pollen, considered directly linked to flower
development, has been found in the fossil record perhaps
as long ago as 130 million years.
While there is only hard evidence of such flowers
existing about 130 million years ago, there is some
circumstantial evidence that they may have existed 250
million years ago. A chemical used by plants to defend their
flowers, oleanane, hasl5een detected in fossil plants that
old, including gigantopterids, which evolved at that time
and bear many of the traits of modem, flowering plants,
though they are not known to be flowering plants
themselves, because only their stems and prickles have
been found preserved in detail, one of the earliest examples
of petrification.
The apparently sudden appearance of relatively
modem flowers in the fossil record posed such a problem
for the theory of evolution that it was called an "abominable
mystery" by Charles Darwin. However the fossil record has
grown since the time of Darwin, and recently discovered
176 Plant Ecology

angiosperm fossils such as Archaefructus, along with further


discoveries of fossil gymnosperms, suggest how
angiosperm characteristics may have been acquired in a
series of steps.
Several groups of extinct gymnosperms, particularly
seed ferns, have been proposed as the ancestors of
flowering plants but there is no continuous fossil evidence
showing exactly how flowers evolved. Some older fossils,
such as the upper Triassic Sanmiguelia, have been
suggested. Based on current evidence, some propose that
the ancestors of the angiosperms diverged from an
unknown group of gymnosperms during the late Triassic
(245-202 million years ago). The relationship of the earlier
gigantopterids to flowering plants is still enigmatic.
A close relationship between Angiosperms and
Gnetophytes, suggested on the basis of morphological
evidence, has been disputed on the basis of molecular
evidence that suggest Gnetophytes are more closely related
to other gymnosperms. Recent DNA analysis (molecular
systematics) show that Amborella trichopoda, found on the
Pacific island of New Caledonia, belongs to a sister group
of the other flowering plants, and morphological studies
suggest that it has features which may have been
characteristic of the earliest flowering plants.
The great angiosperm radiation, when a great diversity
of angiosperms appear in the fossil record, occurred in the
mid-Cretaceous (approximately 100 million years ago).
However, a study in 2007 estimated that the division of the
five most recent (the genus Ceratophyllum, the family
Chloranthaceae, the eudicots, the magnoliids, and the
monocots) of the eight main groups occurred around 140
million years ago. By the late Cretaceous, angiosperms
appear to have become the predominant group of land
plants, and many fossil plants recognizable as belonging to
modern families (including beech, oak, maple, and
magnolia) appeared.
Ecology of Seed Plants 177

However, some authors have proposed an earlier origin


for angiosperms, sometime in the Paleozoic (251 million
years ago or more).
It is generally assumed that the function of flowers,
from the start, was to involve mobile animals in their
reproduction processes. Pollen can be scattered without
bright colors and obvious shapes. Expending energy on
these structures would appear to be a liability, unless they
provide sigriificant benefit.
Island genetics provides one proposed explanation for
the sudden, fully developed appearance of flowering
plants. Island genetics is believed to be a common source
of speciation in general, especially when it comes to radical
adaptations which seem to have required inferior
transitional forms. Flowering plants may have evolved in
an isolated setting like an island or island chain, where the
plants bearing them were able to develop a highly
specialized relationship with some specific animal. Such a
relationship, with a hypothetical wasp carrying pollen from
one plant to another much the way fig wasps do today,
could result in both the plant(s) and their partners
developing a high degree of specialization. Note that the
wasp example is not incidental; bees, which apparently
evolved specifically due to mutualistic plant relationships,
are descended from wasps.
Animals are also involved in the distribution of seeds.
Fruit, which is formed by the enlargement flower parts, is
frequently a seed disbursal tool which depends upon
animals, who eat or otherwise disturb it, incidentally
scattering the seeds it contains. While many such
mutualistic relationships remain too fragile to survive
competition with mainland animals and spread, flowers
proved to be an unusually effective means of production,
spreading to become the dominant form of land plant life.
Flowers are derived from leaf and stem components,
arising from a combination of genes normally responsible
178 Plant Ecology

for forming new shoots. The most primitive flowers are


thought to have had a variable number of flower parts,
often separate from (but in contact with) each other. The
flowers would have tended to grow in a spiral pattern, to
be bisexual (in plants, this means both male and female
parts on the same flower), and to be dominated by the ovary
(female part). As flowers grew more advanced, some
variations developed parts fused together, with a much
more specific number and design, and with either specific
sexes per flower or plant, or at least ovary inferior".
II

Flower evolution continues to the present day; modern


flowers have been so profoundly influenced by humans
that some of them cannot be pollinated in nature. Many
modern, domesticated flowers used to be simple weeds,
which only sprouted when the ground was disturbed.
Some of them tended to grow with human crops, perhaps
already having symbiotic companion plant relationships
with them, and the prettiest did not get plucked because of
their beauty, developing a dependence upon and special
adaptation to human affection.
REFERENCES

Gifford, Ernest M., Adriance S. Foster. Morphology and Evolution of


Vascular Plants. Third edition. WH Freeman and Company, New
York. 1989.
Cronquist, Arthur. An Integrated System of Classification of Flowering
Plants. Columbia Univ. Press, New York. 1981.
Raven, P.H., R.F. Evert, S.E. Eichhorn. Biology of Plants, 7th Edition. W.H.
Freeman. 2004.
Simpson, M.G. Plant Systematics. Elsevier Academic Press. 2006.
Stewart W. N. & Rothwell G. W. Paleobotany and the Evolution of Plants.
Cambridge Univ. Press, NY, USA. 521pp.
Taylor T. N. & Taylor E. L. The Biology and Evolution of Fossil Plants.
Prentice Hall, NJ, USA. 982pp. 1993.
8
Plant Community and
Ecosystem Dynamics

A community is the set of all populations that inhabit a


certain area. Communities can have different sizes and
boundaries. These are often identified with some difficulty.
An ecosystem is a higher level of organisation the
community plus its physical environment. Ecosystems
include both the biological and physical components
affecting the community/ecosystem. Ecologists find that
within a community many populations are not randomly
distributed. This recognition that there was a pattern and
process of spatial distribution of species was a major
accomplishment of ecology. Two of the most important
patterns are open community structure and the relative
rarity of species within a community.
If species within a community have similar geographic
range and density peaks, the community is said to be a
closed community, a discrete unit with sharp boundaries
known as ecotones. An open community, however, has its
populations without ecotones and distributed more or less
randomly. In a forest, where we find an open community
structure, there is a gradient of soil moisture. Plants have
different tolerances to this gradient and occur at different
places along the continuum. Where the physical
environment has abrupt transitions, we find sharp
boundaries developing between populations. For example,
180 Plant Ecology

an ecotone develops at a beach separating water and land.


Open structure provides some protection for the
community. Lacking boundaries, it is harder for a
community to be destroyed in an all or nothing fashion.
Species can come and go within communities over time, yet
the community as a whole persists. In general, communities
are less fragile and more flexible than some earlier concepts
would suggest.
Most species in a community are far less abundant than
the dominant species that provide a community its name:
for example oak-hickory, pine, etc. Populations of just a few
species are dominant within a community, no matter what
community we examine. Resource partitioning is thought
to be the main cause for this distribution.
CLASSIFICATION OF COMMUNITIES

There are two basic categories of communities: terrestrial


(land) and aquatic (water). These two basic types of
community contain eight smaller units known as biomes.
A biome is a large-scale category containing many
communities of a similar nature, whose distribution is
largely controlled by climate:
Terrestrial biomes: tundra, grassland, desert, taiga,
temperate forest, tropical forest.
Aquatic biomes: marine, freshwater.
Terrestrial Biomes
Tundra and Desert
The tundra and desert biomes occupy the most extreme
environments, with little or no moisture and extremes of
temperature acting as harsh selective agents on organisms
that occupy these areas. These two biomes have the fewest
numbers of species due to the stringent environmental
conditions. In other words, not everyone can live there due
to the specialised adaptations required by the environment.
Plant Community and Ecosystem Dynamics 181

Tropical Rain Forests


Tropical rain forests occur in regions near the equator. The
climate is always warm (between 20° and 25° C) with plenty
of rainfall (at least 190 em/year). The rain forest is probably
the richest biome, both in diversity and in total biomass.
The tropical rain forest has a complex structure, with many
levels of life. More than half of all terrestrial species live in
this biome. While diversity is high, dominance by a
particular species is low.
While some animals live on the ground, most rain forest
animals live in the trees. Many of these animals spend their
entire life in the forest canopy. Insects are so abundant in
tropical rain forests that the majority have not yet been
identified. Charles Darwin noted the number of species
found on a single tree, and suggested the richness of the
rain forest would stagger the future systematist with the
size of the catalogue of animal species found there.
Termites are critical in the decomposition and nutrient
cycling of wood. Birds tend to be brightly colored, often
making them sought after as exotic pets. Amphibians and
reptiles are well represented. Lemurs, sloths, and monkeys
feed on fruits in tropical rain forest trees. The largest
carnivores are the cats. Encroachment and destruction of
habitat put all these animals and plants at risk.
Epiphytes are plants that grow on other plants. These
epiphytes have their own roots to absorb moisture and
minerals, and use the other plant more as an aid to grow
taller. Some tropical forests in India, Southeast Asia, West
Africa, Central and South American are seasonal and have
trees that shed leaves in dry season. The warm, moist
climate supports high productivity as well as rapid
decomposition of detritus.
With its yearlong growing season, tropical forests have
a rapid cycling of nutrients. Soils in tropical rain forests
tend to have very little organic matter since most of the
182 Plant Ecology

organif: carbon is tied up in the standing biomass of the


plants. These tropical soils, termed laterites, make poor
agricultural soils after the forest has been cleared. About 17
million hectares of rain forest are destroyed each year.
Estimates indicate the forests will be destroyed within 100
years. Rainfall and climate patterns could change as a
result.

Temperate Forests
The temperate forest biome occurs south of the taiga in
eastern North America, eastern Asia, and much of Europe.
Rainfall is abundant and there is a well-defined growing
season of between 140 and 300 days. The eabtern United
States and Canada are covered by this biome's natural
vegetation, the eastern deciduous forest. Dominant plants
include beech, maple, oak; and other deciduous hardwood
trees. Trees of a deciduous forest have broad leaves, which
they lose in the fall and grow again in the spring.
Sufficient sunlight penetrates the canopy to support a
well-developed understory compQsed of shrubs, a layer of
herbaceous plants, and then often a ground cover of mosses
and ferns. This stratification beneath the canopy provides
a numerous habitats for a variety of insects and birds. The
deciduous forest also contains many members of the rodent
family, which serve as a food source for bobcats, wolves,
and foxes. This area also is a home for deer and black bears.
Winters are not as cold as in the taiga, so many amphibian
and reptiles are able to survive.

Shrubland (Chaparral)
The shrubland biome is dominated by shrubs with small
but thick evergreen leaves that are often coated with a thick,
waxy cuticle, and with thick underground stems that
survive the dry summers and frequent fires. Shrublands
occur in parts of South America, western Australia, central
Chile, and around the Mediterranean Sea. Dense shrubland
Plant Community and Ecosystem Dynamics 183

in California, where the summers are hot and very dry, is


known as chaparral. This Mediterranean-type shrubland
lacks an understory and ground litter, and is also highly
flammable. The seeds of many species require the heat and
scarring action of fire to induce germination.

Grasslands
Grasslands occur in temperate and tropical areas with
reduced rainfall or prolonged dry seasons. Grasslands
occur in the Americas, Africa, Asia, and Australia. Soils in
this region are deep and rich and are excellent for
agriculture. Grasslands are almost entirely devoid of trees,
and can support large herds of grazing animals. Natural
grasslands once covered over 40 percent of the earth's land
surface. In temperate areas where rainfall is between 10 and
30 inches a year, grassland is the climax community
because it is too wet for desert and too dry for forests.
Most grasslands have now been utilised to grow crops,
especially wheat and corn. Grasses are the dominant plants,
vvhile grazing and burrowing species are the dominant
animals. The extensive root systems of grasses allows them
to recover quickly from grazing, flooding, drought, and
sometimes fire.
Temperate grasslands include the Russian steppes, the
South American pampas, and North American prairies. A
tall-grass prairie occurs where moisture is not quite
sufficient to support trees.
Animal life includes mice, prairie dogs, rabbits, and
animals that feed on them (hawks and snakes). Prairies
once contained large herds of buffalo and pronghorn
antelope, but with human activity these once great herds
ahve dwindled.
The savanna is a tropical grassland that contains some
trees. The savanna contains the greatest variety and
numbers of herbivores (antelopes, zebras, and wildebeests,
184 Plant Ecology

among others). This environment supports a large


population of carnivores (lions, cheetahs, hyenas, and
leopards). Any plant litter not consumed by grazers is
attacked by termites and other decomposers. Once again,
human activities are threatening this biome, reducing the
range for herbivores and carnivores.

Deserts
Deserts are characterised by dry conditions (usually less
than 10 inches per year; 25 cm) and a wide temperature
range. The dry air leads to wide daily temperature
fluctuations from freezing at night to over 120 degrees
during the day. Most deserts occur at latitudef. of 300 N or
S where descending air masses are dry. Some deserts occur
in the rainshadow of tall mountain ranges or in coastal
areas near cold offshore currents. Plants in this biome have
developed a series of adaptations to conserve water and
deal with these temperature extremes. Photosynthetic
modifications are another strategy to life in the drylands.
The Sahara and a few other deserts have almost no
vegetation. Most deserts, however, are home to a variety of
plants, all adapted to heat and lack of abundant water
(succulents and cacti). Animal life of the Sonoran desert
includes arthropods (especially insects and spiders),
reptiles (lizards and snakes), running birds (the roadrunner
of the American southwest and Warner Brothers cartoon
fame), rodents (kangaroo rat and pack rat), and a few larger
birds and mammals (hawks, owls, and coyotes).

Taiga (Boreal Forest)

The taiga is a coniferous forest extending across most of the


northern area of northern Eurasia and North America. This
forest belt also occurs in a few other areas, where it has
different names: the montane coniferous forest when near
mountain tops; and the temperate rain forest along the
Pacific Coast as far south as California.
Plant Community and Ecosystem Dynamics 185

The taiga receives between 10 and 40 inches of rain per


year and has a short growing season. Winters are cold and
short, while summers tend to be cool. The taiga is noted for
its great stands of spruce, fir, hemlock, and pine. These trees
have thick protective leaves and bark, as well as needlelike
(evergreen) leaves can withstand the weight of
accumulated snow.
Taiga forests have a limited understory of plants, and
a forest floor covered by low-lying mosses and lichens.
Conifers, alders, birch and willow are common plants;
wolves, grizzly bears, moose, and caribou are common
animals. Dominance of a few species is pronounced, but
diversity is low when compared to temperate and tropical
biomes.

Tundra
The tundra covers the northernmost regions of North
America and Eurasia, about 20% of the Earth's land area.
This biome receives about 20 cm (8-10 inches) of rainfall
annually. Snow melt makes water plentiful during summer
months. Winters are long and dark, followed by very short
summers. Water is frozen most of the time, producing
frozen soil, permafrost. Vegetation includes no trees, but
rather patches of grass and shrubs; grazing musk ox,
reindeer, and caribou exist along with wolves, lynx, and
rodents. A few animals highly adapted to cold live in the
tundra year-round (lemming, ptarmigan).
During the summer the tundra hosts numerous insects
and migratory animals. The ground is nearly completely
covered with sedges and short grasses during the short
summer. There are also plenty of patches of lichens and
mosses. Dwarf woody shrubs flower and produce seeds
quickly during the short growing season. The alpine tundra
occurs above the timberline on mountain ranges, and may
contain many of the same plants as the arctic tundra.
186 Plant Ecology

Climate and Terrestrial Biomes


Climate controls biome distribution by an altitudinal
gradient and a latitudinal gradient. With increases of either
altitude or latitude, cooler and drier conditions occur.
Cooler conditions can cause aridity since cooler air can hold
less water vapor than can warmer air.
Deserts can occur in warm areas due to a blockage of
air circulation patterns that form a rain shadow, or from
atmospheric circulation patters. Warm air rises, producing
low pressure areas. Cooler air sinks, producing high
pressure areas. The tropics tend to be atmospheric low
pressure zones the arctic areas atmospheric highs. Relative
humidity is a measure of how much water an air mass at
a given temperature can hold. In short, warm air can hold
more moisture than can cold air. This basic physical feature
of air helps explain the distribution of some of the world's
great deserts.
The warm, moist air masses in the tropics rise upward
in the atmosphere as they heat. The pressure of air rising
forces air in the upper atmosphere to flow away north and
south. This air at higher elevations is cooler and loses much
of its moisture as rainfall. When the air masses begin to
descend they heat up and begin to draw moisture from the
lands they descend upon, at 30 degrees north and south of
the equator. Many of the world's deserts are at
approximately 30 degrees latitude. Rain shadow deserts
also form when cool, dry air masses descend after passing
over a tall mountain range, such as the Coast Range and
Sierras in California. The Sonoran desert in Arizona is a
doubly caused desert, being at 30 degrees latitude as well
as in the rain shadow of California mountains.

Aquatic Biomes
Conditions in water are generally less harsh than those on
land. Aquatic organisms are buoyed by water support, and
Plant Community and Ecosystem Dynamics 187

do not usually have to deal with desiccation. Despite


covering 71 % of the Earth's surface, areas of the open ocean
are a vast aquatic desert containing few nutrients and very
little life. Clearcut biome distinctions in water, like those on
land, are difficult to make. Dissolved nutrients controls
many local aquatic distributions. Aquatic communities are
classified into: freshwater communities and marine
(saltwater or oceanic) communities.

Marine Biame
The marine biome contains more dissolved minerals than
the freshwater biome. Over 70% of the Earth's surface is
covered in water, by far the vast majority of that being
saltwater. There are two basic cat~gories to this biome:
benthic and pelagic. Benthic communities (bottom
dwellers) are subdiv.ided by depth: the shore/shelf and
deep sea. Pelagic communities include planktonic (floating)
and nektonic (swimming) organisms. The upper 200 meters
of the water column is the euphotic zone to which light can
penetrate.

Coastal Communities
Estuaries are bays where rivers empty into the sea. Erosion
brings down nutrients and tides wash in salt water; forms
nutrient trap. Estuaries have high production for organisms
that can tolerate changing salinity. Estuaries are called
"nurseries of the sea" because many young marine fish
develop in this protected environment before moving as
adults into the wide open seas.

Seashores
Rocky shorelines offer anchorage for sessile organisms.
Seaweeds are main photosynthesizers and use holdfasts to
anchor. Barnacles glue themselves to stone. Oysters and
mussels attach themselves by threads. Limpets and
188 Plant Ecology

periwinkles either hide in crevices or fasten flat to rocks.


Sandy beaches and shores are shifting strata. Permanent
residents therefore burrow underground. Worms live
permanently in tubes. Amphipods and ghost crabs burrow
above high tide and feed at night.

Coral Reefs
Areas of biological abundance in shallow, warm tropical
waters. Stony corals have calcium carbonate exoskeleton
and may include algae. Most form colonies; may associate
with zooxanthellae dinoflagellates. Reef is densely
populated with animal life. The Great Barrier Reef of
Australia suffers from heavy predation by crown-of-thorns
sea star, perhaps because humans have harvested its
predator, the giant triton.

Oceans
Oceans cover about three-quarters of the Earth's surface.
Oceanic organisms are placed in either pelagic (open water)
or benthic (ocean floor) categories. Pelagic division is
divided into neritic and three levels of pelagic provinces.
Neritic province has greater concentration of organisms
because sunlight penetrates; nutrients are found here.
Epipelagic zone is brightly lit, has much photosynthetic
phytoplankton, that support zooplankton that are food for
fish, squid, dolphins, and whales. Mesopelagic zone is .'
semi-dark and contains carnivores; adapted organisms tend
to be translucent, red colored, or luminescent; for example:
shrimps, squids, lantern and hatchet fishes. The
bathypelagic zone is completely dark and largest in size; it
has strange-looking fish. Benthic division includes
organisms on continental shelf (sublittoral), continental
slope (bathyal), and the abyssal plain.
Sublittoral zone harbors seaweed that becomes sparse
where deeper; most dependent on slow rain of plankton
Plant Community and Ecosystem Dynamics 189

and detritus from sunlit water above. Bathyal zone


continues with thinning of sublittoral organisms. Abyssal
zone is mainly animals at soil-water interface of dark
abyssal plain; in spite of high pressure, darkness and
coldness, many invertebrates thrive here among sea urchins
and tubeworms.
Thermal vents along oceanic ridges form a very unique
community. Molten magma heats seawater to 350°C,
reacting with sulfate to form hydrogen sulfide (H2S).
Chemosynthetic bacteria obtain enEJrgy by oxidizing
hydrogen sulfide. The resulting food chain supports a
community of tubeworms and clams.

Freshvvater Bio~e

The fresh\water biome is subdivided into two zones:


running waters and standing waters. Larger bodies of
freshwater are less prone to stratification (where oxygen
decreases with depth). The upper layers have abundant
oxygen, the lowermost layers are oxygen-poor. Mixing
between upper and lower layers in a pond or lake occurs
during seasonal changes known as spring and fall overturn.
Lakes are larger than ponds, and are stratified in
summer and winter. The epilimnion is the upper surface
layer. It is warm in summer. The hypolimnion is the cold
lower layer. A sudden drop in temperature occurs at the
middle of the thermocline. Layering prevents mixing
between the lower hypolimnion (rich in nutrients) and the
upper epilimnion (which has oxygen absorbed from its,
surface). The epilimnion warms in spring and cools in fall,
causing a temporary mixing. As a consequence,
phytoplankton become more abundant due to the increased
amounts of nutrients.
Life zones also exist in lakes and ponds. The littoral
zone is closest to shore. The limnetic zone is the sunlit body
of the lake. Below the level of sunlight penetration is the
190 Plant Ecology

dark profundal zone. At the soil-water interface we find the


benthic zone. The term benthos is applied to animals and
other organisms that live on or in the benthic zone. Rapidly
flowing, bubbling streams have insects and fish adapted to
oxygen-rich water. Slow moving streams have aquatic life
more similar to lake and pond life.
COMMUNITY DENSITY

Communities are made up of species adapted to the


conditions of that community. Diversity and stability help
define a community and are important in environmental
studies. Species diversity decreases as we move away from
the tropics. Species diversity is a measure of the different
types of organisms in a community (also referred to as
species richness).
Latitudinal diversity gradient refers to species richness
decreasing steadily going away from the equator. A hectare
of tropical rain forest contains 40-100 tree species, while a
hectare of temperate zone forest contains 10:-30 tree species.
In marked contrast, a hectare of taiga contains only a paltry
1-5 species! Habitat destruction in tropical countries will
cause many more extinctions per hectare than it would in
higher latitudes.
Environmental stability is greater in tropical areas,
where a relatively stable/constant environment allows
more different kinds of species to thrive. Equatorial
communities are older because they have been less
disturbed by glaciers and other climate changes, allowing
time for new species to evolve. Equatorial areas also have
a longer growing season. The depth diversity gradient is
found in aquatic communities. Increasing species richness
with increasing water depth. This gradient is established by
environmental stability and the increasing availability of
nutrients.
Community stability refers to the ability of
communities to remain unchanged over time. During the
Plant Community and Ecosystem Dynamics 191

1950s and 1960s, stability was equated to diversity: diverse


communities were also stable communities. Mathematical
modeling during the 1970s showed that increased diversity
can actually increase interdependence among species and
lead to a cascade effect when a keystone species is removed.
Thus, the relation is more complex than previously thought.
CHANGE IN COMMUNITIES

Biological communities, like the organisms that comprise


them, can and do change over time. Ecological time focuses
on community events that occur over decades or centuries.
Geological time focuses on events lasting thousands of
years or more.
Community succession is the sequential replacement of
species by immigration of new species and local extinction
of older ones following. a disturbance that creates
unoccupied habitats for colonization. The initial rapid
colonizer species are the pioneer community. Eventually a
climax community of more or less stable but slower
growing species eventually develops.
During succession productivity declines and diversity
increases. These trends tend to increase the biomass (total
weight of living tissue) in a community. Succession occurs
because each community stage prepares the environment
for the stage following it.
Primary succession begins with bare rock and takes a
very long time to occur. Weathering by wind and rain plus
the actions of pioneer species such as lichens and mosses
begin the buildup of soil. Herbaceous plants, including the
grasses, grow on deeper soil and shade out shorter pioneer
species. Pine trees or deciduous trees eventually take root
and in most biomes will form a climax community of plants
that are stabile in the environment. The young produced by
climax species can live in that environment, unlike the
young produced by successional species.
192 Plant Ecology

Secondary succession occurs when an environment has


been disturbed, such as by fire, geological activity, or
human intervention. This form of succession often begins
in an abandoned field with soil layers already in place.
Compared to primary succession, which must take long
periods of time to build or accumulate soil, secondary
succession occurs rapidly. The herbaceous pioneering
plants give way to pines, which in turn may give way to
a hardwood deciduous forest.
Early researchers assumed climax communities were
determined for each environment. Today we recognise the
outcome of competition among whatever species are
present as establishing the climax community.
Climax communities tend to be more stable than
successional communities. Early stages of succession show
the most growth and are most productive. Pioneer
communities lack diversity, make poor use of inputs, and
lose heat and nutrients. As succession proceeds, species
variety increases and nutrients are recycled more. Climax
communities make fuller use of inputs and maintain
themselves, thus, they are more stable. Human activity
replaces climax communities with simpler communities.
Communities are composed of species that evolve, so
the community must also evolve. Comparing marine
communities of 500 million years ago with modern
communities shows modem communities composed of
quite different organisms. Modem communities also tend
to be more complex, although this may be a reflection of
the nature of the fossil record as well as differences between
biological and fossil species.
The basic effect of human activity on communities is
community simplification, an overall reduction of species
diversity. Agriculture is a purposeful human intervention
in which we create a monoculture of a single favored (crop)
species such as com. Most of the agricultural species are
derived from pioneering communities. Inadvertent human
Plant Community and Ecosystem Dynamics 193

intervention can simplify communities and produce


stressed communities that have fewer species as well as a
superabundance of some species. Disturbances favor early
successional species that can grow and reproduce rapidly.
ECOSYSTEMS AND COMMUNITIES

Ecosystems include both living and nonliving components.


These living, or biotic, components include habitats and
niches occupied by organisms. Nonliving, or abiotic,
components include soil, water, light, inorganic nutrients,
and weather. An organism's place of residence, where it can
be found, is its habitat. A niche is is often viewed as the role
of that organism in the community, factors limiting its life,
and how it acquires food. Producers, a major niche in all
ecosystems, are autotrophic, usually photosynthetic,
organisms. In terrestrial ecosystems, producers are usually
green plants. Freshwater and marine ecosystems frequently
have algae as the dominant producers.
Consumers are heterotrophic organisms that eat food
produced by another organism. Herbivores are a type of
consumer that feeds directly on green plants. Since
herbivores take their food directly from the producer level,
we refer to them as primary consumers. Carnivores feed on
other animals and are secondary or tertiary consumers.
Omnivores, the feeding method used by humans, feed on
both plants and animals. Decomposers are organisms,
mostly bacteria and fungi that recycle nutrients from
decaying organic material. Decomposers break down
detritus, nonliving organic matter, into inorganic matter.
Small soil organisms are critical in helping bacteria and
fungi shred leaf litter and form rich soil. .
Even if communities do differ in structure, they have
some common uniting processes such as energy flow and
matter cycling. Energy flows move through feeding
relationships. The term ecological niche refers to how an
organism functions in an ecosystem. Food webs, food
194 Plant Ecology

chains, and food pyramids are three ways of representing


energy flow. Producers absorb solar energy and convert it
to chemical bonds from inorganic nutrients taken from
environment. Energy content of organic food passes up
food chain; eventually all energy is lost as heat, therefore
requiring continual input. Original inorganic elements are
mostly returned to soil and producers; can be used again
by producers and no new input is required.
Energy flow in ecosystems, as with all other energy,
must follow the two laws of thermodynamics. Recall that
the first law states that energy is neither created nor
destroyed, but instead changes from one form to another
(potential to kinetic). The second law mandates that when
energy is transformed from one form to another, some
usable energy is lost as heat. Thus, in any food chain, some
energy must be lost as we move up the chain.
The ultimate source of energy for nearly all life is the
Sun. Recently, scientists discovered an exception to this
once unchallenged truism: communities of organisms
around ocean vents where food chain begins with
chemosynthetic bacteria that oxidise hydrogen sulfide
generated by inorganic chemical reactions inside the
Earth's crust. In this special case, the source of energy is the
internal heat engine of the Earth.
Food chains indicate who eats whom in an ecosystem.
Represent one path of energy flow through an ecosystem.
Natural ecosystems have numerous interconnected food
chains. Each level of producer and consumers is a trophic
level. Some primary consumers feed on plants and make
grazing food chains; others feed on detritus. The population
size in an undisturbed ecosystem is limited by the food
supply, competition, predation, and parasitism. Food webs
help determine consequences of perturbations: if titmice
and vireos fed on beetles and earthworms, insecticides that
killed beetles would increase competition between birds
and probably increase predation of earthworms, etc.
Plant Community and Ecosystem Dynamics 195

The trophic structure of an ecosystem forms an


ecological pyramid. The base of this pyramid represents the
producer trophic level. At the apex is the highest level
consumer, the top predator. Other pyramids can be
recognised in an ecosystem. A pyramid of numbers is based
on how many organisms occupy each trophic level. The
pyramid of biomass is calculated by multiplying the
average weight for organisms times the number of
organisms at each trophic level. An energy pyramid
illustrates the amounts of energy available at each
successive trophic level. The energy pyramid always shows
a decrease moving up trophic levels because:
Only a certain amount of food is captured and eaten
by organisms on the next trophic level.
Some of food that is eaten cannot be digested and exits
digestive tract as undigested waste.
Only a portion of digested food becomes part of the
organism's body; rest is used as source of energy.
Substantial portion of food energy goes to build up
temporary ATP in mitochondria that is then used to
synthesize proteins, lipids, carbohydrates, fuel
contraction of muscles, nerve conduction, and other
functions.
Only about 10% of the energy available at a particular
trophic level is incorporated into tissues at the next
level. Thus, a larger population can be sustained by
eating grain than by eating grain-fed animals since 100
kg of grain would result in 10 human kg but if fed to
cattle, the result, by the time that reaches the human
is a paltry 1 human fg!
A food chain is a series of organisms each feeding on the
one preceding it. There are two types of food chain:
decomposer and grazer. Grazer food chains begin with
algae and plants and end in a carnivore. Decomposer chains
are composed of waste and decomposing organisms such
as fungi and bacteria.
196 Plant Ecology

Food chains are simplifications of complex


relationships. A food web is a more realistic and accurate
'depiction of energy flow. Food webs are networks of
feeding interactions among species. The food pyramid
provides a detailed view of energy flow in an ecosystem:
The first level consists of the producers. All higher levels
are consumers. The shorter the food chain the more energy
is available to organisms.
Most humans occupy a top carnivore role, about 2% of
all calories available from producers ever reach the tissues
of top carnivores. Leakage of energy occurs between each
feeding level. Most natural ecosystems therefore do not
have more than five levels to 'their food pyramids. Large
carnivores are rare because there is so little energy available
to them atop the pyramid.
Food generation by producers varies greatly between
ecosystems. Net primary productivity (NPP) is the rate at
which producer biomass is formed. Tropical forests and
swamps are the most productive terrestrial ecofystems.
Reefs and estuaries are the most productive aquatic
ecosystems. All of these productive areas are in danger
from human activity. Humans redirect nearly 40% of the
net primary productivity and directly or indirectly use
nearly 40% of all the land food pyramid.
REFERENCES

Grime, J.P. "Biodiversity and Ecosystem Function: The Debate Deepens."


Science Vol. 277. no. 533029 Aug 1997 pp. 1260 -1261. 25 May 2007
Lindeman, R.L. 194;2. 'The trophic-dynamic aspect of ecology". Ecology
23: 399-418.
Patten, B.C. 1959. "An Introduction to the Cybernetics of the Ecosystem:
The Trophic-Dynamic Aspect". Ecology 40, no. 2.: 221-231.
Robert Ulanowicz (1997). Ecology, the Ascendant Perspective. Columbia
. Univ. Press.
Tansley, A.G. 1935. "The use and abuse of vegetational concepts and
terms". Ecology 16: 284-307.
9
Ecology of Weeds and Invasive Plants

A "plant growing out .of place," that is, plants growing


where they are not wanted, at least by some people, is a
common, accepted explanation for what weeds are. This
notion of undesirability imparts so much human value to
the idea of weediness that it is usually necessary to
recognise who is making the determination as well as the
characteristics of the plants themselves.
"

Weeds are recognized worldwide as an important type


of undesirable, economiC pest, especially in agriculture.
However, the value of any plant is unquestionably
determined by the perceptions of its viewers. These
perceptions also influence the human activities directed at
this category of vegetation.
Weeds are little more than plants that have aroused a
level of human dislike at some particular place or time. \
Unfortunately, the anthropomorphic view of weeds
provides little insight into why and where they exist, their '
interactions and associations with crops, native plants, and
other organisms, or even how to manage them effectively.
Weeds are found worldwide and have proven to be
successful organisms in the environments that they inhabit.
Therefore, it is important to explore whether weeds posses
common traits that distinguish them from other plants or
whether they are only set apart by local notions of
usefulness.
198 Plant Ecology

Characteristics of Weeds

A list of biological characteristics that describe weeds was


proposed in the 1970s and continues to be used today, but
it seems unlikely that any plant species could possess all of
those "ideal" weedy traits. However, Herbert Baker,
botanist and originator of the list, suggests that a species
might possess various combinations of the characteristics,
resulting in a range of weediness from minor to major
weeds. These are:
Germination requirements fulfilled in many
environments
Discontinuous germination (internally controlled) and
great longevity of seed
Rapid growth through vegetative phase to flowering
Continuous seed production for as long as growing
conditions permit
Self-compatibility but not complete autogamy or
apomixis
Cross-pollination, when it occurs, by unspecialized
visitors or wind
Very high seed output in favorable environmental
circumstances
Production of some seed in a wide range of
environmental conditions; tolerance and plasticity
Adaptations for short-distance dispersal and long-
distance dispersal
If perennial, vigorous vegetative reproduction or
regeneration from fragments
If perennial, brittleness, so as not to be drawn from the
ground easily
Ability to complete interspecifically by special means.
Ecological success in the form of weediness cannot be
measured solely from the perspective of noxiousness. The
Ecology of Weeds and_Invasive Plants 199

number of individuals, the range of habitats occupied, and


the ability to continue the species through time must be
considered foremost when evaluating success of a species.
Weeds can be described in either anthropomorphic or
biological terms. Weeds emerge from such descriptions as
organisms that may possess a particular suite of biological
characteristics but also have the distinction of negative
human selection. Thus, a definition of a weed as any plant
that is objectionable or interferes with the activities or
welfare of man seems to describe sufficiently this category
of vegetation. A sample of definitions of weeds published
over the past century was presented by Randall, who also
argued that the most important criterion was problem-
causing plants that interfere with land use.
Zimmerman believes that the term "weed" should be
used to describe plants that (1) colonize disturbed habitats,
(2) are not members of the original plant community, (3) are
locally abundant, and (4) are economically of little value (or
are costly to control). Aldrich defines weeds as plants that
originated under a natural environment and, in response to
(human) imposed or natural conditions, are interfering
associates of crops and human activities.
Each of these definitions implies that weeds have some
common biological traits but also a level of relative
undesirability as determined by particular people. Whether
or not a plant is a weed depends on the context in which
someone finds it and on the perspectives and objectives of
those involved in dealing with it. Rejma'nek, on the other
hand, believes that weeds, colonisers, and naturalized
species (including invasive plants) reflect three overlapping
concepts. In his view, weeds are plants growing where they
are not desired (anthropomorphic defi-nition), colonisers
occur early in succession (ecological definition), and
invasive plants are plants that become locally established
and spread to areas where they are not native
(biogeographical definition).
200 Plant Ecology

The most important criterion for weediness is


interference at some place or time with the values and
activities of people-farmers, foresters, land managers, and
many other segments of human society. However, the
abundance of weeds is often of more concern than the mere
presence of them. For instance, farmers and land managers
are usually less concerned about the occurrence of a few
isolated plants in a field, even noxious ones, than the
occupation of land by vast numbers of weeds. Therefore,
the relative abundance of plants, their location, and the
potential use of the land they occupy should also be
considered in weed definitions. When abundance is
applied as a criterion for weediness, it implies a condition
of the land as well as a class of vegetation and a form of
human discrimination. Weed abundance also may be an
indicator or symptom of land mismanagement or neglect.

Agrestals
Agrestals are weeds of tilled, arable land. They require the
nearly continual disturbance of agriculture to occupy the
land. Holzner et al. indicate that every cropping system, for
example, cereals, root crops, and orcl:tards, also has its
special complement of weeds, which may be·eifuer natfve·:::
plants or exotics that have been naturalised into the local
flora. A list of the 76 worst agricultural weeds in the world
was developed by Holm and his associates and has become
the standard by which agrestals are compared. The top 18
weeds on this list are given in Table 1. An additional 104
of the weeds that cause the greatest impacts on agriculture
was reviewed by Holm et al. in 1997. As a group these 180
agricultural weeds are estimated to cause over 90% of the
loss of crop productivity worldwide.
Agrestals have evolved as either specialists or
colonizers during the course of agricultural history.
Specialised weeds (specialists) have evolved a narrow
adaptation to a single crop or sometimes crop cu1tivar and
Ecology of Weeds and Invasive Plants 201

its particular growing conditions. Perhaps the most extreme


• example of how human activities influence weed species
distribution and composition are crop mimics. These are
weeds that have evolved life cycles or morphological
features so similar to a crop that the two species cannot be
distinguished or separated easily.

Table 1. Scientific and Common names of certain annual weed species


considered the world's 18 worst

Species Common name


Amaranthus hybridus Smooth pigweed
Amaranthus spinosus Spiny amaranthus
Avenafatua Wild oat
Chenopodium album Common lambsquarters
Convolvulus arvensis Field bindweed
Cynodon dactylon Bermudagrass
Cyperus esculentus Yellow nutsedge
Cyperus rotundus Purple nutsedge
Digitaria sanguinalis Large crabgrass
Echinochloa colonum Junglerice
Echinochloa crus-galli Bamyardgrass
Eichhornia crassipes Waterhyacinth
Eleusine indica Goosegrass
Imperata cylindrica Cogongrass
Paspalum conjugatum Sour paspalum
Portulaca oleracea Common purslane
Rottboellia exaltata Itchgrass
Sorghum halepense Johnsongrass

Since agrestals that are specialists have evolved along with


the cultural practices of a particular crop, any change in
practices usually disfavors the weed. Colonisers, on the
other hand, are plants with characteristics that allow them
to rapidly occupy and dominate disturbed areas. Weeds are
major constraints to crop production, yet as primary
producers, they also can be important components in an
202 Plant Ecology

. agroecosystem. It is in this context that weeds are


sometimes perceived as an ecological "good".
Awareness of the importance of weeds on arable land
for their role in other trophic levels is growing as natural
landscapes become rare or disappear due to the expansion
of human-occupied landscapes. The weed flora in many
parts of the world has changed over the past century, with
some species declining in abundance while others have
increased. These changes in the weed flora reflect improved
agricultural efficiency, the use of different crops in arable
rotations, and the use of more broad-spectrum herbicide
combinations..
Many weed species of arable land support a high
diversity of insects, so the reduction in abundance of weed
host plants can affect associated insects and, therefore, the
abundance of other taxa. For example, in the United
Kingdom a number of insect groups and farmland-
associated birds (notably the grey partridge, Perdix perdix)
have undergone marked population decline, which is
associated with changes in agricultural practices over the
past 30 years. Thus, it seems that weeds may have a general
role in supporting biodiversity within agroecosystems.
CHARACfERISTICS OF INVASIVE PLANTS

Invasive plants, unlike agricultural weeds, are generally


defined as those that can successfully establish, become
naturalised, and spread to new natural habitats apparently
without further assistance from humans. They are also
generally nonnative or exotic in the new habitat and are
often relatively new introductions to an ecoregion. Invasive
plants respond readily to human-induced changes in the
environment such as disturbance but also may initiate
environmental change through their dominance on the
landscape. In addition, the spatial and temporal extent of
their impact may be expressed at scales ranging from local
to global.
Ecology of Weeds and Invasive Plants 203

Some ecological impacts believed to be caused by


invasive plants are:.;:; follows:
Reduction of biodiversity
Loss or encroachment upon endangered and
threatened species and their habitats
Loss of habitat for native insects, birds, and other
wildlife
Loss of food sources for wildlife
Changes to natural ecological processes such as plant
community succession
Alteratio~s to the frequency and intensity of natural
fires
Disruptions of native plant-animal associations such as
pollination, seed dispersal, and hvst-plant
relationships
It is widely believed that the most effective way to limit
plant invasions is to prevent the introduction of exotic
species, which may be difficult because of the ongoing
expansion in global travel and trade, changes in
environments at all scales (local to global), and increasing
development of land for human use.
Although the traits of an "ideal weed" have also been
ascribed to invasive plants, few empirical studies have
tested this concept. The biological characteristics of
invasive plants appear in many cases to be dependent upon
the habitat in which they occur. Thus, general descriptions
of invasive plants remain inconclusive. Some useful
generalisations have been made, however, from reviews of
empirical evidence or broadscale analyses of floras or
databases. For example, Reichard and Hamilton, using a
regression tree analysis of biological and environmental
traits of invasive plants, suggest that species known to be
invasive elsewhere should be limited in introduction to a
new area with a similar environment, where they might
.Iso be invasive.
204 Plant Ecology

-CLASSIFICATION SYS~ OF WEEDS AND INVASIVE PLANTS

Massive amounts o~ money, time, and energy are expended


on weeds and inv~.sive plants because of their economic
and ecological costs and impacts on agricultural and
natural systems. Because of the magnitude of these effects,
it is important that scientists and land managers consider
carefully the metaphors they use to describe these two
categories of vegetation. Larson points out that metaphors
allow people to understand abstract or perplexing subjects
in term of something they already know about, a common
referent.
Thus, weeds and especially invasive plants are often
described in militaristic terms, which probably date to
Elton's classic The Ecology of Invasions by Animals and
Plants. Davis points out that such terms as alien, exotic,
invader, and invasion commonly used by invasion
ecologists contrast markedly to ~e less evocative terms
such as coloniser, founding population, introduced plant,
nonnative, spread, or migration, which could be used to
describe weeds and "invasive" plants. It should be noted
that a similarly militaristic terminology has been used for
decades in the pest management field.
From a management point of view, there is little doubt
that the "invasion" terminology and metaphors have been
useful in pointing out the significance of weeds to land
managers and policymakers. From a strictly scientific point
of view, however, it is difficult to argue against returning
to the more value-neutral terminology used by Baker and
Stebbins in their early classic, The Genetics of Colonising
Species. Since this text is designed to fulfill a dual role for
both scientists and land mangers and because the notion of
"weed" is itself value laden, we have chosen to use the
language of both scientists and managers that is in
conventional use to discuss this important class of
vegetation.
Ecology of Weeds and Invasive Plants 205

Botanical classification is the systematic grouping of


plants using criteria that distinguish among types of
vegetation. These criteria may be biologically meaningful,
based on phylogenetic or evolutionary evidence, or
artificial and based on structural or other visible or
functional attributes. Some common methods used to
classify weeds are by taxonomic relationships, life history,
habitat, physiology, and degree of undesirability. Weeds
and invasive plants can also be classified by ecological
behavior related to invasion and evolutionary strategies
related to carbon allocation.

Taxonomic Classification
Systematics is the scientific study of biological organisms
and their evolutionary relationships. Ideally, organisms are
classified systematically according to their presumed
genetic relationships, although often this information is
unknown. The basis of modem classification is taxonomy,
the identification, naming, and grouping of plants
according to their traits in common. The accepted
taxonomic system used today classifies organisms into a
hierarchy of categories: kingdom, phylum (also called
division in some botany texts), class, order, family, genus,
and species.
Recent evidence has shown that an additional category,
the domain, occurs above the level of the kingdom; the
three recognised domains are Bacteria, Archaea, and
Eukarya. All land plants are placed in the domain Eukarya
and the kingdom Plantae. Most weeds occur in the phylum
Anthophyta (angiosperms, flowering plants), although
notable exceptions occur (e.g., some ferns, which are
seedless, and conifers, seed plants that have no flowers, are
considered weeds). Angiosperms are further divided into
the classes Dicotyledones (dicots) and Monocotyledones
(monocots).
206 Plant Ecology

The next level of classification is the order. Although


systematists do not agree on the exact number of orders, the
commonly accepted Cronquist system recognises 64 orders
of dicots and 19 orders of monocots. The orders are divided
further into families, which, like classes and orders, are
comprised of plants whose morphological similarities are
greater than their differEnces.
Approximately 383 angiosperm families are currently
recognised (318 dicot and 65 monocot). The level of genus
includes plants that have common characteristics and that
are presumed to be genetically related. The narrowest
category of classification is the species, which consists of
plants that can interbreed freely (the biological species
concept). For practical purposes, however, most species are
grouped largely on the basis of anatomical and
morphological characteristics (the morphological species
concept). At this point in taxonomic classification, the plant
group is given a name, called a scientific name or Latin
binomial, which consists of both the genus and species
names of the plant.
There are approximately 250,000 species of flowering
plants in the world (depending upon which authority is
used). However, less than 250 of these, about 0.1%, are
troublesome enough to be called major agricultural weeds
throughout the world. It is far more difficult to estimate the
number of invasive plant species in nonagricultural
habitats worldwide. In the United States, by one estimate,
introduced invasive plants comprise from 8 to 47% of the
total flora of most states. Of the 250 recognised major
agricultural weeds, nearly 70% occur in only 12 plant
families and over 40% are found in only two families,
Poaceae (grass family) and Asteraceae (aster or composite
family).
Although these observations are fruitful areas of
speculation for plant evolutionary biologists, it should be
noted that about 75% of world food production is provided
Ecology of Weeds and Invasive Plants 207

by only a dozen crops: barley, maize, millet, oats, rice,


sorghum, sugarcane, wheat, cassava, soybean, sweet
potato, and white potato. Eight of these crops (the first eight
in the list above) are also members of the grass family.
The distribution of both the world's worst agricultural
weeds and its major crops is quite taxonomically restricted,
again pointing to the extreme discrimination and selection
that humans apply to vegetation. It is sometimes necessary
to distinguish only broadly among weed species, for
example when broad-scale methods of weed control are
used. In such situations, distinction among grasses and
sedges (monocot) and broadleaf (dicot) plants may be
sufficient, and a much abbreviated system of classification
is satisfactory. Such a system was once in common use by
weed control specialists; a typical description of weeds by
this method is shown below: '
Dicots. Plants whose seedlings produce two cotyledons
or seed leaves. Usually typified by netted leaf venation
and flowering parts in fours, fives, or multiples thereof.
Examples include mustards (Brassica spp.),
nightshades (Solanum spp.), and morningglory
(Convolvulus spp.). Commonly called broadleaved
plants.
Monocots. Plants whose seedlings bear only one
cotyledon. Typified by parallel leaf venation and
flower parts in threes or multiples of three. Most weeds
are found in only two groups, grasses and sedges,
although other groups exist. Grasses. Leaves usually
have a ligule or at times an auricle. The leaf sheaths are
split around the stem with the stem being round or
flattened in cross section with hollow internodes.
Sedges. Leaves lack ligules and auricles and the leaf
sheaths are continuous around the stem. In many
species the stem is triangular in cross section with solid
internodes.
208 Plant Ecology

Classification by Life History


Another method used to classify weeds is by the life cycle
of the plant. The length of life, seas~ of growth, and\time
and method of reproduction are used to classify weeds in
this way. Annuals. An annual plant completes its life cycle
from seed to seed in one year or less. Annuals are often
divided into two groups, winter and summe!", according to
the plant's time of germination, maturation, and death:
Winter annuals. These plants uS!lally germinate in the
fall or winter, grow throughout the spring, and set seed.
and die by early summer..
Summer annuals. These plants germinate in the sP9Jtg,
grow throughout the summer, set seed by autumn, and
die before winter.
In mild climates, however, it is usual for some winter
annuals to germinate in late summer or autumn and for
some summer annuals to live throughout the winter.
Annual plants are the largest single category of weeds.
Biennials. These plants live longer than one but less than
two years. During the first growth phase, biennials develop
vegetatively from a seedling into a rosette. Because of this
growth habit, biennials sometimes can be confused with
winter annuals. After a cold period, vegetative growth
resumes, and floral initiation, seed production, and death
occur.
Biennials are often large plants when mature and have
thick fleshy roots. Relatively few weed species are
biennials, but some annual plants may behave as biennials
under certain conditions and some biennials may behave
as short-lived perennials in mild climates. Perennials.
Perennial plants live for longer than two years and may
reproduce several times before dying. These plants are
characterised by renewed vegetative growth year after year
from the same root system:
Ecology of Weeds and Invasive Plants 209

Simple herbaceous perennials. Simple herbaceous


perennials reproduce almost exclusively from seed and
normally do not reproduce vegetatively. However, if
the root system of these plants is injured or cut, each
piece usually regenerates into another plant.
Dandelion, plantain, and sulfur cinquefoil are
examples of simple herbaceous perennials.
Creeping herbaceous perennials. Creeping herbaceous
perennials survive over the winter and produce new
vegetative structures from asexual reproductive organs
such as rhizo(Iles, tubers, stolons, bulbs, corms, and
roots. These plants also reproduce sexually from seed.
Most aquatic. weeds, except algae, are creeping
perennial plants.
Woody plants. This is a special category of perennial
weed. Plants in this group are characterised by stems
that have secondary growth, producing wood and
bark, which results in an incremental increase in
diameter each year. Some tree, some shrub, and many
vine species are considered to be woody weeds.

Figure 1. Habitats of aquatic weeds

Classification by Habitat
Weeds can be classified according to where they grow.
210 Plant Ecology

Most weeds are terrestrial, that is, tound on land, but some
are restricted to the aquatic environment. Some weeds only
infest a particular crop or cropping system, complex of
plant communities, or growing condition. Therefore, it is
common to find lists and descriptions of weeds that are
usually found in particular environments, such as arable
land, pastures and rangeland, forests, ' rights-of-way, or
wildlands. Weeds can be classified by their habitat as
follows:
Aquatic weeds. Aquatic weeds are plants that are
modified structurally to live in water. They have been
categorised further based on their location in the
aqueous environment.
Floating weeds. These plants rest upon the water surface.
Their roots hang freely into the water or sometimes
attach to the bottom of shallow ponds or streams.
Emergent weeds. These typical plants of natural
marshlands are often found along the shorelines of
ponds and canals. They stand erect and are always
rooted into very moist soil. .
Submerged weeds. Although a few floating stems or
leaves may exist on the water surface, these plants
grow completely under water.
Some weeds and invasive plantS occur mainly in riparian
habitats, along rivers, streams, or other watercourses. These
terrestrial plants, such as Japanese knotweed (PolygQJ\um
cuspidatum), Himalaya blackberry (Rubus armenicus),
reed canarygrass (Phalaris arundinacea), and saltcedar
(Tamarix spp.), require the frequent disturbance or high
water table associated with rivers, streams, lakes, or ponds.
These plants can alter the hydrology of an area and also
reduce human access to areas where they occur.

Physiological Classification
Plants differ in their responses to temperature, light, day
Ecology of Weeds and Invasive Plants 211

length, and other factors of the environment. These


differences in plant physiology and biochemistry have also
been used as a basis for weed classification.
Most plants, called C3 plants, use the Calvin-Benson
cycle exclusively as a method of fixing carbon dioxide,
water, and light energy into sugars. This terminology is
used because the first stable product of photosynthesis in
such plants (phosphoglyceric acid) has three carbon atoms.
In some plants, called C 4 plants, the first stable
photosynthetic products are four-carbon atom sugars, such
as oxaloacetate, malate, and aspartate.
This physiological distinction may not seem significant
as a means of categorising weeds. However, these
differences in photosynthetic pathway result in substantial
biochemical, anatomical, and morphological variation
among species. Because of these differences, C4 weeds are
often more efficient at photosynthesis and can be more
competitive than C3 weeds and crops, especially in hot, dry
climates. Of the 18 worst weeds in the world noted by Holm
et al., 14 have the C4 pathway of carbon fixation.

Classification by Day Length


Classification by day length is based on a photoperiodic
response of flower initiation in plants. Three distinct classes
of day length response are known: sho~t day, long day, and
day neutral. Although these responses are named for the
length of the light period, it is now known that plants detect
and respond to the length of the dark period (e.g., short-
day plants are actually long-night plants).
Weeds that have a short-day response to day length,
such a-s lambsquarters (Chenopodium album) and
cocklebur (Xanthium spp.), are stimulated to flower when
days are short and maintain vegetative growth when days
are long. Long-day weeds, like henbane (Hysocyamus'
niger) and dogfennel (Eupatorium capillifolium), maintain
212 Plant Ecology

vegetative growth when days are short but are induced to


flower under long-day conditions. Other weeds (e.g.,
nightshades) remain vegetative or flower irrespective of the
photoperiodic condition.

Undesirability Classification
The term noxious weed is a legal term that refers to any
plant species capable of becoming detrimental, destructive,
or difficult to control. Legally, a noxious weed is any plant
designated by a federal, state, or county government as
injurious to public health, agriculture, recreation, wildlife,
or property. Many states, provinces, and countries maintain
at least one official list of such weeds so tbat their
introduction can be prevented or restricted.
Noxious weeds usually create a particularly
undesirable condition in crops, forest plantations, grazed
rangeland, or pastures. For example, the presence of
noxious weed seed in seed crops can prevent the sale and
distribution of that crop across national and international
boundaries. Poisonous weeds, which can be landscape
ornamentals or occur in pastures and rangeland, represent
a special kind of undesirability, since they can be a direct
threat to human or animal health.

Ecological Classification
Weeds, and in particular invasive plants, are often
classified using ecological categories related to population
behavior. The flora includes many weeds, which may also
be colonisers (taxa appearing early in vegetation
succession) or naturalised species (exotic species that form
sustainable populations without direct human assistance).
By this classification scheme, invasive plants are a subset
of naturalised species that are spreading. Not all
naturalised taxa are invasive, however, nor are all
colonisers considered to be weeds.
Ecology of Weeds and Invasive Plants 213

Groves and Cousens and Mortimer divide the process


of invasion by an exotic species into the phases' of
introduction, colonisation, and naturalisation. These three
phases of invasion are defined as follows:
Introduction. As a result of dispersal, propagules arrive
at a site beyond their previous geographical range and
establish populations of adult plants.
Colonisation. The plants in the founding population
reproduce and increase in number to form a colony that
is self-perpetuating.
Naturalisation. The species establishes new self-
perpetuating populations, undergoes widespread
dispersal, and becomes incorporated into the resident
flora.
Richardson et al., however, argue that colonisation as used
by Cousens and Mortimer is a component of naturalisation,
and the term invasion should be distinguished from
naturalisation and used to describe widespread dispersal
and incorporation of an exotic species into the resident
flora.
Such differences of opinion on terminology pertaining
to invasion will likely diminish as further knowledge is
gained about the ecological processes involved. Weed
species can be organised according to evolutionary
strategies that are based on genetically determined patterns
of carbon resource allocation. One prevalent theory holds
that two fundamental external factors limit the amount of
plant material (vegetation) that can accumulate within an
area. These factors are stress and disturbance. When the
extremes of these factors are considered, the following
possible strategies of evolutionary development emerge:
Stress tolerators. These are plants that survive in
unproductive environments by reducing their biomass
allocation for vegetative growth and reproduction and
increasing their allocation to maintenance and defense.
214 Plant Ecology

They exhibit characteristic!> -hat ensure the endurance


of relatively mature individuals in
Competitors. These are plants that have evolved
characteristics that maximise the capt..lre of
environmental resources in productive but relatively
undisturbed conditions. These plants have extensive
vegetative growth and are abundant during the early
and intermediate stages of succession.
Ruderals. Ruderals are plants that are found in highly
disturbed but potentially productive environments.
These plants are usually herbs, characteristically
having a short life span, rapid growth, and high seed
production. They occupy the earliest stages of
succession.
Most herbaceous weed species fall into one of two
combined strategies, competitive ruderals or stress-tolerant
competitors. Plants possessing the competitive ruderal
strategy have rapid early growth rates and competition
between individual plants occurs before flowering. Such
plants occupy fertile sites and periodic disturb?l1ce (e.g.,
annual tillage) favors their abundance and distribution.
Many annual, biennial, and herbaceous perennial weed
species found on arable land fit the criteria for the
competitive ruderal tactic. Stress-tolerant competitors are
primarily trees or shrubs, although some perennial herbs
also fall into this category. Common characteristics of these
weeds are rapid dry-matter production, large stem
extension, and high leaf area production.
WEEDS AND INVASIVE PLANTS IN PRODUCTION SYSTEMS

Some weed species have even achieved worldwide


prominence. Most weeds are important, however, from a
more local perspective. The local distribution of weeds is
influenced by biotic and abiotic environmental factors that
determine habitat types and human activities. Abiotic
factors that affect weed occurrence are soil type, soil pH,
Ecology of Weeds and Invasive Plants 215

soil moisture, light quantity and quality, precipitation


pattern, and variation in air, soil, and water temperatures.
Disturbed areas also are higher in susceptibility to
invasion than habitats that exist for long periods of time in
late succession. Biological factors, such as the incidence of
insects and diseases on either weeds or associated crops,
grazing activities of animals, and plant competition, also
can influence the distribution of weeds. It is for all of these
reasons that human land uses, such as farming, forestry,
range management, and recreation, are major causes of
local and regional patterns of weed distribution. Plant
species react in different ways when their habitats are
disturbed by humans; some species flourish because of the
disturbance, whereas others migrate or die and are
replaced.

Weeds and Agricultural Land


Weeds must have been known to early farmers because
hoes and other grubbing" implements, artifacts of those
1/

ancient times, have been found at archeological sites. In


addition, many references account for the detrimental
effects of weeds on crop yields, from the early writings of
Theophrastus and the Bible to more recent books. Even
today, weeds are considered to be just an incidental part of
food production in most parts of the world, where farmers
are simply people with hoes. The use of modern mechanical
and chemical tools to control weeds is actually little more
than a century old, even though weeds have been
associated with humans since agriculture began.
Human action is the most important factor determining
the occurrence and distribution of agricultural weed
species. TMany agrestals that accompanied crops for
centuries in Europe have now become locally extinct,
retreating to their climatic optimum where most survive
outside cultivated fields. Other we~ds have increased in
both prominence and abundance as agricultural practices
216 Plant Ecology

change. Holzner and Immenon suggest several causes for


such changes in weed species composition:
Improved seed cleaning, which results in the local
eradication of "specialists" that are unable to grow
outside arable land and depend on being sown with the
crop
Abandonment of crops, which leads to loss of
specialised weeds
"Leveling" of environmental conditions, which results
in a uniform weed flora . Increased reliance on crop
monocultures, which tends to simplify the weed flora
Combine harvesting, which allows some weed species
to shed seed in the field and distributes the seed of
others
Reduced-tillage and "no-tillage" operations, which
promote perennial species
Reduced competitive ability of short-stature crops and
crops treated with chemical growth regulators
Extensive use of herbicides, which causes sensitive
species to become locally extinct or to evolve resistance
to the chemical

Weed Control
A goal of agriculture for the last half century or more has
been to develop efficient methods of weed control in crops,
forest plantations, rangelands, and noncrdp situations. The
search for cost-effective ways to control weeds has often
focused on tillage and herbicides as a means to reduce labor
requirements and production costs or increase yields.
Below are some reasons to control weeds in cropland.
The threat of weeds to crop productivity accounts for
most of the human effort devoted to weed control. It is
estimated that 10-15% of the total market value of farm
products in the United States is lost because of weeds. This
Ecology of Weeds and Invasive Plants 217

loss amounts to about $8 billion to $10 billion per year.


Direct losses to forests and rangeland are more difficult to
estimate than agricultural losses. Walstad and Kuch believe
that nearly 30% reduction in wood productivity could
result because of weed occupation during the early stages
of forest plantation formation. The U.S. Forest Service
estimates that about 3.5 million acres of National Forest
System lands are infested with invasive plants.
Weeds have a detrimental effect on crop quality as well
as quantity, especially crops that must meet size, color,
nutrient content, or contamination-free standards. For
example, yields of alfalfa hay in California are often highest
during the first cutting when annual weeds are present.
However, hay quality is also low when weeds are present
in the crop. For example, protein content can fall from over
20% to below 10% when the hay contains large amounts of
weeds. Such decreases in grade or quality often mean
lowered revenue for growers, since a premium price is
usually paid for commodities of high quality.
In some cropping systems, the crop seed and weed seed
are so similar in weight and shape that separation at harvest
is. difficult. Examples are alfalfa and dodder (Cuscuta spp.)
seed, soybean seed and nightshade fruits, and pea seed that
are mixed with the immature flowers of Canada thistle
(Cirsium arvensis). If the weed material is not removed
from these crops by screening, lower price for the
commodity will result. For seed crops, the presence of a few
noxious weed seed, even less than 1%, usually makes the
commodity unmarketable.
Weed control is a major reason for many cultural
practices associated with crop production. For example,
weeds are killed during plowing and cultivation (tillage) to
prepare seedbeds for planting. A report by the U.S.
National Research Council indicates that 92-97% of the
acreage planted to com, cotton, soybean, and citrus are
treated with herbicides each year. In addition 87% of all
218 Plant Ecology

citrus acreage and 75% of potato and vegetable crops


acreage in the United States are chemically treated for weed
control. According to the US Environmental Protection
Agency, 60% of the total pesticide sales in the United States
in 1999 was for herbicides. There is no doubt that weed
control is a costly endeavor in the production of most crops.
Weeds also interfere with harvesting operations, often
making harvest more expensive and less efficient. For
example, weeds sometimes get wrapped around rollers or
cylinders of mechanical harvesters, causing equipment
breakdowns and longer harvest times. Up to 50% loss in
efficiency and 20% loss of yield can result from weed
presence at harvest time.
Some weed species act as alternate hosts or harbor
insects, pathogens, nematodes, or rodents thdt are crop
pests. Numerous specific examples exist of various pest
organisms that benefit from the presence of weeds. For
example, aphids and cabbage root maggots live on wild
mustard, later attacking cabbage and other cole crops.
Nightshades are hosts of the Colorado potato beetle.
Disease organisms, such as maize dwarf mosaic and maize
chlorotic dwarf virus, use Johnsongrass (Sorghum
halepense) rhizomes to overwinter.
Black stem rust uses barberry (Berberis thunbergii),
quackgrass (Agropyron repens), and wild oat (Avena fatua)
as hosts prior to infesting cereal crops. Rodent damage to
orchards can be prevented by weeding around trees before
winter. It also is possible for weeds to aid in thE' prevalence
or spread of certain bene-fidal organisms that are used to
control other pests. In such cases, the weeds act as an
alternate source of food or cover for the beneficial
organisms, allowing them to survive when thE' preferred
host is not available. Improve A nimal Health.
Some weeds are poisonous to animals. However, plants
toxic to one species of animal may be harmless to others.
For example, larkspur (Delphinium spp.) will kill cattle if
Ecology of Weeds and Invasive Plants 219

eaten in sufficient quantity, but sheep and horses are


relatively unaffected by this rangeland weed. In contrast,
fiddleneck (Amsinckia spp.) is highly toxic to horses, while
other livestock are relatively tolerant of it. It is estimated
that up to 10% of range-grazing livestock may become
afflicted by poisonous plants at some time during each
growing season:
In addition to direct poisoning, animals may experience
other discomforts from association with certain weed
species. Some plants contain chemicals that make animals
abnormally sensitive to the sun, a phenomenon called
photosensitization. Other plants contain teratogenic
materials that result in fetal malformations. For example,
malformed lambs can result if false hellebore (Veratrum
californicum) is ingested by sheep around the fourteenth
day of gestation. Bracken fern (Pteridium aquilinum)
causes a disease of cattle called "red water" because of the
blood-colored urine that is its symptom. This weed causes
cancer of the bladder if eaten in sufficient quantities.
Weeds affect a number of human activities that are
difficul t to assess in monetary terms. The presence of weeds
can reduce real estate values because of the unkempt and
unsightly appearance of the property. Dense moisture-
holding weed growth aids the deterioration of wooden and
metal structures and machinery, further reducing property
value. In fire-prone ecosystems, weeds can provide fuel to
carry fire, further endangering structures and property.
Access and enjoyment of recreation areas are also reduced
by weed presence.
Some rivers and lakes in the tropics and subtropics are
clogged by aquatic weeds, making travel on them nearly
impossible. Ross and Lembi provide an interesting example
of how weeds influence transportation costs. They indicate
that in 1969 and 1970,487,000 tons of wild oat seed were
inadvertently transported from Canada to the United States
. along with 16 million tons of grain. The transportation costs
220 Plant Ecology

for the wild oat were estimated at $2 million, which did not
include the $2 million cost for cleaning the grain to remove
contamination.
Weeds are kept free from highway intersections to
prevent accidents. Airports and railways also keep signs
and lights free of weeds so that maximum visibility can be
maintained. Power line rights-of-way are kept free of tall
growing vegetation to prevent power outages if trees
contact power lines during storms and to increase access to
downed power lines.
Toxicants or irritants produced by weeds can cause
serious health problems for some people. These discomforts
or illnesses include hay fever, dermatitis, and direct
poisoning. Hay fever afflicts millions of people each year.
It is caused by an adVerse effect of proteins associated with
the pollen of certain plants on the respiratory system of
susceptible people. Ragweed is best known for causing hay
fever. However, pollen from many other broadleaved
plants, grasses, trees, and shrubs causes similar allergic
reaction~. Each year, many people are troubled by poison
ivy (Rhus radicans), poison oak (R. diversiloba), and poison
sumac (R. vernix). These plants produce and store a toxic
substance called urushiol that causes intense itching and
rash upon contact with the skin. Many plants contain toxic
substances that when ingested cause sickness or death to
humans. Toxic substances in weeds include alkaloids,
glycosides, oxalates, resins and resinoids, volatile oils, acrid
juices, phytotoxins (toxalbumens), and minerals. There are
few poisons, including synthetic substances and minerals,
that approach the strength and violence of illnesses caused
by some plant-produced toxins.

Weeds in Forests
There are many natural conditions such as climate, soil type
and fertility, topography, and events like hurricanes and
wildfire that shape forested landscapes. Following
Ecology of Weeds and Invasive Plants 221

"catastrophic" disturbances, it is common for forests to


undergo a sequence of vegetation changes that result in a
forest nearly identical to the one previously destroyed. This
process of natural forest reestablishment through
successive changes in vegetation composition is called
secondary succession.
Following a radical disturbance, like a fire or clearcut,
a new patch in the physical environment is once again
available for colonisation by plants. In such situations,
"pioneer" tree (e.g., poplar, birch, alder, and some conifers)
or shrub species are quick to colonise the disturbed areas
and can dominate them for years to decades. This rapid
recolonisation by usually native pioneer species, although
a normal stage in succession, can delay the revegetation of
disturbed sites with more economically desirable trees.
The major disturbance to forests of any region is the
harvesting of wood by humans. It was estimated in 1989
that each year the world loses 37 million acres of forest in
this manner and current estimates remain unchanged. In
temperate conifer forests logging, especially without any
follow up reforestation activities, led to the gradual
replacement of conifers by less desirable herbaceous, shrub
or hardwood species.
In Canada large-scale weed problems have occurred
due to exploitation forestry, which strives to maximise
profits and minimize costs. Weed problems were
exacerbated by poor choice of forested stands to harvest,
season and method of harvesting, intensity of utilisation,
and lack of attention to regeneration. Walstad and his
associates similarly indicated that hardwoods occupy 32%
of the prime timberland in western Oregon that was once
dominated by conifers.

Forest Regeneration

Most forests regenerate naturally following disturbance


given enough time. However, logging activities and land
222 Plant llcology

clearing are the principal disturbance factors that both set


up and modify the natural patterns and time frames of
succession so that native and exotic weed species are
favored and even dominate many/forest types. The ability
of a site to regenerate, as well as ihe composition of species
following such disturbances, is most dependent on the
type, frequency, and severity of the tree removal operation.
In the coastal Douglas-fir forests of the U.S. Pacific
Northwest, the impact of both native and exotic plants is
currently restricted to the earliest stages of forest succession
that follow logging and fires. Ruderal exotic forbs, such as
Canada thistle or woodland groundsel (Senecio jacobaea),
and some exotic shrubs, such as Scotch broom (Cytisus
scoparius), displace native early seral vegetation in some
locations and reduce tree regeneration in others.
Though exotic plants are typically eliminated from the
plant community after a few years to a decade of forest
stand development", exotic shade-tolerant species are
capable of persisting and/or invading forest understories
if relatively open stand conditions are maintained through
clearcutting or severe silvicultural thinning. In particular,
false-brome (Brachypodium sylvaticum) poses a serious
threat to forest understory communities in that region.
Several techniques, collectively known as artificial
regeneration, have been used successfully to replant many
logged-over areas in many countries. This method usually
involves collecting seed of preferred tree species,
germinating and growing the seedling trees in nurseries,
outplanting them to field sites, and following this by
intensive chemical weed control. Wagner et al. (2006),
surveying 60 studies, found that the most intensive
vegetation management treatments always improved crop
tree growth, although results varied by location, tree
species, and length of time from experiment initiation.
Despite these successes in projected crop tree biomass yield,
important questions still remain about the ecological
Ecology of Weeds and Invasive Plants 223

(Balandier et al. 2006), social, and economic desirability of


converting vast acreages of naturally regenerated forests
into tree farms.
Weeds in Rangelands
The destruction and replacement of vegetation by humans
are now common occurrences over most of the world, with
a loss in primary productivity and floristic diversity often
being the result. The invasion of exotic plants is both a cause
and a consequence of such environmental manipulation.
However, it is rare that invaders cause the replacement of
most or all of the plant and animal species in a disturbed
ecosystem.
A possible exception to this generalisation is rangeland
weeds. In this system of production, species replacement
following disturbance has been so complete that only a
sketchy picture of predisturbance conditions remain. We
offer the sagebrush (Artemisia tridentata)-cheatgrass
(Bromus tectorum) steppe as an example.
The chance introduction of cheatgrass before the turn
of the last century to the Great Basin of North America
altered the entire native shrub ecosystem of that region.
D' Antonio and Vitousek after Billings indicate that its
introduction provides a classical case of biological
impoverishment where the concomitant environmental
change allows successful replacement of indigenous
vegetation. In this case, native perennial bunchgrasses and
shrubs, particularly sagebrush, were first grazed by large
herbivores, then invaded by cheatgrass, and subsequently
subjected to range fires.
Original Vegetation and Early Land Use History of
Great Basin. Billings and others indicate that the western
Great Basin was not part of the bison range of the North
American Great Plains because the rhizomatous Col grasses
on which the bison thrived cannot grow on the summer-
dry steppes of this region.
224 Plant Ecology

Rather, perennial C3 bunchgrasses of the genera Poa,


Festuca, Agropyron, and Stipa dominated the grass stratum
of this sagebrush ecological formation. Apparently, the
native bunchgrasses of the region also did not carry fire
well because range fires in the sagebrush-bunchgrass
steppe, in contrast to the Great Plains, were rare. The native
ungulate herbivores were antelope, deer, desert bighorn
sheep, and elk which, because of their smaller size and
numbers than bison, created a relatively light impact on the
sagebrush-grass community.
WEED AND INVASIVE PLANTS IN LESS MANAGED HABITATS

Certain forests, deserts, prairies, beaches, marshes,


estuaries, and riparian areas have been protected from
disturbance or designated as wilderness throughout the
world. Wilderness and similarly managed natural areas,
such as national parks and monuments, provide many
benefits to society. These benefits include the preservation
of biodiversity, unique natural features, and watersheds as
well as opportunities for recreation and personal
fulfillment.
Although land management agencies place a high
priority on protection of natural ecosystems and wilderness
areas, some of these benefits are threatened by increasing
levels of human activity within and outside areas
designated for protection. The introduction of exotic species
into such areas is of particular concern due to the potential
for irreversible impacts on the natural ecosystems that such
areas represent. Three research areas were identified by the
Aldo Leopold Wilderness Research Institute to address the
question of exotic plant invasion into wilderness:
Understanding the introduction, spread, and
distribution of exotic species within wilderness
Understanding the effects of exotic species on
wilderness values
Ecology of Weeds and Invasive Plants 22S

- Identifying and evaluating management options and


their consequences
Parks and her associates) examined the patterns of invasive
plant diversity in mountainous ecoregions of the
northwestern United States. Their analysis found that
altered riparian systems an~ disturbed forests were
especially vulnerable to exotic plant invasion.
Conversely, alpine areas, forests, and grasslands
designated as wilderness were still relatively unaffected by
invasive plants, with introductions often being restricted to
campsites, roads, or trails. The predominance of wilderness
throughout much of the western United States is believed
to contribute to the lower incidence of invasive plants in
mountainous ecoregions of that area compared to other
regions. Human settlement and intense land use at low
elevations were identified as factors that enhance invasive
plant introductions.
There are many examples of the widespread regional or
even global distribution of weeds. One of the earliest
examples is that of Hitchcock and Clothier which describes
the distribution of native and introduced weeds in Kansas
as that land was being developed for agriculture. A similar
study was accomplished by Mason, who described the
occurrence of wild oat throughout several provinces of
central Canada.
These studies are augmented by more recent
descriptions of widespread infestations of weed species, for
example, leafy spurge (Euphorbia esula), purple loosestrife
(Lythrum salicaria), downy brome (also known as
cheatgrass) (B. tectorum), Paterson's curse (Echium
platagineum), and lantana (Lantana camara). The ability to
disperse widely is a common characteristic of many weed
and invasive plant species, which has been exacerbated in
recent decades by increasingly global movement of humans
and goods.
226 Plant Ecology

Any harmful organism that is spreading or has the


capacity to spread poses a threat to uninfested areas
without regard for ownership boundaries. Thus, a
spreading species represents a problem to more people than
just those whose land it currently occupies. Such situations
make a strong case for legislation (weed laws), quarantine
districts, or other governmental interventions to reduce or
slow the spread of weeds and invasive plants. Furthermore,
governmental objectives for weed suppression may be less
constrained by cash flow than those of individual farmers,
rant.pers, or forest land owners.
REFERENCES

Baskin, Yvonne. A Plague of Rats and Rubbervines: The Growing Threat Of


Species Invasions. Island Press, 2003.
Burdick, Alan. Out of Eden: An Odyssey of Ecological Invasion. Farrar Straus
and Giroux, 2005.
Coates, Peter. American Perceptions of Immigrant and Invasive Species:
Strangers on the Land. University of California Press, 2007.
Elton, Charles S. The Ecology ofInvasions by Animals and Plants. University
of Chicago Press, -2000.
Janick, Jules. Horticultural Science. San Francisco: W.H. Freeman, 1979.
Lockwood, Juli~; Martha Hoopes, Michael Marchetti Invasion Ecology.
Blackwell Publishing, 2006.
McNeeley, Jeffrey A. The Great Reshuffling: Human Dimensions Of Invasive
Alien Species. World Conservation Union (IUCN), 200l.
Van Driesche, Jason; Roy Van Driesche Nature Out of Place: Biological
Invasions In The Global Age. Island Press, 2004.
Terrill, Ceiridwen. Unnatural Landscapes: Tracking Invasive Species.
University of Arizona Press, 2007.
10
Phage Ecology and Plants

Bacteriophages (phages), potentially the most numerous


"organisms" on Earth, are the viruses of bacteria, more
generally, of prokaryotes. Phage ecology is the study of the
interaction of bacteriophages with their environments.
Phage ecology is increasingly an important component of
sessions and symposiums associated with phage meetings
as well as general microbiological meetings.
Phages are obligate intracellular parasites meaning that
they are able to reproduce only while infecting bacteria.
Phages therefore are found only within environments that
contain bacteria. Most environments contain bacteria,
including our own bodies. Often these bacteria are found
in large numbers.
As a rule of thumb, many phage biologists expect that
phage population densities will exceed bacterial densities
by a ratio of 10-to-1 or more. As there exist estimates of
bacterial numbers on Earth of approximately 1030, there
consequently is an expectation that 1031 or more individual
virus particles exist, making phages the most numerous
category of "organisms" on our planet.
Bacteria appear to be highly diverse and there possibly
are millions of species. Phage-~cological interactions
therefore are quantitatively vast: huge numbers of
interactions. Phage-ecological interactions are also
qualitatively divers\.!: There are huge numbers of
2.28 Plant Ecology

environment types, bacterial-host types, and also


individual phage types.
The phage consists of a nucleic acid core that is made
up, depending on the phage, of DNA or, less often, of RNA.
Surrounding this nucleic-acid genome is '.a proteinbased
capsid. The capsid plays three important roles in the phage
life cycle: (i) protecting the phage genome during the
extracellular search; (ii) effecting phage adsorption, which
is the attachment of the virion particle to a susceptible
bacterium; and (iii) the subsequent delivery of the phage
genome into the cytoplasm of the now-infected bacterium.
The extracellular search occurs via phage diffusion
through an aqueous milieu. During this period the phage
must avoid physical damage while waiting to encounter a
susceptible bacterium. The likelihood that an individual
phage will find a bacterium to adsorb is a function of time,
the phage diffusion rate, and the local density of phage-
susceptible bacteria, with more bacteria resulting in faster
phage adsorption. A slightly different set of parameters
governs the likelihood of phage attack on bacteria, with
adsorption a function of time and the phage diffusion rate,
but also of phage density, with more phage resulting in
more bacteria infected.

Figure 1. Structure of Bacteriophage


Phage Ecology and Plants 229

Phage adsorption to bacteria furthermore is a functIon


of phagebacteria chemical and physical interaction. Phage
display proteins with high affinity to specific bacterial
surface molecules-an association analogous to antigen
recognition by immune systems. The host range of most
phages, Le., the species that they are capable of
productively infecting, consequently is relatively narrow-
typically limited to only a single bacterial genus, species, or,
often, even to onsly a limited number of strains within a
given species. Thus, while total phage densities can be
enormous-as many as 100 million or more per gram of soil
or m1 of aquatic environment the actual density of phages
capable of infecting a particular bacterial strain usually is
much smaller.
Following uptake, the phage genome can rapidly
subvert host-cell functions, directing bacterial metabolism
during the phage latent period towards phage production.
'Oepending on the phage, these virions either may
,accumulate within the bacterial cytoplasm or, for
filamentous phage, virions instead are released-over the
course of an extended latent period- across an otherwise
intact cell envelope.
PHAGE ECOLOGY

Phage ecology is the study of the interaction of phage with


their biotic and abiotic environments. The scale of phage
ecology is at once both exhilarating and intimidating. As a
guiding principle toward understanding phage ecology we
therefore seek generalizations, plus look to more
established scientific discipiines for guidance, the most
obvious being general ecology. Toward that end we can
speak of phage "organismal" ecology, population ecology,
community ecology, and ecosystem ecology. Phage ecology
from these perspectives will be described in turn.
Phage ecology also may be considered (though mostly
less well formally explored) from perspectives of phage
230 Plant Ecology

behavioral ecology, evolutionary ecology, functional


ecology, landscape ecology, mathematical e~ology,
molecular ecology, physiological ecology and spatial
ecology.
Phage ecology additionally draws from microbiology,
particularly in terms of environmental microbiology, but
also from an enormous catalog of study of phage and
phage-bacterial interactions in terms of their physiology
and, especially, their molecular biology.
Following the traditions of ecology we can differentiate
phage ecology into four general categories:
Phage organismal ecology is the study of the
adaptations that phage employ to increases their
likelihood of transmission between hosts such as virion
desiccation resistance, ability during infection to repair
ultraviolet (UV) light-mediated nucleic-acid damage,
and so on.
Phage population ecology is the study of phage life-
history characteristics, particularly as they apply to
phage growth and intraspecific (between-phage)
competition. U~derstanding phage population growth
within spatially structured environments, such as
within the phyllosphere or rhizosphere, is particularly
challenging.
Phage community ecology focuses on the stability of
phage-containing environments such as the propensity
of phage to drive phage-sensitive bacteria to extinction.
Phage community ecology is complicated by the
continuous co-evolution of bacteria with their phage
predators. In addition, phage play important roles in
the horizontal transfer of DNA between bacteria.
Phage ecosystem ecology considers the phage impact
on energy flow and nutrient cycling within ecosystems.
Phage, for example, can disrupt the soil bacteria
responsible for nitrogen cycling.
Phage Ecology and Plants 231

Phage Organismal Ecology


Phage "organismal" ecology is primarily the study of the
evolutionary ecological impact of phage growth
parameters:
latent period, plus
eclipse period
rise period
burst size, plus
rate of intracellular phage-progeny maturation
adsorption constant, plus
rates of virion diffusion
virion decay (inactivation) rates
host range, plus
resistance to restriction
resistance to abortive infection
various temperate-phage properties, including
rates of reduction to lysogeny
rates of lysogen induction
the tendency of at least some phage to enter into (and
then subsequently leave) a not very well understood
state known (inconsistently) as pseudolysogeny.
Another way of envisioning phage "organismal" ecology is
that it is the study of phage adaptations that contribute to
phage survival and transmission to new hosts or
environments. Phage organismal" ecology is the most
II

closely aligned of phage ecology disciplines~ith the


classical molecular and molecular genetic analyses of
bacteriophage.
From the perspective of ecological subdisciplines, we
can also consider phage behavioral ecology, functional
ecology, and physiological ecology under the heading of
phage "organismal" ecology. However, as notEi!d, these
232 Plant Ecology

subdisciplines are not as well developed as more general


considerations of phage "organismal" ecology. Phage
growth parameters often evolve over the course of phage
experimental adaptation studies.
In the mid 191Os, when phage were first discovered, the
concept of phage was very much a whole-culture
phenomenon, where various types of bacterial cultures (on
solid media, in broth) were visibly cleared by phage action.
Though from the start there was some sense, especially by
Felix d'HereIle, that phage consisted of individual
"organisms", in fact it wasn't until the late 1930s through
the 1940s that phage were studied, with rigor, as
individuals, e.g., by ~lectron microscopy and single-step
growth experiments: Note, though, that for practical
reasons much of "organismal" phage study is of their
properties in bulk culture (many phage) rather than the
properties of individual phage virions or individual
infections.
This, somewhat whole-organismal view of phage
biology saw its heyday during the 1940s and 1950s, before
giving way to much more biochemical, molecular genetic,
and molecular biological analyses of phage, as seen during
the 1960s and onward. This shift, paralleled in much of the
rest of microbiology, represented a retreat from a much
more ecological view of phages. However, the organismal
view of phage biology lives on as a foundation of phage
ecological understanding. Indeed, it represents a key thread
that ties together the ecological thinking on phage ecology
with the more "modern" considerations of phage as
molecular model systems.
The basic experimental toolkit of phage "organismal"
ecology consists of the single-step growth (or one-step
growth; example) experiment and the phage adsorption
curve (example). Single-step growth is a means of
determining the phage latent period (example), which is
approximately equivalent (depending on how it is defined)
Phage Ecology and Plants 233

to the phage period of infection. Single-step growth


experiments also are employed to determine a phage's
burst size, which is the number of phage (on average) that
are produced per phage-infected bacterium.
The adsorption curve is obtained by measuring the rate
at which phage virion particles attach to bacteria. This is
usually done by separating free phage from phage-infected
bacteria in some manner so that either the loss of not
currently infecting (free) phage or the gain of infected
bacteria may be measured over time.

Phage Population Ecology


A population is a group of individuals which either do or
can interbreed or, if incapable of interbreeding, then are
recently derived from a single individual. Population
ecology considers characteristics that are apparent in
populations of individuals but either are not apparent or
are much less apparent among individuals. These
characteristics include so-called intraspecific interactions,
that "is betwee~ individuals making up the same
population, and cah include competition as well as
cooperation.
Competition can be eith~r in terms of rates of
population growth or in terms of retention of population
sizes. Respectively, these are population-density
independent and· dependent effects. Phage population
ecology considers issues of rates of phage population
growth, but also phage-phage interactions as can occur
. when two or more phage adsorb an individual bacterium.

Phage Community Ecology


A conuriimity consists of all of the biol9gical individuals
found within a given enviroIiment, particularly when more
than one species is present. Community ecology studies
those characteristics of "Communities that either are not
234 Plant Ecology

apparent or which are much less apparent if a community


consists of only a single population.
Community ecology thus deals with interspecific
interactions. Interspedfic interactions, like intraspecific
interactions, can range from cooperative to competitive but
also to quite antagonistic. An important consequence of
these interactions is coevolution.
The interaction of phage with bacteria is the primary
concern of phage community ecologists. Phage, however,
are capable of interacting with species other than bacteria,
e.g., such as phage-encoded exotoxin interaction with
animals. Phage therapy is an example of applied phage
community ecology.

Phage Ecosystem Ecology


An ecosystem consists of both the biotic and abiotic
components of an environment. Abiotic entities are not
alive and so an ecosystem essentially is a community
combined with the non-living environment within which
that ecosystem exists. Ecosystem ecology naturally differs
from community ecology in terms of the impact of the
community on these abiotic entities, and vice versa. In
practice, the portion of the abiotic environment of most
concern to ecosystem ecolOgists is inorganic nutrients and
energy.
Phage impact the movement of nutrients and energy
within ecosystems primarily by lysing bacteria. Phage can
also impact abiotic factors via the encoding of exotoxins.
Phage ecosystem ecologists are primarily concerned with
the phage impact on the global carbon cycle, especially
within the context of a phenomenon know as the microbial
loop.
PHAGE TEMPERANCE AND VIRULENCE

Population and organismal ecologies are concerned with·


the adaptations that organisms employ to enhance their·
Phage Ecology and Plants 235

Darwinian fitness over the course of their life cycles. For


phage, the life-cycle steps most under their control are the
durability of the yirion particle, the breadth of the host
range, and the details of the infection strategy. In this
section we take a population-ecology approach to
contrasting two infection strategies: temperance versus
virulence. Subsequently we consider the impact of plants
on lysogeny.
For lytic phage, progeny release can occur only
following the total destruction (lysis) of the host bacterial
cell. One can differentiate lytic phage into two types,
temperate phage and obligately lytic (virulent) phage. Only
temperate phage can display lysogeny, an infection that
stalls shortly after the introduction of the phage genome
into the host cell, which then (in most cases) integrates as
a prophage into the bacterial chromosome. During
lysogeny phages neither produce virions nor lyse bacteria.
A temperate phage does not obligately enter into a
lysogenic relationship with its host bacterium; in fact many
temperate phage infections result in the immediate
production of phage progeny, i.e., a lytic cycle rather than
lysogeny. This decision is determined by characteristics of
the infecting phage and the metabolic state of the host.
When not induced, a phage in the lysogenic state replicates
as a giant gene complex along with the host cell's genome.
Lysogeny typically results in bacterial resistance to
infection by similar phage. Mutant phage that are able to
bypass this resistance are described as vir mutants.
One advantage that can be associated with such
virulence is an ability to actively infect homologous
lysogens (bacteria lysogenised by the same phage), though
establishment of this Vir phenotype can require multiple
phage mutations. Note, though, that a diversity of virulent
phage exist that are not vir mutants but instead are
unrelated to temperate phage. In addition to providing a
safe home to the temperate-phage genome, and blocking
236 Plant Ecology

the replication of non-virulent homologous phage,


lysogeny has the potential to alter the phenotype of the host
cell, a process known as phage (or lysogenic) conversion.
PHAGE LYSOGENY

What advantages are bestowed upon a phage that displays


a temperate lifestyle? We can describe obligately lytic
phage as essentially semelparous, with the acquisition of a
bacterial cell resulting in only a single reproductive
episode. As such these phage are perhaps best understood
as adapted to a so-called r strategy of population growth,
with a life-history emphasis on rapid population increase
when resources are plentiful.
Aiding in this strategy is their (i) avoidance of
physiological tradeoffs required for the display of lysogeny,
(ii) a commitment of all progeny to lytic growth (rather than
some fraction to lysogen formation), or (iii) an ability to
infect lysogens. These advantages, however, come with
requirements for long-term survival as free phage when
bacteria are less numerous, and/or dissemination to new
environments to find new bacterial hosts. Temperate phage
can also display the semelparous infection strategy of
virulent phage. A fraction of infections, however, will
instead result in lysogeny.
Lysogeny represents an iteroparity of sorts, i.e., more
than one reproductive episode per lifetime, at least so long
as one is willing to accept a clonally related population of
lysogtt?S as a single individual and a sporadic induction of
lytic cycles within this population as consecutive
repro.ducti'l:~ events.
Such a life-history approach is the more K-like strategy
whereby temperate prophage, by less-rapidly killing off
their bacterial resource, may more readily sustain their
population size at an environment's carrying capacity.
Filamentous phage similarly display a more semelparous
life-history strategy than virulent phage, with (i) a longer-
Phage Ecology and Plants 237

term maintenance of the host infection, (ii) only a few phage


progeny released over a given interval, and (iii) with
multiple intervals over which these phage progeny are
released.
Though lysogeny often is framed as an adaptation to
survival within relatively unstable environments, i.e.,
during "hard times", one could similarly argue that
lysogeny bestqws-a competitive advantage on phage that
is useful when bacterial populations are relatively stable,
i.e., not fluctuating in size. Lysogeny thusly can be an
effective K strategy so long as lysogens are not being
actively killed by lytic phage.
Exposure to . lytic phage may be less likely in well-
structured environments that limit phage diffusion and in
which lysogen--nrlcrocolonies exist as relatively few cells,
with low lysogen number minimising the potential for
temperate-phage mutation to anti-lysogen virulence. Of
course, arguing that wellstructured environments can favor
lysogeny over virulence is nearly a restatement of the "hard
times" hypotheSIs .whereby the "long periods of dearth"
required for this dominance is posited to be a consequence
of environmental structure rather than solely of temporal
variation in resource availability.

Lysogeny and Plants


Plants can induce pacteriallysogens (that is, cause them to
initiate their lytic cycle), a strategy that plants could employ
towards the elimination of bacterial pathogens. An extract
of inulberry leaves could induce lysogens of Pseudomonas
syringae. It may be· that these induced prophage are
"abandoning ship" in response to the plant's release of
antibacterial compounds, with the phage simply following
an evolutionary algorithm based on the logic that it is better
to take one's chances as a free phage than to continue to
infect a dying bacterium.
238 Plant Ecology

Also consistent with a shift away from lysogeny in the


lysogeny-lytic cycle balance, Menzel et al have reported a
negative impact of certain plant-growth regulators and
herbicides on the initial establishment of lysogeny. For
many bacteria, plant association marks not a high
likelihood of plant-induced bacterium death but, instead~ a
period of effective bacterial growth.
Given heterogeneous bacterial populations it could be
advantageous for temperate phage to lyse host bacteria
when times are good since not only are healthy, uninfected
bacteria potentially present, but those bacteria also could
contain nonhomologous prophage capable of infecting and
then lysing their own uninduced lysogens. That is, more
effective bacterial growth could tip an environment from
conditions that disfavor phage lytic growth and favor
lysogen survival (Le., environments in which it is better,
from the phage perspective, to be a bacterium) to conditions
that favor lytic growth and thereby disfavor a continued
display of lysogeny.
PHAGE RHIZOSPHERE

From the phage perspective much of what is of interest in


understanding the phage-plant interaction has to do with
the extracellular search, a province of phage organismal
ecology. How exactly do phage manage to find new
bacteria to infect before succumbing to the ravages of
environmentally induced virion decay? Answering this
question within the rhizosphere-the region consisting of
plant roots and surrounding soil-involves examining the
impact of soil structure and chemistry on the mobility and
survival of free phage and their hosts.
In most crop environments, with a few notable
exceptions, most of the time the soil is only partially
hydrated. The lack of a continuous aqueous phase greatly
complicates predictions regarding the diffusion of free
phage, as for any soil colloid. This situation is complicated
Phage Eco~ogy and Plants 239

ftu$er by the propensity of free phage to become trapped


within biofilms or reversibly adsorbed to particles, such as
clays, that are commonly found in the soil. This non-specific
and often reversible phage adsorbtion within soils is a
function of sorbent and virion surface chemistry, virion
size, and pH. Moreover, acidic soils can permanently
inactivate free phage.
We can only speculate whether these phage-soil
interactions help or hinder the phage in their search for a
suitable host. Since sorption to solid substrates lowers the
diffusion rates of free -phage, it localises them to specific
spatial regions, and thereby limits the maximum rate at
which they can encounter a suitable host cell. On the other
hand, adsorption to clays has been suggested to ,have a
protective effec~- by holding phage within 'a hydrated
environment. Both states are essentially in opposition only
within undisturbed soil. '
If the soil itself is disseminated, becomes fully hydrated,
or is otherwise well mixed, then associated phage, whether
free or bound, may be disseminated. Soil-adsorbed phage
thus could serve as a viable infectivity pool that is tapped
only as bacteria grow, diffuse, or swim into the phage
vicinity, or if the soil particle i:self is transferred into or onto
a bacterium-containing environment.
Phage can attack bacteria directly associated with plant
roots. Given the close proximity of roots to soil it seems
obvious that phage attack must occur via phage diffusion
from either surrounding soil or from neighboring roots .
. Rhizosphere bacteria, however, may gain an upper hand in
what likely is a constant battle between the phage predator
and bacteria prey by some combination of:
(i) relatively' low viable counts of phage capable of
infecting specific bacteria;
(ii) relatively low rates of phage diffusion within soil,
particularly under drier conditions or following phage
adsorption to ~oil;
Plant Ecology

(iii) relatively high rates of free-phage inactivation within


soil; and
(iv) physical (or spatial) refuges that provide bacteria with
a degree of physical protection from phage.
Indeed, one can envisage phage replication as equivalent
to a nuclear chain reaction with anything damping the
mobility or production of the phage "neutron" serving to
limit depletion of the bacterial "fuel": Significant bacterial
depletion occurs only once bacteria first have achieved a
critical "mass".
Phage researchers have significant experience handling
phage within environments containing relatively low
phage densities-and in which diffusion (and mixing) is
limited-since these are the conditions under which phage
growth typically occurs in solid media. Solid-media growth
involves mixing a small number of phage with a large
excess of bacteria, which is then poured into a thin layer
over a regular plate of agar growth medium. The soft-agar
-ih the top layer impedes phage an~ bacterial diffusion
while the bottom agar layer maintains a relatively constant
chemical and physical state for bacterial and phage growth.
Within the soft agar layer, phage populations grow as
plaques, which are expanding regions of phageinduced
bacterial lysis, each originating from a single phage
infection. Plaques appear transparent or translucent against
the background of the typically more-opaque bacterial lawn
growing within the agar-based substrate.
Of particular relevance to understanding host-phage
ecology within the rhizosphere therefore is determination
of the degree to which phage growth in soil systems
approximates the better understood solid-phase phage
growth in laboratory media. Though within agar plaque-
development theory as well as techniques for plaque-
growth quantification have been fairly well developed, to
our knowledge similarly finescale investigation has not
been attempted within a soil-based medium.
Phage Ecology and Plants 241

Individual soils likely vary' spatially and temporally


with regardto plaque-like growth-particularly as a function
of soil composition, degree of hydration, density and
physiological state of host bacteria, and rates of free-phage
inactivation. Nevertheless, we predict that the basic
principle of phage solid-phase population growth, i.e., a
phagediffusion mediated expanding sphere of bacterial
infection, could still apply.
Thus, without active mixing of soil, e.g., via the action
of invertebrates or other localized soil disruptions, we
speculate that bacterial microcolonies within the
rhizosphere may display periods of boom or bust with
regard to phage attack, with increasing microcolony size,
or mere time, increasing the likelihood of phage-
microcolony encounter. Infection of one bacterium within
a localised bacterial clone could result in the destruction of
part or all of a genetically homogeneous bacterial,
microcolony.
Means by which such coordinated attack may be
thwarted could include:
(i) variation in the physiological or anatomical state ot the
bacteria making up a clone, so that not all bacteria are
equally susceptible to phage attack (including the
formation of spores for those species that are able to,
as well as hyphael aging for streptomycetes;
(ii) display of motility such that bacteria progeny minimise
co-location and thereby avoid serial infection (though
with the caveat that active movement through soil
might increase the likelihood that individual bacteria
encounter phage); and
(iii) sequestration away from the soil such as within root
nodules colonised by rhizobia or perhaps following
bacterial penetration into plants upon infection.
Rarely encountering microcolonies could be antithetical to
the prosperity or even survival of obljgately lytic phage, but
242 Plant Ecology

could provide numerous stable, otherwise phage-free


niches for temperate-phage survival as bacteriallysogens.
PHAGE PHYLLOSPHERE

How phage interact with their bacterial hosts in the


phyllosphere, the aerial plant structures, is even less
understood than the phage ecology of the rhizosphere. The
phyllosphere presents a less hospitable environment-
relative to the rhizosphere-given the exposure to UV,
intense visible light, and desiccation that is likely on many
above-soil plant surfaces.
In the case of phages of the plant-pathogen Erwinia
amylovora, some studies have noted that phage are less
readily isolated from the aerial portions of trees, even
during times of active infection by the host. By contrast,
phage could almost always be isolated from the soil around
infected trees. That would suggest that the phage reservoir
is located in the soil, possibly with phage multiplying on
stray bacteria that fall from the tree to the ground below.
Other phage studies on the same host bacterium, however,
have found abundant E. amylovora phages in the
phyllosphere of infected trees.
If the virion reservoir is the soil for at least some phage
that attack plant epiphytes, then how do phage reach the
phyllosphere of a tree, which may have foliage 3 or more
feet off the ground? One possible explanation is that phage
invade the phyllosphere upon plant germination and then
remain a part of plant normal flora. Alternatively, phages
could move from plant to plant within the phyllosphere
with soil remaining a phyllosphere phage sink rather than
source habitat.
The habitats of otherwise plant-associated phage can
range beyond plants or plant-associated soil. Irrigation
waters and agricultural drainage are known to contain
phage capable of inf~cting plant-associated bacteria.
Erwinia phage have been isolated from lakes as well as
Phage Ecology and Plants 243

from sewage. The latter phage perhaps display broad host


ranges, attacking enteric bacteria associated with humans
as well as the bacteria associated with plants. Perhaps
similarly, phage infecting phytopathogenic Pseudomonas
have been isolated from sewage while Agrobacterium-
infecting phage have been isolated from feces.
Erwinia phage have also been isolated from a com flea
beetle. This association between phage and insect is of
interest since arthropods are known vectors for viruses that
directly infect plants. However, the degree to which
arthropods are responsible for phage transmission remains
an open question.
Indeed, phage movement within and between plants as
well as between soil and plants presumably follows paths
similar to those employed by bacteria, i.e., carriage by
animals, dust, soil, seeds, and water, including splashing
caused by hard rain, plus various human activities such as
pruning. A case can be made that rain, by promoting
epiphytic bacterial growth, can simultaneously supply
phage with (i) healthy hosts, (ii) water in which to diffuse
between bacteria, and (iii) a means of disseminating about
individual plants .as well as among populations of pla:nts.
PHAGE AND AGRICULTURE

Though plants are surrounded by phage, the vast bulk of


the phage impact on plants is mediated through plant-
associated bacteria. Plant-associated symbiotic bacteria can
range from helpful (mutuals) to harmful (pathogens), and
the phage impact on bacteria also can range from
mutualistic to parasitic.
Despite these complications, the phage impact, either
negative or positive on plants, tends to be limited (i) to
phage-induced bacterial lysis, (ii) to selection for phage-
resistance within bacterial communities, or (iii) to phage-
associated modification of bacterial phenotypes (phage
conversion). These we consider in order:
244 Plant Ecology

There have been a number of reports of phage


presencE!, in relatively small-scale experiments, that
result in reduced plant growth or reduction in plant
nitrogen content. 'Experimentally induced decline in
the plant-protective bacterium, Pseudomonas
fluorescens, has also been noted. However, it is
uncertain how often wild phages negatively impact on
plant growth under naturally occurring conditions. We
speculate that extrapolation of observations from small
to large scales is challenging due to ignorance of
naturally occurring phage-bacteria dynamics and
spatial heterogeneities that exist across large plots.
Indeed, we are aware of only one report suggesting that
phage may have obstructed plant growth on a large
scale in a non-experimental agricultural setting, as
mediated by reductions in soil rhizobia. Further
exploration of this "extrapolation" issue could be
difficult assuming reluctance to conduct large-scale
field tests of phage that are antagonistic to beneficial
bacteria.
It is typically assumed that reduced plant growth
correlated with phage presence is a consequence of
. phage antagonism against beneficial bacteria, e.g.,
phage-induced lysis. Lytic phage can also indirectly
reduce the fitness of susceptible bacteria. This fitness
reduction can be manifest either as a density decline of
bacteria inhabiting specific niches, or by a decline only
of susceptible bacteria, the latter suggesting a
replacement of susceptible bacteria by similar but
phage-resistant bacteria. The above-cited Demolon and
Dunez study, for example, has been much discussed in
the plantphage literature, ,with some authors
concluding that the observed negative phage impact
resulted from selection for a phage-resistant bacterial
phenotype that was less effective at nitrogen fixation.
This interpretation is consistent with more modern
Phage Ecology and Plants 245

phage community-ecology theory, which posits that


bacterial resistance to phage attack often comes at some
metabolic cost to bacteria. Consequently, a bacterium.
that does not display phage resistance may bEt ablEtio
invade and even drive to extinction a phage-resistant
population of otherwise-identical bacteria, at least so
long as phage are not present.
From a plant's perspective, this change in bacterial
prevalence upon phage attack is irrelevant unless phage-
sensitive and phageresistant bacteria display differences in
their ability to interact with plants. Such differences are
often observed, though it is important to note that for many
studies only a fraction of bacterial- mutations to phage
resistance result in significant change in plant-interaction
phenotypes.
Phage T4- and phage fEC2-resistant, mostly
lypopolysaccharide (LPS)-defective mutants of the soft-rot
Erwinias E. carotovora and E. chrysanthemi, for example,
generally do not display a loss of virulence. Given a relative
rarity of avirulence in phage-resistant mutants, an invasion
of phage into a pathogen population and subsequent
selection for phage-resistant variants may not impact
negatively on the virulence of the surviving bacteria.
Rather, the community impact may be seen as a decHpe of
the pathogen population followed by recovery by similarly
virulent but phageresistant bacteria.
Phage-resistant mutants of Ralstonia solanacearum,
which display various defects in LPS synthesis, were
predominantly avirulent in tobacco seedlings, as were
approximately 50% of selected phage-resistant
Xanthomonas campestris mutants. Mutational phage
resistance resulting from a loss of pili also has been shown
to reduce Rhizobium nodulation in clover. Mutation to
phage-resistance by P. fluorescens similarly can result in
reduced protection of radishes from Fusarium. wilt. The
mechanism by which this loss occurs, however, appears to
246 Plant Ecology

be not so much from a decline in P. fluorescens colonising


ability as due to insufficient induction of cross-reactive
systemic resistance by plants.
Like phage resistance, modification of bacterial
phenotype can result from phage conversion, Le., the
expression of prophage genes over the course of
lyso-genic infection. Little is known of phage
conversion positively affecting plant growth. There
exists some evidence for the converse, however:
lysogeny negatively impacting soybean-
Bradyrhizobium japonicum interaction. An example of
phage conversion that could possibly be interpreted as
positively affecting plant fitness, by poisoning a plant
predator, occurs in the case of toxic annual ryegrass,
Lolium rigidum. The developing seeds of this grass are
susceptible to infection and gall formation by a
nematode, Anguina funesta. If these nematodes are
carrying certain strains of the bacterium Clavibacter
toxicus, then the galls will contain corynetoxin. Related
to tunicamycin-like antibiotics, corynetoxin inhibits
protein glycosylation and can be fatal to graz~ng
animals (Le., sheep) that consume the infected grass.
Some phages are associated with the ability of C. toxicus
to produce corynetoxin. It is not clear how the phage causes
the production of the toxin, as the complicated glycolipid
structure of corynetoxins would require an elaborate
metabolic pathway to be expressed from the phage genome.
However, associations between phage and bacterial
virulence factors are quite common, being responsible for
much of the virulence associated with such important
human pathogens as Corynebacterium diphtheria,
Escherichia coli, and Vibrio cholerae. Note that the
likelihood of phage conversion of non-toxogenic C. toxicus
strains into toxigenic strains seems to be enhanced via the
. application various herbicides at the same time as phage
exposure. Also of interest, phage Xf and Xf-2 infection of
Phage Ecology and Plants 247

X. campestris results in an increase in this bacterium's


virulence towards rice.
PHAGE THERAPY

Phage have been proposed as plant-pathogen control


agents in a process known as phage therapy: the application
of specific phages to specific ecosystems in order to reduce
the population size of specific bacteria. That is, phage
therapy is a form of biological control-the use of one
organism to suppress another. Like other methods of
biological control, one advantage of phage therapy is a
reduction in the usage of chemical agents against pest
species, which, in the case of phage, means a reduction in
the usage of chemical antibiotics.
Phage therapy was ~xPlored extensively by early phage
workers as a means of controlling plant pathogens.
Circumstances in which phage therapy of plants or plant
products has been attempted include against Salmonella
associated with freshcut fruit, to disinfest Streptomyces
scabies-infected potato seedtuber, against bacterial leaf spot
of mungbeans, against Xanthomonas pruniassociated
bacterial spot of peaches, to control of X. campestris
infections of peach trees as well as cabbage and pepper
diseases, to control Ralstonia solanacearum, and to control
soft rot and fire blight associated with Erwinia.
Phage therapy has been used successfully against
bacterial blotch of mushrooms caused by Pseudomonas
tolaasii. In studies notable for the employment of phage
hostrange mutants, phage therapy has also been employed
against bacterial blight of geraniums and bacteria spot of
tomatoes, both caused by pathovars of X. campestris. Phage
can also be used to bias the survival of more-effective
mutualistic bacteria. Basit et al., for example, have isolated
phage that are ineffective against a preferred inoculum of
B. japonicum but effective against naturally occurring
competitors. By coating seeds with phage effective only
248 Plant Ecology

against these potential competitors they can enhance


nitrogen fixation.
Though seemingly effective in certain situations, it is
likely that phage therapy againSt bacterial plant pathogens
will not prove to be a "magic bullet" in all cases. Johnson
proposed a general biological control model which suggests
that the success of a particular treatment will be influenced
by agent and target densities. An important component of
this model is the possibility of the target residing in spatial
refuges into which the biological control agent cannot
penetrate.
We would propose several additional factors that could
contribute to the success or failure of a potential phage
therapy system, such as the location or niche in which the
target pathogen population resides (including the potential
for refuges), the presence of adequate water as a medium
for virion diffusion, rates of virion decay, the timing of
phage application, phage in situ infection fecundity, and
the relative fitness of phageresistant bacterial mutants.
Furthermore, due to the diversity of bacteria and their
phages, extrapolation of phage therapy practices from one
pathogen system to different systems will not always be
practicable.
REFERENCES

Abedon, S. T. 'Phage Ecology", In R Calendar and S. T. Abedon (eds.),


The Bacteriophages. Oxford University Press, Oxford. 2006.
Ackermann, H.-W., and M. S. DuBow. Viruses of Prokaryotes, CRC Press,
Boca Raton, Florida. 1987.
Chibani-Chennoufi, 5., A. Bruttin, M. L. Dillmann, and H. Briissow.
"Phage-host interaction: an ecological perspective" J. Bacteriol.
186:3677-3686. 2004.
Goyal, S. M., C. P. Gerba, and G. Bitton. Phage Ecology. CRC Press, Boca
Raton, Florida. 1987.
Paul, J. H., and C. A. Kellogg. "Ecology of bacteriophages in nature", p.
211-246. In C. J. Hurst (ed.), Viral Ecology. Academic Press, San
Diego. 2000.
11
Ecology of Plant Diseases

Plant pathology is the scientific study of plant diseases


caused by pathogens (infectious diseases) and
environmental conditions (physiological factors).
Organisms that cause infectious disease include fungi,
oomycetes, bacteria, viruses, viroids, virus-like organisms,
phytoplasmas, protozoa, nematodes and parasitic plants.
Not included are insects, mites, vertebrate or other pests
that affect plant health by consumption of plant tissues.
Plant pathology also involves the study of the
identification, etiology, disease cycle, economic impact,
epidemiology, how plant diseases affect humans and
animals, pathosystem genetics and management of plant
diseases.
Plant pathogens cause mortality and reduce fecundity
of individual plants, drive host population dynamics, and
affect the structure and composition of natural plant
communities. Pathogens are responsible for both numerical
changes in host populations and evolutionary changes
through selection for resistant genotypes. Unking such
ecological and evolutionary dynamics has been the focus of
a growing body of literature on the effects of plant diseases
in natural ecosystems. A guiding principle is the
importance of understanding the spatial and temporal
scales at which plants and pathogens interact. This chapter
describes the ecology of plant diseases in natural
250 Plant Ecology

ecosystems, focusing in particular on the effects of diseases


on populations of plants, the maintenance of plant species
diversity as well as the process of rapid evolutionary
changes in host-pathogen symbioses.
ORGANISMS CAUSING PLANT DISEASES

Phytopathogenic Fungi

The majority of phytopathogenic fungi belong to the


Ascomycetes and the Basidiomycetes. The fungi reproduce
both sexually and asexually via the production of spores.
These spores may be spread long distances by air or water,
or they may be soil bourne. Many soil bourne spores,
normally zoospores and capable of living saprotrophically,
caring out the first part of their lifecycle in the soil. Fungal
diseases can be controlled through the use of fungicides in
agriculture, however new races of fungi often evolve that
are resistant to various fungicides.

Oomycetes
The oomycetes are fungal-like organisms that until recently
used to be mistaken for fungi. They include some of the
most destructive plant pathogens including the genus
Phytophthora which includes the casual agents of potato
lat~ blight and sudden oak death. Despite not being closely
related to the fungi, the oomycetes have developed very
similar infection strategies and so many plant pathologists
group them with fungal pathogens.

Bacteria
Most bacteria that are associated with plants are actually
saprotrophic, and do no harm to the plant itself. However,
a small number, around 100 species, are able to cause
disease. Bacterial diseases are much more prevalent in sub-
tropical and tropical regions of the world.
Ecology of Plant Diseases 251

Most plant pathogenic bacteria are rod shaped (bacilli).


In order to be able to colonise the plant they have specific
pathogenicity factors. There are 4 main bacterial
pathogenicity factors:
1. Cell wall degrading enzymes. Used to break down the
plant cell wall in order to release the nutrients inside.
Used by pathogens such as Erwinia to cause soft rot.
2. Toxins. These can be non-host specific, and damage all
plants, or host specific and only cause damage on a host
plant.
3. Phytohormones. For example agrobacterium changes the
level of Auxin to cause tumours.
4. Exopolysaccharides. These are produced by bacteria and
block xylem vessels, often leading to the death of the
plant.

Phytoplasmas and Spiroplasmas


Phytoplasma, formerly known as 'Mycoplasma-like
organisms' or MLOs, are specialised bacteria that are
obligate parasites of plant phloem tissue, and some insects.
They were first discovered by scientists in 1967 when they
were named mycoplasma like organisms or MLOs. They
can't be cultured in vitro in cell-free media. They are
characterised by their lack of a cell wall, a pleiomorphic or
filamentous shape, normally with a diameter less than 1
micrometer, and their very small genomes.
Phytoplasmas are pathogens of important crops,
including coconuts and sugarcane, causing a wide variety
of symptoms that ranges from mild yellowing to death of
infected plants. They are most prevalent in tropical and
sub-tropical regions of the world. Phytoplasmas require a
vector to be transmitted from plant to plant and this
normally takes the form of sap sucking insects such as leaf
hoppers in which they are also able to replicate.
252 Plant Ecology

Phytoplasmas were first identified in 1967 in plants that


were thought to be infected with viruses, but ultrathin
sections of the plants phloem revealed the presence of
mycoplasma like organisms. Being mollicutes, phytplasmas
lack cell walls and instead are bound by a triple layered unit
membrane. The cell membranes of all phytoplasmas
studied so far usually contain a single immunodominant
protein that makes up the majority of the protein content
of the cell membrane. Their shape is normally pleiomorphic
or filamentous and normally have a diameter of less than
1 micrometer. Like other prokaryotes, DNA is free in the
cytoplasm. They are believed to reproduce through Binary
fission.

Symptoms
A common symptom caused by phytoplasma infection is
phyllody, the production of leaf like structures in place of
flowers. Evidence suggests that the phytoplasma
downregulates a gene involved in petal formation and
genes involved in the maintenance of the apical meristem.
This causes sepals to form where petals should. Other
symptoms, such as the yellowing of leaves, are thought to
be caused by the phytoplasma's presence in the phloem
affecting its function, and changing the transport of
carbohydrates.
Phytoplasma infected plants may also suffer from
virescence - the development of green flowers due to the
loss of pigment in the petal cells. Many phytoplasma
infected plants gain a bushy or witch's broom appearance
due to changes in normal growth patterns caused by the
infection. Most plants show apical dominance but
phytoplasma infection can cause the proliferation of
auxiliary (side) shoots and an increase in size of the
internodes. Such symptoms are actually useful in the
commercial productiun of poinsettia. The infection is
necessary to produce more axillary shoots that enable to
Ecology of Plant Diseases 253

production of pionsettia plants that have more than one


flower.
Phytoplasmas may cause many other symptoms that
are induced because of the stress placed on the plant by
infection rather than specific pathogencity of the
phytoplasma. Photosynthesis, especially photosystem II, is
inhibited in many phytoplasma infected plants.
Phytoplasma infected plants often show yellowing which
is caused by the breakdown of chlorophyll, whose
biosynthesis is also inhibited.

Transmission
Movement between plants
The phytoplasmas are mainly spread by insects of the
families Cicadellidea (leafhoppers) and Fulgoridea
(planthoppers) which feed on the phloem tissues of
infected plants picking up the phytoplasmas and
transmitting them to the next plant they feed on. For this
reason the host range of phytoplasmas is strongly
dependent upon its insect vector. Phytoplasmas contain a
major antigenic protein that makes up the majority of their
cell surface proteins and this has been shown to interact
with insect microfilament complexes and,is believed to the
determining factor is insect-phytoplasma interation.
Phytoplasmas may overwinter in insect vectors or perrinial
plants. Phytoplasmas can have varying affects on their
insect hosts, examples of both reduced and increased fitness
have been seen.
Phytoplasmas will be found in most of the major organs
of an infected insect host once they are established. They
will enter the insects body through the stylet and then move
through the intestine and bein absorbed into the
haemolymph. From here they proceeded to colonise the
salivary glands, a process that can take up to three weeks.
The time between phytoplasmas being taken up by the
254 Plant Ecology

insect and the phytoplasmas reaching an infectious titre in


the salivary gland is called the latency period. Phyto-
plasmas can also be spread via vegetative propergation
such as the grafting of a piece of infected plant onto a
healthy plant.
Movement within plants
Phytoplasmas are able to move within the pholem from
source to sink and they are able to pass through sieve tube
elements, but spread more slowly than solutes, for this and
other reasons I.110Vement by passive translocation is not
supported.
Detection and Diagnosis
Before molecular techniques were developed the diagnosis
of phytoplasma diseases was difficult due to the fact that
they could not be cultured. Thus classical diagnostic
techniques such as observation of symptoms were used.
Ultrathin sections of the phloem tissue from suspected
phytoplasma infected plants would also be examined for
their presence.Another diagnostic technique used was to
treat infected plants with antibiotics such as tetracycline to
see if this cured the plant.
Molecular diagnostic techniques for the detection of
phytoplasma began to emerge in the 1980s and included
ELISA based methods. In the early 1990's peR-based
methods were developed that were far more sensitive than
those that used ELISA, and RFLP analysiS allowed the
accurate identification of different strains and species of
phytoplasma. There are also techniques that allow the
assessement of the level of infection.
Control
Phytoplasmas are normally controlled by the breeding and
pla!lting of disease resistance varieties of crops (believed to
Ecology of Plant Diseases 255

the most economically viable option) and by the control of


the insect vector. Tissue culture can be used to produce
clones of phytoplasma infected plants that are healthy. The
chances of gaining healthy plants in this manner can be
enhanced by the use of cryotherapy, freezing the plant
samples in liquid nitrogen before using them for tissue
culture.
Work has also been carried out investigating tHe effecti-
veness of plantibodies targeted against phytoplasmas.
Tetracyclines are bacteriostatic to phytoplasmas, that is
they inhibit their growth. However without continuous use
of the antibiotic disease symptoms will reappear. Thus
tetracycline is not a viable control agent in agriculture, but
is used to protect ornamental coconut trees.

Spiroplasma
Spiroplasma is a genus of Mollicutes, a group of small
bacteria without cell walls. Spiroplasma shares the simple
metabolism, parasitic lifestyle, fried-egg colony
morphology and small genome of other Mollicutes, but has
a distinctive helical morphology, unlike Mycoplasma. It has
a spiral shape and moves in a corkscrew motion. Most
spiroplasmas are found either in the gut or hemolymph of
insects, or in the phloem of plants. Spiroplasmas are
fastidious organisms, which require a rich culture medium.
Typically they grow well at 30°C, but not at 37°C. A few
species, notably Spiroplasma mirum, grow well at 37°C
(human body temperature), and cause cataracts and
neurological damage in suckling mice. The best studied
species of spiroplasmas are Spiroplasma citri, the causative
agent of Citrus Stuborn Disease, and Spiroplasma kunkelii,
the causative agent of Com Stunt Disease.
There is some disputed evidence for the role of
spiroplasmas in the etiology of Transmissible Spongiform
Encephalopathies (TSEs), due primarily to the work of Dr.
Bastian, summarized be~ow. Other researchers, such as
256 Plant Ecology

Leach et al. have failed to replicate this work, while the


prion model for TSEs has gained very wide acceptance. The
most recent work of Alexeeva et aL appears to refute the
role of spiroplasmas in the best small animal scrapie model
(hamsters). Bastian et aL have responded to this challenge
with the isolation of a spiroplasma species from scrapie-
infected tissue, grown it in cell-free culture, and
demonstrated its infectivity in ruminants.
According to Frank 0. Bastian, MD:
"spiroplasmas contain internal fibrillar proteins, that have
morphological and immunological similarities to scrapie-
and CJD-related fibrillar proteins. This comparison is
noteworthy since mycoplasmologists consider these fibril
proteins unique to this prokaryoteJn vivo and in vitro
experimental Spiroplasma infections produce cytopathic
effects similar to those of the scrapie agent Experimental
Spiroplasma brain infection in the suckling rat is
characterized by vacuolar encephalopathy with localization
of the microbe to gray matter.[... ] Spiralins are chemically
bound to Spiroplasma-associated fibrils (SpFs) and are
separated with difficulty.' SpFs are unique internal fibrils of
spiroplasmas with a molecular weight of 55 kDa. Recently,
SpFs have been shown to bear close morphological
resemblance to scrapie-associated fibrils (SAPS), ' and show
cross-reactivity using SAP antibody."

In addition, a Spiroplasma species had been shown to kill


males of the Plain Tiger butterfly on infection, leading to
interesting consequences for population genetics and
consequently speciation similar to the effects caused by
some strains of Wolbachia

Plant Viruses
There are many types of plant virus, and some are even
~symptomatic. Normally plant viruses only cause a loss of
yield. Therefore it is not economically viable to try to
:ontrol them, the exception being when they infect
:>erennial species, such as fruit trees.
Ecology of Plant Diseases 257

Most plant viruses have small, single stranded RNA


genomes. These genomes may only encode 3 or 4 proteins:
a replicase, a coat protein, a movement protein to allow cell
to cell movement and sometimes a protein that allows
transmission by a vector.
Plant viruses must be transmitted from plant to plant
by a vector. This is normally an insect, but some fungi,
nematodes and protozoa have been shown to be viral
vectors. Plant viruses need to be transmitted by a vector,
most often insects such as leafhoppers. One class of viruses,
the Rhabdoviridae, have been proposed to actually be
insect viruses that have evolved to replicate in plants. The
chosen insect vector of a plant virus will often be the
determining factor in that virus' host range: it can only
infect plants that the insect vector feeds upon. This was
shown in part when the old world white fly made it to the
USA, where it transferred many plant viruses onto new
hosts.
Depending on the way they are transmitted, plant
viruses are classified as non-persistent, semi-persistent and
persistent. In non-persistent transmission, viruses become
attached to the distal tip of the stylet of the insect and on
the next plant it feeds on, it inoculates it with the virus.
Semi-persistent viral transmission involves the virus
entering the foregut of the insect. Those viruses that
manage to pass through the gut into the haemolymph and
then to the salivary glands are known as persistent. There
are two sub-classes of persistent viruses: propergative and
circulative. Propergative viruses are able to replicate in both
the plant and the insect (and may have originally been
insect viruses), whereas circulative can not.
Many plant viruses encode within their. genome
polypeptides with domains essential for transmission by
insects. In non-persistent and semi-persistent viruses, these
domains are in the coat protein and another protein known
as the helper component. A bridging hypothesis has been
258 Plant Ecology

proposed to explain how these proteins aid in insect-


mediated viral transmission. The helper component will
.bind to the specific domain of the coat protein, and then the
insect mouthparts--creating a bridge.
In persistent propergative viruses, such as tomato
spotted wilt virus (TSWV), there is often a lipid coat
surrounding the proteins that is not seen in the other classes
of plant viruses. In the case of TSWV, 2 viral proteins are
expressed in this lipid envelope. It has been proposed that
the viruses bind via these proteins and are then taken into
the insect cell by receptor-mediated endocytosis.
The discovery of plant viruses causing disease is often
accredited to Martinus Beijerinck who discoved, in 1898,
that even after passing infective tree sap through a
porcelain filter remained infectious but was sterile of
microorganisms.
After the initial discovery of the 'viral concept' there
was need to classify any other known viral diseases based
on the mode of transmission even though microscopic
observation proved fruitless. In 1939 Holmes published a
classification list of 129 plant viruses. This was expanded
and in 1999 there were 977 officially recognised, and some
provisional, plant virus species.
The pUrification of the TMV (the first purification) was
first performed by Wendell Stanley, who published his
findings in 1936. He later was accredited with the Nobel
Prize in Chemistry in 1946. In the 1950s a discovery by two
labs simultaneously proved that the purified RNA of the
TMV was infectious which reinforced the argument, that
had a lot of opposition at the time, that RNA was carrying
genetic information to ('ncie for the production of new
infectious particles.
More recently the research has been focused on the
manipulation and modification of plant virus genomes do
discover function and for commercial gain in the
agriculture business by using viral-derived sequences to
Ecology of Plant Diseases 259

provide understanding of novel forms of resistance. The


recent boom in technology allowing humans to manipulate
plant viruses has really helped bring the subject out of an
Aristolean science age and into the 21st century.
Structure
Viruses are very small and can only be seen under an
electron microscope. The ~tructure of a virus is given by its
coat of proteins, which surround the viral genome.
Assembly of viral particles takes place spontaneously. Over'
50% of known plant viruses are rod shaped. Exact length
is normally dependent on the genome but it is usually
between 300-500 nm with a diameter of 15-20 nm. Protein
subunits can be placed around the circumference of a circle
to form a disc. In the presence of the viral genome, the discs
are stacked, then a tube is created with room for the nucleic
acid genome in the m1ddle
The second most common structure amongst plant
viruses a~e isometric particles. They are 40-50 nm in
diameter. In cases when there is only a single coat protein,
the basic structure consists of 60 T subunits, where T is an
integer. Some viruses may have 2 coat proteins are the
formation of the particle is analogous to a football.
There are three genera of Geminiviridae that possess
geminate particles which are like two isometric particles
stuck together. A very small number of plant viruses have,
in addition to their coat proteins, a lipid envelope. This is
derived from the plant cell membrane as the virus particle
buds off from the cell.

Transmission of plant viruses

Through sap
It implies direct transfer of sap by contact of and wounded
plant with a healthy one. Such process occurs during
260 Plant Eco~ogy

agricultural practices by tools, hands, or by animal feeding


on the plant. Generally TMV, potato viruses and cucumber
mosaic viruses are transmitted via sap.

Insects
Plant viruses need to be transmitted by a vector, most often
insects such as leafhopperso One class of viruses, the
Rhabdoviridae, have been proposed to actually be insect
viruses that have evolved to replicate in plantso The chosen
insect vector of a plant virus will often be the determining
factor in that virus' host range: it can only infect plants that
the insect vector feeds upon. This was shown in part when
the old world white fly made it to the USA, where it
transferred many plant viruses onto new hosts.
Depending on the way they are transmitted, plant
viruses are classified as non-persistent, semi-persistent and
persistent. In non-persistent transmission, viruses become
attached to the distal tip of the stylet of the insect and on
the next plant it feeds on, it inoculates it with the virus.[2]
Semi-persistent viral transmission involves the virus
entering the foregut of the insect.
Those viruses that manage to pass through the gut into
the haemolymph and then to the salivary glands are known
as persistent. There are two sub-classes of persistent
viruses: propergative and circulative. Propergative viruses
are able to replicate in both the plant and the insect,
whereas circulative can not.
Many plant viruses encode within their genome
polypeptides with domains essential for transmission by
insects. In non-persistent and semi-persistent viruses, these
domains are in the coat protein and another protein known
as the helper component. A bridging hypothesis has been
proposed to explain how these proteins aid in insect-
mediated viral transmission. The helper component will
bind to the specific domain of the coat protein, and then the
insect mouthparts - creating a bridge.
Ecology of Plant Diseases 261

In persistent prop ergative viruses, such as tomato


spotted wilt virus (TSWV), there is often a lipid coat
surrounding the proteins that is not seen in the other classes
of plant viruses. In the case of TSWV, 2 viral proteins are
expressed in this lipid envelope. It has been proposed that
the viruses bind via these proteins and are then taken into
the insect cell by receptor-mediated endocytosis.

Plasmodiophorids
A number of viral genera are transmitted, both persistently
and non-persistently, by soil bourne zoosporic protozoa.
These protozoa are not phytopathogenic themselves, but
parasitic. Transmission of the virus takes place when they
become associated with the plant roots.

Seed and pollen borne viruses


Plant virus transmission from generation to generation
occurs in about 20% of plant viruses. When viruses are
transmitted by seeds, the seed is infected in the generative
cells and the virus is maintained in the germ cells and
sometimes, but less often, in the seed coat. When the
growth and development of plants is delayed because of
situations like unfavourable weather, there is an increase in
the amount of virus infections in seeds. There does not seem
to be a correlation between the location of the seed on the
plant and its chances of being infected.
Little is known about the mechanisms involved in the
transmission of plant viruses via seeds, although it is
known that it is environmentally influenced and that seed
transmission occurs because of a direct invasion of the
embryo via the ovule or by an indirect route with an attack
on the embryo mediated by infected gametes. These
processes can occur concurrently or separately depending
on the host plant. It is unknown how the virus is able to
directly invade and cross the embryo and boundary
between the parental and progeny generations in the ovule.
262 Plant Ecology

Many plants species can be infected through seeds


including but not limited to the families Leguminoseae,
Solanacease, Compositae, Rosaceae, Curcurbitaceae,
Gramineae.

Nematodes
Nematodes are small, multicelluar wormlike creatures.
Many live freely in the soil, but there are some species
which parasitize plant roots. They are mostly a problem in
tropical and subtropical regions of the world, where they
may infect crops. Root knot nematodes have quite a large
host range, whereas cyst nematodes tend to only be able to
infect a few species. Nematodes are able to cause radical
changes in root cells in order to facilitate their lifestyle.

Protozoa
There are a few examples of plant diseases caused by
protozoa. They are transmitted as zoospores which are very
durable, and may be able to survive in a resting state in the
soil for many years. They have also been shown to transmit
plant viruses. When the motile zoospores come into contact
with a root hair they produce a plasmodium and invade the
roots.
PHYSIOLOGICAL PLANT DISORDERS

Physiological plant disorders are caused by non-


pathological disorders such as poor light, weather damage,
water-logging or a lack of nutrients, and affect the
functioning of the plant system. Physiological disorder are
distinguished from plant diseases caused by pathogens,
such as a virus or fungus. Whilst the symptoms of
physiological disorders may appear disease-like, they can
usually be prevented by altering environmental conditions.
However, once a plant shows symptoms of nutrient
deficiency it is likely that that season's yields will be
reduced.
Ecology of Plant Diseases 263

Causes of physiological disorders can be identified by


examining:
Where symptoms first appear on a plant- on new
leaves, old leaves or all over?
The pattern of any discolouration or yellowing- is it all
over, between the veins or around the edges? If only
the veins are yellow deficiency is probably not
. involved.
Note general patterns rather than looking at individual
plants- are the symptoms distributed throughout a
group of plants of tPte same type growing together. In
the case of a deficiency all of the plants should be
similarly effected, although distribution will depend
on past treatments applied to the soil.
Soil analysis, such as determining pH, can help to
confirm the presence of physiological disorders. Recent
conditions, such as heavy rains, dry spells, frosts, etc,
may also help to determine the cause of plant
disorders.
WEATHER DAMAGE

Frost and cold are major causes of crop damage to tender


plants, although hardy plants can also suffer if new growth
is exposed to a hard frost following a period of warm
weather. Symptoms will often appear overnight, affecting
many types of plants. Leaves and stems may tum black, and
buds and flowers mar be discoloUfed, and frosted blooms
may not produce fruit. Many annual plants, or plants
grown in frost free areas, can suffer from damage when the
air temperature drops below 400 Fahrenheit. Tropical plants
may begin to experience cold damage when the tempe-
rature is 42-480 Fahrenheit, symptoms include wilting of the
top of the stems and/ or leaves, and blackening or softening
of the plant tissue.
Frost or cold damage can be avoided by ensuring that
tender plants are properly hardened before planting, ~nd
264 Plant Ecology

that they are not planted too early in the season, before the
risk of frost has passed. Avoid planting susceptible plants
in frost pockets, or where they will receive early morning
sun. Protect young buds and bloom with horticultural
fleece if frost is forecast. Cold, drying easterly winds can
also severely inhibit spring growth even without an actual
frost, thus adequate shelter or the use of windbreaks is
important.
Drought can cause plants to suffer from water stress
and wilt. Adequate irrigation is required during prolonged
hot, dry periods. Rather than shallow daily watering,
during a drought water should be directed towards the
roots, ensuring that the soil is thoroughly soaked two or
three times a week. Mulches also help preserve soil
moisture and keep roots cool.
Heavy rains, particularly after prolonged dry periods,
can also cause roots to split, onion saddleback, tomatoes
split and potatoes to become deformed or hollow. Using
mulches or adding organic matter such as leaf mold,
compost or well rotted manure to the soil will help to act
as a 'buffer' between sudden changes in conditions. Water-
logging can occur on poorly drained soils, particularly
following heavy rains. Plants can become yellow and
stunted, and will tend to be more prone to drought and
diseases. Improving drainage will help to alleviate this
problem. Hail can cause damage to soft skinned fruits, and
may also allow brown rot or other fungi to penetrate the
plant. Brown spot markings or lines on one side of a mature
apple are indicative of a spring hailstorm. Plants affected
by salt stress are unable to take water from soil, due to an
osmotic imbalance between soil and plant.
NUTRIENT DEFICIENOES

Poor growth and a variety of complaints such as leaf


discolouration can be caused by a lack of plant foods. This
may be due to shortages of necessary nutrients, or because
Ecology of Plant Diseases 265

the nutrients are present but not available to the plant. Th~
latter can be caused by incorrect pH, shortages of water or
an excess of another nutrient. Generally, the key to avoiding
nutrient deficiencies is to ensure that the soil is healthy and
contains plenty of well rotted organic matter rather than by
feeding or treating individual plants.

Boron Deficiency
Boron (B) deficiency is a rare disorder affecting plants
growing above a granite bedrock, which is low in boron.
Boron may be present but locked up in soils with a high pH,
and the deficiency may be worse in wet seasons. Symptoms
include dying growing tips and bushy stunted growt~.
Crop-specific symptoms include:
Beetroot: rough, cankered patches on roots, internal
brown rot.
Cabbage: distorted leaves, hollow areas in stems.
Cauliflower: poor development of curds, and brown
patches. Stems, leafstalks and midribs roughened.
Celery: leaf stalks develop cracks on the upper surface,
inner tissue is reddish brown.
Pears: new shoots die back in spring, fruits develop
hard brown flecks in the skin.
Strawberries: Stunted growth, foliage small, yellow and
puckered at tips. Fruits are small and pale.
Swede (rutabaga> and turnip: brown or grey concentric
rings develop inside the roots.
Arecaceae: brown spots on fronds & lower productivity.
Boron deficiency can be avoided by improving the moisture
retaining capacity of light soils, and ensuring pH is kept
below 7. Borax can be raked into the soil at 35 g/m2.
Calcium Deficiency
Calcium (Ca) deficiency is a plant disorder that can be
266 Plant Ecology

caused iJy insufficient calcium in the growing medium, but


is more frequently a product of a compromised nutrient
mobility system in the plant. This may be due to water
shortages, which slow the transportation of calcium to the
plant, or can be caused by excessive usage of potassium or
nitrogen fertilizers.
Calcium deficiency symptoms appear initially as
generally stunted plant growth, necrotic leaf margins on
young leaves or curling of the leaves, and eventual death
of terminal buds and root tips. Generally the new growth
of the plant is affected first. The mature leaves may be
affected if the problem persists.
Crop-specific symptoms include:
Apple. 'Bitter pit'-fruit skins develop pits, brown
patches appear in flesh and taste becomes bitter. Can
occur when fruit is in storage. Bramley apples are
particularly susceptible.
Cabbage and Brussels sprouts. Internal browning.
Carrot. Cavity spot'-oval spots develop into craters
which may be invaded by other disease causing
organisms. ,
Celery. Stunted growth, central leaves stunted.
Tomatoes and peppers. 'Blossom end rot'-Symptoms
start as sunken, dry decaying areas at the blossom end
of the fruit, furthest away from the stem, not all fruit
on a truss is necessarily affected. Sometimes rapid
growth from high-nitrogen fertilizers may cause
blossom end rot.

Treatment
Calcium deficiency can be rectified by adding Agricultural
lime to acid soils, aimittg at a pH of 6.5, unless the plant
in question specifically prefers acidic soil. Organic matter
should be added to the soil in order to improve its moisture-
retaining capacity.
Ecology of Plant Diseases 267

Plant damage is difficult to reverse, so take corrective


action immediately. Make supplemental applications bf
calcium nitrate at 200 ppm nitrogen. Test and correct the pH
if needed because calcium deficiency is often associated
with low pH.

Iron Deficiency
Iron (Fe) deficiency is a plant disorder also known as 'lime-
induced chlorosis'. A deficiency in the soil is rare. Iron can
be unavailable if pH is too high or if the soil is waterlogged,
or has been overfertilised with phosphorus. Can be
confused with manganese deficiency.
Any plants may be affected, but raspberries and pears
are particularly susceptible, as well as most acid-loving
plants such as azaleas and camellias.
Symptoms include leaves turning yellow or brown in
the margins between the veins which may remain green,
while young leaves may appear to be bleached. Fruit is of
poor quality and quantity.
Iron deficiency can be avoided by choosing appropriate
soil for the growing conditions (e.g., avoid growing acid
loving plants on lime soils), or by adding well-rotted
manure or compost.

Magnesium Deficiency
Magnesium (Mg) deficiency is a plant disorder. Magnesium
can be easily washed out of light soils in wet seasons.
Excessive potassium fertiliser usage can cause also Mg to
become unavailable to the growing plant. This disorder
partic~arly affects potatoes, tomatoes, apples, currants <;md
gooseberries, and chrysanthemums.
Symptoms include, yellowing between leaf veins,
which stay green, giving a marbled appearance. This begins
with older leaves from late June, but spreads to younger
growth. Can be confused with virus, or natural aging in the
268 Plant Ecology

case of tomato plants. Fruits are small and woody. Mg


deficiency can be rectified in the short term by applying a
foliar feed fortnightly, with epsom salts diluted at a rate of
200g per 10 litres of water (807 per 21;2 gal) after flowering.
In the longer term add dolomitic limestone if soil pH
allows, or other Mg containing rocks such as Kieserite.
Reduce usage of potash fertilsers if this may be contributing
to the problem.

Manganese Deficiency
Manganese (Mn) deficiency is a plant disorder that is often
confused with, and occurs with, iron deficiency. Most
common in poorly drained soils, also where organic matter
levels are high. Manganese may be unavailable to plants
where pH is high.
Affected plants include onion, apple, peas, French
beans, cherry and raspberry, and symptoms include
yellowing of leaves with smallest leaf veins remaining
green to produce a 'chequered' effect. The plant may seem
to grow away from the problem so that younger leaves may
appear to be unaffected. Brown spots may appear on leaf
surfaces, and severely affected leaves turn brown and
wither.

Nitrogen Deficiency
Nitrogen (N) deficiency in plants can occur when woody
material such as sawdust is added to the soil. Soil
organisms will utilise any nitrogen in order to break this
down, thus making it temporarily unavailable to growing
plants. 'Nitrogen robbery' is more likely on light soils and
those low in organic matter content, although all soils are
susceptible. Cold weather, especially early in the season,
can also cause a temporary shortage.
All vegetables apart from nitrogen fixing legumes are
prone to this disorder. Symptoms include poor plant
Eculogy of Plant Diseases 269

growth, leaves are pale green or yellow in the case of


brassicas. Lower leaves show symptoms first. Leaves in this
state are said to be etiolated with reduceq chlorophyll.
Flowering and fruiting may be delayed.
Prevention and control of nitrogen deficiency can be
achieved in the short term by using grass mowings as a
mulch, or foliar feeding with manure, and in the longer
term by building up levels of organic matter in the soil.
Sowing green manure crops such as grazing rye to cover
soil over the winter will help to prevent nitrogen leaching,
while leguminous green manures such as winter tares will
fix additional nitrogen from the atmosphere.

Phosphorus Deficiency
Phosphorus (P) deficiency is a plant disorder that is most
common in areas of high rainfall, especially on acid, clay
or poor chalk soils. Cold weather can cause a temporary
deficiency.
All plants may be affected, although this is an
uncommon disorder. Particularly susceptible are carrots,
lettuce, spinach, apples, currants and gooseberries.
Symptoms include poor growth, and leaves that turn blue/
green but not yellow-oldest leaves are 'affected first. Fruits
are small and acid tasting. Phosphorus deficiency may be
confused with nitrogen deficiency. It can be controlled by
applying organic sources of phosphorus such as rock
phosphate. Plants that are naturally adapted to low levels
of available soil phosphorus, however, are more likely to
suffer from phosphate poisoning: the key is to provide the
right level for any particular plant type, neither too high nor
too low.

Potassium Deficiency
Potassium deficiency, also known as potash deficiency, is
a plant disorder that is most common on light, sandy soils,
270 Plant Ecology

as well as chalky or peaty soils with a low clay content. It


is also founden--heavy clays with a poor structure.
Plants require potassium ions (K+) for protein synthesis
and for the opening and closing of stomata, which is
regulated by proton pumps to make surrounding guard
cells either turgid or flaccid. A deficiency of potassium ions
can impair a plants ability to maintain these processes.
The deficiency most commonly affects fruits and
vegetables, notably potatoes, tomatoes, apples, currants,
and gooseberries, and typical symptoms are brown
scorching and curling of leaf tips, and yellowing of leaf
veins. Purple spots may also appear on the leaf undersides.
Deficient plants may be more prone to frost damage and
disease, and their symptoms can often be confused with
wind scorch or drought. Prevention and cure can be
achieved in the shorter term by feeding with home-made
comfrey liquid, adding seaweed meal, composted bracken
or other organic potassium-rich fertilisers. In the longer
term the soil structure should be improved by adding
plenty of well rotted compost or manure. Wood ash has
high potassium content, but should be composted first as
it is in a highly soluble form.
EPIDEMIC DISEASES OF PLANTS

Plant Disease epidemiology is the study of disease in plant


populations. Much like diseases of humans and animals,
plant diseases occur due to pathogens such as bacteria,
viruses, fungi, oomycetes, nematodes, phytoplasmas,
protozoa, and parasitic plants. Plant disease
epidemiologists strive for an understanding of the cause
and effects of disease and develop strategies to intervene
in situations where crop losses may occur. Typically
successful intervention will lead to a low enough level of
disease to be acceptable, depending upon the value of the
crop.
Ecology of Plant Diseases 271

Plant disease epidemiology is often looked at from a


multi-disciplinary approach, requiring biol9gical,
statistical, agronomic and ecological perspectives. Biology
is necessary for understanding the pathogen and its life
cycle. It is also necessary for understanding the physiology
of the crop and how the pathogen is adversely affecting it.
Agronomic practices often influence disease incidence for
better or for worse.
Ecological influences are numerous. Native species of
plants may serve as reservoirs for pathogens that cause
disease in crops. Statistical models are often applied in
order to summarize and describe the complexity of plant
disease epidemiology, so that disease processes can be more
readily understood. For example, comparisons between
patterns of disease progress for different diseases, cultivars,
management strategies, or environmental settings can help
in determining how plant diseases may best be managed.
Policy can be influential in the occurrence of diseases,
through actions such as restrictions on imports from
sources where a disease occurs.
In 1963 J. E. van der Plank published "Plant Diseases:
Epidemics and Control", a seminal work that created a
theoretical framework for the study of the epidemiology of
plant diseases. This book provides a theoretical framework
based on experiments in many different host pathogen
systems and moved the study of plant disease
epidemiology forward rapidly, especially for fungal foliar
pathogens. Using this framework we can now model and
determine thresholds for epidemics that take place in a
homogeneous environment such as a mono-cultural crop
field.
Disease epidemics in plants can cause huge losses in
yield of crops as well threatening to wipe out an entire
species such as was the case with Dutch Elm Disease and
could occur with Sudden Oak Death. An epidemic of potato
late blight, caused by Phytophthora infestans, led to the
Great Irish Famine and the loss of many lives.
272 Plant Ecology

Commonly the elements of an epidemic are referred to


as the "disease triangle": a susceptible host, pathogen, and
conducive environment. For disease to occur all three of
these must be present. Below is an illustration of this point.
Where all three items meet there is disease. The fourth
element missing from this illustration for an epidemic to
occur, is time. As long as all three of these elements are
present disease can initiate, an epidemic will only ensue if
all three continue to be present. Anyone of the three might
be removed from the equation though. The host might out-
grow susceptibility as with high-temperature adult-plant
resistance, the environment changes and is not conducive
for the pathogen to cause disease, or the pathogen is
controlled through a fungicide application for instance.
Sometimes a fourth factor of time is added as the time
at which a particular infection occurs, and the length of time
conditions remain viable for that infection, can also play an
important role in epidemics. The age of the plant species
can also playa role, as certain species change in their levels
of disease resistance as they mature; a process known as
ontogenic resistance.
If all of the criteria are not met, such as a susceptible
host and pathogen are present but the environment is not
conducive to the pathogen infecting and causing disease,
disease cannot occur. For example, com is planted into a
field with com residue that has the fungus Cercospora zea-
maydis, the causal agent of Grey leaf spot of com, but if the
weather is too dry and there is no leaf wetness the spores
of the fungus in the residue cannot germinate and initiate
--infection.
Likewise, it stands to reason if the host is susceptible
and the environment favours the development of disease
but the pathogen is not present there is no disease. Taking
the example above, the com is planted into a ploughed field
where there is no com residue with the fungus Cercospora
zea-maydis, the causal agent of Grey leaf spot of com,
Ecology of Plant Diseases 273

present but the weather means long periods of leaf wetness,


there is no infection initiated.
When a pathogen requires a vector to be spread then
for an epidemic to occur the vector must be plentiful and
active. Monocyclic epidemics are caused by pathogens with
a low birth rate and death rate meaning they only have one
infection cycle per season. They are typical of soil born
diseases such as Fusarium wilt of flax. Polycyclic epidemics
are caused by pathogens capable of several infection cycles
a season. These are most often caused by airborne diseases
such as powdery mildew. Bimodal polycyclic epidemics
can also occur. For example in brown rot of stone fruits the
blossoms and the fruits are infected at different times.
For some diseases it is important to consider the disease
occurrence over several growing seasons, especially if
growing' the crops in monoculture year after year or
growing perennial plants. Such conditions can mean that
the inoculum produced in one season can be carried over
to the next leading to a build of an inoculum over the years.
In the tropics there are no dear cut breaks between growing
seasons as there are in temperate regions and this can lead
to accumulation of innoculum.
Epidemics that occur under these conditions are
referred to as polyetic epidemics and can be caused by both
monocylcic and polycyclic pathogens. Apple powdery
mildew is an example of a polyetic epidemic caused by a
polycyclic pathogen and Dutch Elm disease a polyetic
epidemic caused by a monocyclic pathogen.
REFERENCES

Agrios, George. Plant Pathology. Academic Press. 2005.


Arneson, PA "Plant disease epidemiology: temporal aspects". Plant
Health Instructor. American Phytopathological Society. 2001.
Dickinson, M . Molecular Plant Pathology. BIOS Scientific Publishers.
2003.
George N. Agrios. Plant Pathology, Academic Press. New York. 1997.
274 Plant Ecology

Madden, Laurence; Gareth Hughes, Frank Van Den Bosch. Study of Plant
Disease Epidemics. American Phytopathological Society. 2007.
Milton Zaitlin. Discoveries in Plant Biology, New York 14853, USA. Pp.:
105-110. 1998.
Drenth, A "Fungal epidemics - does spatial structure matter?". New
Phytologist 163: 4-7. 2004.
Shultz, T.R; Line, R.F "High-Temperature, Adult-Plant Resistance to
Wheat Stripe Rust and Effects on Yield Components". Agronomy
Journal 84: 170-175. American Society of Agronomy. 1992.
12
Plant Ecology and Climate Change

In contemporary ecology, there are at least four prominent


research speciality areas that study vegetation change:
succession ecology, invasion biology, gap/patch
dynamics, and global change effects on plant
communities.
The underlying processes studied in each of these areas
are basically the same. First, colonization, establishment,
turnover, persistence, and spread' are fundamental events
and processes that interact to produce vegetation change
in all four subdisciplines; second, whatever the nature of
vegetation change, it is often initiated or greatly in.uenced
by, disturbance and/ or changes in interactions with other
trophic levels; third, local and long-distance dispersal
allow new species to enter existing plant communities;
fourth, facilitation and inhibition, as well as interactions
with species from other trophic levels, strongly in.uence
vegetation change; and .fth, in all cases, changes in
community composition affect, and are affected by,
ecosystem processes (Figure 1).
These four research areas focus on different causes of
vegetation change, e.g., species introduced from other
regions of the world, disturbances that create gaps and
initiate succession, and global change. Given that these
four research areas seekto illuminate the mechanisms that
cause vegetation change, and that the phenomena under
276 Plant Ecology

study often interact (e.g., gaps and climate change may


facilitate invasions), one would. expect there to be
considerable information exchange among these research
areas.
Time

I ntrod uction of New Dis p&f&al Within Introduction of Hew


Native Species The Community Non-Natiw Species.

OngOing or
Inlt'l11lltll!nt
C1ISp&rsal
111~llr
tll!ltl!tl!l!tl!!l!l

Mutualisms
Herbivory
Disease
Predation
Re produ ction
DisperSoal

Climate Change

Resou rce Microbes


Fluctuations Pollinators
HerolVofll.

Changes in establishment. spread. pgrsisfllnce,and ecosystem procssSGs


).
Vegetation change

Figure 1. The same factors change plant communities regardless of the


speciality area in which the research is conducted.
Plant Ecology and Climate Change 277

PLANTS AND CLIMATE

Climatic extremes appear more important than climatic


averages in predicting patterns of vegetation distribution.
The poleward limits of many tree species are determined
by frost sensitivity and freezing tolerance. For example,
-40°C is the lower limit of supercooling for most ring-
porous hardwood trees. Below this temperature
intracellular freezing occurs, killing the cells and therefore
the organism. The northern limit of these tree species
correlates closely with the record low temperature of -
40°C. Conifers, which have a different mechanism of
freezing tolerance, can withstand the temperature of
liquid nitrogen and have a correspondingly more extreme
latitudinal limit to their distribution.
Similarly, chilling-sensitive broad-leaved tropical
evergreen trees are killed at 15°C, and their poleward limit
coincides with record low temperatures of 15°C.
Woodward recognised several major physiognomic types
of vegetation: conifer forest, deciduous forests,
broadleaved evergreen forests, shrubland, and herb-
dominated vegetation, based on the assumptions that
climate governs the low-temperature or low-moisture limit
of distribution, and that the mesic limits of distribution are
determined by the environmental tolerance of a
competitively superior but less tolerant physiognomic
type, Woodward successfully predicted the global
distribution of most major physiognomic types of
vegetation.
These observations suggest that the lowtemperature or
low-moisture limit of a physiognomic type is governed by
the fundamental niche (Le., phYSiological tolerance) of the
plants, whereas additional limits of plant distribution may
be determined by other, more complex factors, probably
mediated by competition. These patterns also indicate that
extreme events, which are a function of environmental
fluctuation, are more useful than average conditions in
278 Plant Ecology

predicting northern limits of physiognomic types. As


climate changes, the location and frequency of these
extreme events will certainly change, leading to changes
in distribution of biomes.
Understanding of patterns of environmental
variability are critical to predictions of future vegetation
distribution. Although extreme events predict
distributions of general physiognomic types, they are
inadequate to predict finer distribution patterns of species
or functional groups of species, i.e., groups of species
which show similar responses to change in environment.
The greater diversity of response by species than by
physiognomic types occurs because each species has a
unique range of environmental conditions under which it
occurs, which seldom coincides precisely with their limits
of physiological tolerance. Correlation of the current
distribution of plants with their current average climatic
conditions has been the main basis of explaining past
patterns of vegetation change and predictions of future
vegetation distribution.
These studies have led to several important
generalisations:
(i) Each species shows a unique pattern of distribution
with climate and responds most strongly to different
patterns of climatic factors, so that complex climatic
changes cause species to migrate with different
patterns and to form new associations. Thus, current
plant communities are temporary associations among
species that last only hundreds or a few thousand
years. These communities did not exist as entities in
the past, and there is no reason to expect their
continued coexistence in the future. Because of the
highly individualistic response of species to climate
and because this response depends on interactions
with other species in the community, it is difficult to
determine which aspects of climate are particularly
Plant Ecology and Climate Change 279

critical to the distribution of a species, making


predictions of future distribution difficult.
(ii) Changes in the distribution of a species may lag
significantly behind climatic changes where dispersal
limits the rate of species migration. Thus, predictions
of the future response to climate requires some
understanding of factors governing the regeneration
phase.
till) Projections of the future response of vegetation to
climate are quite sensitive to availability of soil
resources, so that knowledge of climate alone is
inadequate to predict future species distribution.
(iv) Finally, in complex and more diverse communities
such as dry or wet tropical forest there are so many
important species that we can never hope to
understand their climatic controls well enough to
predict future vegetation changes.
What is needed is a simplified classification of species into
functional groups whose environmental controls can be
predicted from general ecological principles. The question
is whether there are any predictable responses of groups
of species that make a functionalgroup classification
practical.

Plants' Interaction with Environment


As plants interact with the environment directly by
exchanging water and energy, they are very sensitive to
storms, droughts and floods . These weather events can
severely damage crop yield. In this unit we look at how
plants are affected by changes in temperature, humidity
and rainfall.
Probably the most important role of plants in the
environment is the production of oxygen (02) and the
absorption of carbon dioxide (C02) from the atmosphere
during the process of photosynthesis. Photosynthesis is the
280 Plant Ecology

basic process for plant life. Plants also affect and change
their surroundings to make them more suitable for growth.
Figure 2 shows some of the more minor ways in which
plants interact with their environment. Environmental
conditions, such as light intensity, temperature, water
availability and wind strength, affect plant growth.

Leavtllt ~ bnlnd\_
absorb sound ilnd block Branches, leaves
erosion-causing rainfall provide shade and reduoe
windspMd

EvapotraMpiration
from Il!!iIves cools
surrounding ilir

Rooa
~Iizesoil.
prevent erosion

Figure 2. Plant's interact with environment

Plants also modify the environment around them, they


release water which cools the air, breakdown the soil to
make it suitable for their roots and for other plants and
animals and decrease the speed of the wind. As you know,
plants need water to live and grow. High temperatures
reduce the availability of water and decrease crop yields.
Young plants are especially vulnerable to extreme weather
conditions.
Plant Ecology and Climate Change 281

Let's now look at some of the mechanisms plants use


to cope with extreme weather events such as high
temperatures, floods or droughts.

High Temperatures
High temperatures affect crops directly by increasing the
rate at which they loose water (their evaporation rate), in
the same way that high temperatures make us sweat.
Plants have very small pores, called stomata, spread over
their leaves. These stomata help the plants control how
much water they contain. In figure below you can see that
the stomata are made up of two guard cells which open
or close the pore depending on how much water the plant
needs. During dry periods, the guard cells are closed so
that the plants do not loose too much water. Under normal .
weather conditions, the stomata are open.

Gucrd cells

Stana

Stanatal q:>enlng

Figure 3. Working of stomata

Each plant type has different structural characteristics and


so all plants don't grow well at the same temperature.
When the optimal temperature value for a particular plant
is exceeded, the plant tends not to grow as well leading
to a drop in yield. Most plants are very sensitive to high
temperatures, although the extent varies depending on the
age of the plant and its ability to withstand poor situations.
282 Plant Ecology

High temperatures also increase the amount of water lost


from the soil by evaporation. This also affects plant growth
as the soil is their main source of water.

Precipitation
Precipitation (rainfall) is the primary source of soil
moisture, and rainfall amount is probably the most
important factor determining the crop yield. A change in
climate can cause an increase or a decrease in the amount
of precipitation which falls.

primary root -.....,MIi'--


~ ._ _~ root hairs

__ root tip
rootcap---

Figure 4. Plant roots

Roots are one of the main ways plants get water from the
environment. In many parts of the world, plants have
much longer roots than trunks or branches. Sometimes a
bush that is only around 30 cm high can have roots that
go as deep as two meters into the soil. This happens in
places where there is not much rain during the year, like
the deserts or the very arid dry regions of the world.
Dry periods can badly affect plant growth but the
amount of damage depends on the ability of the plant to
expand its root system and how much water the soil can
hold onto. High humidity, frost and hail may also damage
certain crops.
Plant Ecology and Climate Change 283

High temperatures are normally accompanied by dry


periods and both the warmth and the lack of water can
have negative effects on plant growth. Roots are unable
to find water in the soil and stomata have to close to
prevent water loss from the plant. This causes the
temperature of the plant to rise and this can damage the
plant. When plants struggle to grow because of high
temperatures or lack of water, they are said to be under
stress.
Excessively wet years, on the other hand, may cause
crop yields to fall. The soil becomes waterlogged and the
plant roots rot in the excess water. Intense bursts of rainfall
may also damage young plants, both because of the hard
impact of the rain drops on the tender plants and because
heavy rain can cause soil erosion.
VEGETATION AND NATURAL RESOURCES

Given the difficulties of predicting direct species reponses


to climate and the importance of soil resources in
mediating plant responses to climate, the correlation of
vegetation with soil resources may provide an effective
means of predicting the response of functional groups of
plants to climate change (figure 5).
Soil resource availability will respond in predictable
ways to altered climate. In cold climates, increased
temperature will stimulate microbial activity and nutrient
cycling, thus increasing the availability of commonly
limiting nutrients. such as nitrogen and phosphorus.
In temperate mesic-to-dry climates, increased
temperature and potential evapotranspiration combined
with moderate changes in precipitation will decrease soil
moisture in mid-continental regions. In tropical coastal dry
climates, moderate increases in temperature with high
variations in precipitation are expected.
284 Plant Ecology

Figure 5. Interrelationship between climate, soil resources, and functional


. groups of organisms.

'000

eoo
...E
~

E eoo
j
I 400

i aoo

0
lV1S 1NO

V.ar
1tH 1.
Figure 6. Total annual precipitation of the last fifteen years at the tropical
deciduous forest ill the Biological Reserve of Chamela, Jalisco, Mexico.

Rates of nutrient cycling correlate closely with availability


of most soil resources. Forest clearing for agriculture
creates earlv successional habitat with associated increases
J

in light availability. Local and regional variations in


climatic projections and soil fertility provide logical bases
for refining these predictions< The major point is that
Plant Ecology and Climate Change 285

patterns of change in soil resources can be predicted from


defined scenarios of climatic change. Soil moisture
responds readily and predictably to changes in climate.
However, considerable research is necessary to
determine to the magnitude and timing of the response of
soil fertility to changes in climate. At present it is also
difficult to predict how changes in the seasonality of
climate affect soil resources and the plant phenological
patterns necessary to exploit these soil resources
effectively.
Changes in availability of soil resources lead to
predictable changes in the types of plants that can be
expected. In brief, high-resource environments support a
high relative growth rate (RGR) through high capacities
for photosynthesis and nutrient uptake, which in turn
require high tissue-nitrogen concentrations. Continued
high rates of resource capture require high rates of root
and leaf turnover, a process that can be supported at
relatively low cost in a high-resource environment.
Conversely, in low-resource environments there are
inadequate resources to support rapid growth, so plants
are constrained to grow slowly and are likely to achieve
a small size. Because of low tissue-nutrient concentrations,
plants in low-resource environments have low potentials
to photosynthesize, transpire, and absorb nutrients. Plants
in these environments have high root-to-shoot ratios to
maximise capture of scarce soil resources. Tissue-turnover
rates are low, causing tissue nutrients to be retained for ---
a long time but also requiring effective chemical defense
against herbivores and pathogens. These chemical
defenses reduce litter quality and reinforce the low
nutrient availability in these sites.
Experimental evidence supporting these predictions
has emerged mainly from studies in temperate ecosystems,
although recent studies in tropical rain forests suggest
similar patterns with respect to relative growth rate, root-
286 Plant Ecology

shoot ratios and mineral nutrient requirements. Less is


known about the highly diverse tropical deciduous forest
in which the availability of soil resources is restricted by
seasonal drought. In these forests tree-seedlings occupying
resource-rich disturbed areas (e.g., Heliocarpus pallidus)
show high relative growth rates, high demands for
mineral nutrients and a low root/shoot ratio. These
species also tend to be more sensitive to water and mineral
nutrient stress.
ENVIRONMENTAL RESPONSES OF PLANTS

Differences in resource requirements determine plant


response to environmental fluctuations. Plants adapted to
high availability of soil resources are sensitive to changes
in nutrient supply and other resources such as light and
water. In these habitats plants respond plastically to
localised depletion zones around root systems through
morphological changes in roots, res:ulting in reallocation
of absorptive surfaces from depleted zones into the
resource-rich areas. This system of patch exploitation
requires shoots and roots with a short life span and high
rate of tissue turnover.
A high degree of morphological plasticity associated
with active foraging will be of selective advantage only
where it promotes access to large reserves of light, water,
and mineral nutrients. Conversely, in habitats where
productivity is chronically resource-limited, there is less
morphological plasticity. On infertile soils we expect that
resource capture and survival will depend on successful
exploitation of resource pulses. In unproductive habitats,
in which growth is frequently uncoupled from r~source
capture, plasticity would involve reversible physiological
changes rather than reallocation of biomass to facilitate
exploitation of resource pulses. These generalisations come
from screening programs which document the responses
of plants to fluctuations in resource supply.
Plant Ecology and Climate Change 287

The general predictions of the laboratory experiments


described above are consistent with responses of
communities to field manipulations. In artic tundra, when
light temperature, and nutrients were manipulated to
simulate patterns expected with climatic change, each
species initially showed a species-specific, unique response
to our manipulations, which was difficult to predict from
general principles. However, after nine years of treatment,
there continued to be species-specific, unpredictable
responses to temperature, but relatively predictabl
responses to nutrients: For example Betula nana and other
rapidly growing deciduous shrubs increased growth in
response to nutrient addition, whereas Ledum palustre
and other slowly growing evergreen species responded
negatively to nutrient addition.
Similarly, all species except the most shade-tolerant
understory species responded negatively to reduction in
light intensity. These results suggest several important
conclusions. First, the resource responses observed in the
field were consistent with predictions of laboratory
experiments in that rapidly growing species responded
positively to nutrient addition but slowly growing species
responded negatively to this improvement in resource
supply, presumably through changes in competitive
balance. Secondly, it was easier to predict responses to
altered resource supply than to altered climate.
The relatively minor increase in nutrient availability
that occurred nine years after initiation of our temperature
treatment suggests that our experiment was too short for
climate to strongly alter soil resource supply. A third result
of these experiments was that manipulations which
benefited some species reduced the biomass of other
species. Therefore, the overall production of the ecosystem
was affected much less than the productivity of individual
species or functional groups. Thus, the contrasting
responses of individual species and functional groups
288 Plant Ecology

buffer ecosystem processes such as production and


nutrient cycling, rendering them relatively resistant to
environmental change.
The buffered response of ecosystem processes is also
seen in an unmanipulated tundra ecosystem sampled over
a period- of years. Over a series of years in which the
productivity of individual species varied 2-8-fold, there
was no significant variation in productivity of the total
community.
Addition of water and nutrients to grassland also gives
predictable responses by functional groups of species.
Rapidly growing forbs and grasses responded positively to
nutrient and water addition, whereas the more slowly
growing cactus responded negatively. Moreover, when
grassland production was examined over a series of years,
the productivity of individual species varied much less
than did the productivity of the community as a whole,
again indicating the extent to which ecosystem processes
are buffered by the compensatory responses of individual
species.
ADAPTIVE STRATEGIES OF VEGETATION

Plants evolve a variety of adaptations to the light and


moisture availability within a particular environment in
order to flourish. Plants adaptations include those of leaf
form and canopy structure (the roof of foliage formed by
the crowns of trees). For instance, a hard, needle leaf
structure is an adaptation to extreme temperatures and
low moisture status in winter. The leaves of some rain
forest trees have a special joint .at the bottom of their stalk
that enables them to twist and turn to follow the light as
the sun passes from east to west over head. Deciduous
trees drop their leaves to cut transpiration loss during dry
periods and when temperatures are very cold.
Fleshy "leaves", like those of desert succulents or thick
photosynthetic skin like that of the giant Saguaro cactus
Plant Ecology and Climate Change 289

helps retain moisture. The Baobab tree, found in the wet/


dry tropical (savanna) climate stores water in its trunk to
combat the long drought period experienced in that
climate.
Plants have adapted particular root structures to live
in arid regions. Deep tap roots draw moisture hidden deep
below the surface while extensive near - surface root
systems catch moisture as it infiltrates into soil. Some
desert grasses have rolled surfaces to reduce water loss
from the inner surface and hairs which reduce air
movement.
Canopy structures reflect the environmental
conditions vegetation grows in. The conical canopies of
conifers help shed snow and catch low angle sun rays
during the long winters where they grow. The rain forest
displays a multi-layered canopy. Each layer possesses
organisms adapted to the environmental conditions found
in it. A canopy can be so thick and dense, like that found
in the rain forest, that little light penetrates to the surface.
The lack of light for understory growth creates an open
forest that you can see into for some distance. Where
canopy density is low, more light filters to the surface
creating a thick ground cover and a closed forest.
Rarely is any location dominated by a single specie of
plant. A plant community refers to the associated plant
species that form the natural vegetation of any place. For
instance, a midlatitude forest is comprised of a community
of trees, shrubs, ferns, grasses, and flowering herbs. Plant
communities provide a habitat for animals and
significantly modify the local environment. Plant
communities affect soil type when organic material
decomposes into the soil altering soil moisture retention,
infiltration capacity, soil structure and soil chemiStry.
Trees shade the forest floor, reducing incident' solar
radiation and lowering temperatures of both the soil, and
the air. Reduced incident light decreases evaporation
290 Plant Ecology

keeping soils moister beneath the forest canopy. These


impacts affect animal habitats and the diversity of animal
species which are associated with these plant
communities.
An ecotone is a plant community in a distinct zone of
transition between other more extensive communities.
Ecotones vary in scale, from local (between forest and
field) to global (savannas). Within an ecotone plants of
different environmental tolerances often intermingle. For
instance, grasses adapted to low moisture conditions
intermingle with deciduous trees within a prairie - forest
ecotone.

Principle of Limiting Factors


The plants and animals that succeed in occupying a
particular niche are those that can easily adapt to the
unique environmental conditions of a site. Each plant and
animal in the community has a specific range of tolerance
for particular environmental conditions. Climate factors
are the most important influence over the successful
establishment of plant and animal communities. Two
climatic factors are important, sunlight and moisture .
Not •only is the amount of sunlight available important
but the duration and quality of light are important too. For
instance, at high altitudes the intense ultra violet light may
inhibit the growth of particular plants. The intensity of
light affects photosynthesis and rate of primary
productivity. The duration of sunlight affects the
flowering of plants and the activity patterns of animals.
The availability of water is important for the survival of
most life forms. But plants require water for a number of
life processes like germination, growth and reproduction
too.
Principle" of limiting factors says that the maximum
obtainable rate of photosynthesis is limited by whichever
Plant Ecology and Climate Change 291

basic resource of plant growth is in least supply. The


availability of energy and moisture varies geographically.
At high latitudes the limiting factor is generally energy
availability while in low latitudes moisture is the limiting
factor to growth. The figure below shows the relationship
between potential evapotranspiration (PE), a moisture
index (MI), climate and vegetation.

p Cool Dry Cool Wet

p Moisture Index

Figure 7. Relationship between c1iTrUlte, vegetation potential


evapotranspiration and the moisture index.

Potential evapotranspiration is the optimal amount of


water entering the atmosphere as a result of evaporation
and plant transpiration when there is an unlimited
amount of moisture. Because evaporation and
transpiration depend on energy availability, potential
evapotranspiration is a measure of energy input. High-
values of potential evapotranspiration relate to warm
climates while low values to cool climates. The moisture
index is a measure of moisture availability. High values of
the moisture index means that plenty of water is available.
292 Plant Ecology

Combining the two variables, potential evapo-


transpiration and moisture index we have a notion of
what the climate is like in any part of the diagrams. For
instance, high PE and large values of MI are indicative of
warm and moist climates. Note that tundra and taiga
(mostly conifers) are successfully established over a wide
range of moisture conditions, from dry to moist, but
always in cool environments. Other vegetation systems
have more narrowly defined moisture and (emperature
requirements.
Plants of a particular region have adapted to the
temperature and moisture conditions in which they live.
Most gatd~.ners are familiar with plant hardiness
(growing) zone maps. The zones are based on the
minimum temperature experienced and thus tolerated by
different species of plants. There have been recent signs
that these zones are starting to shift due to global
warming.
PLANT SUCCESSION

Natural vegetation of a particular location evolves in a


sequenCe of steps involving different plant communities.
The evolutionary process is known as plant succession.
Plant succession usually begins with a fairly simple
community known as a pioneer community. The pioneer
community, and each successive community alters the
environment in such a way to permit new communities
to occupy a site. These alterations of the environment
include changes in site microclimate and soil conditions.
A climax community is the result of a long period of plant
succession. Climax communities usually exhibit a good
deal of species diversity and thus are relatively stable
systems. Disturbance renews a successional sequence.
Plant succession was renewed after the explosion of Mt.
St. Helens with the subsequent disruption of biotic
communities that inhabited the region. Human
Plant Ecology and Climate Change 293

disturbance related to tropical deforestation has renewed


the successional sequence of plant communities in the
tropical rain. forest.
RESEARCH ON VEGETATION CHANGE

Most researchers who study vegetation change do so


within a narrow conceptual framework. They do not
regularly make use of the findings and insights of very
similar studies being conducted in other research
subdisciplines, nor do they try to make their findings and
insights easily accessible to researchers in other areas. It
seems obvious that plant ecology would benefit from better
communication among the different research speciality
areas. Indeed, communication across broad disciplinary
horizons within ecology may be of value more generally.
We propose three steps that individual researchers can
take to increase the useful exchange of ideas and
informa tion among these research areas. The fourth step
should be undertaken by the scientific community as a
whole.
Invasion ecology, succession ecology, gap/patch
dynamics, and studies of the effects of global change on
plant communities all study vegetation in flux, that is,
vegetation experiencing changes in species composition.
Thus, each speciality area could be considered a part of
a larger research initiative: the ecology of vegetation
change. There is precedence for this perspective. Luken
recognised vegetation change as the fundamental subject
area relevant to all kinds of vegetation management. Thus,
the first step is to be aware that related speciality areas
exist. This step may be especially important for young
researchers, such as doctoral students, who may have
limited awareness of the scope of their and related
research areas. If researchers began to envision themselves
as studying vegetation change, rather than an invasion
biologist, or succession ecologist, they would be less
294 Plant Ecology

inclined to take a parochial perspective with respect to


their research.
The simplest step individual researchers can take to
increase communication among speciality areas is to
consciously seekout relevant ideas and data from related
research areas. For example, an invasion ecologist
investigating the community-wide consequences of an
introduced species that is altering the soil-microbial
community could recognise the value of seeking out
findings obtained from similar studies conducted within
a succession framework. Or, a researcher studying the
effects of increases in nitrogen deposition on plant
communities could recognise the relevance of the many
invasion studies that have examined the impact of
changing resource levels on plant community structure.
A simple conceptual approach that could integrate
succession and invasion ecology is to consider the various
ways that species can facilitate and/or inhibit one
another. Considering native and introduced species, and
the possibility that each may inhibit or facilitate species in
either group, eight types of potential interactions among
the two groups of species can be identified.
Studies of invasibility have shown that some native
species are able to prevent, or at least slow, the
establishment of introduced species. For example, Wedin
and Tilman showed that the native grass Schizachyrium
scoparium can inhibit the establishment and spread of the
introduced grass Agropyron repens by its ability to depress
levels of soil nitrogen.
Studies in disturbed New Zealand forests have shown
that Hakea sericea is inhibiting the reestablishment of the
native shrub Leptospermum scoparium and native tree
Kunzea ericoides. The latter two examples exemplify the
fourth type of interaction. In a Hawaiian study, Carino
and Daehler showed that an introduced legume,
Chamaecrista nicitans, facilitated the subsequent invasion
Plant Ecology and Climate Change 295

of another introduced species, the grass Pennisetum


setaceum.
The final three types of interactions are not typical
subjects of either succession or invasion ecology, yet they
do occur. Juniperus virginiana, a native tree in the US, has
been found to facilitate the establishment of Rhamnus
cathartica, an introduced and invasive tree, on some
Mississippi River bluffs through a nurse plant effect on
Rhamnus seedlings.
/ In the western United States, crested wheatgrass
(Agropyrum cristatum),- an introduced perennial, is
known to impede the spread and establishment of the
annual cheatgrass (Bromus tectorum), another introduced
species. And, De Pietri showed that Rosa rubiginosa, a
shrub introduced to Argentina, facilitated the re-
establishment of several native woody species in disturbed
subantarctic forests by reducing grazing herbivory on
native seedlings growing beneath the thorny shrubs. This
brief accounting of eight types of interaction between
introduced and native species emph~sizes that it is the
nature of the impact of the species that is important, not
the place of origin.
Regardless of the interaction type under investigation,
all these studies are trying to answer the same two basic
questions: what are the mechanisms that facilitate or
inhibit the establishment and spread of particular plant
species over time, and what are the consequences of these
changes on community structure and ecosystem
processes? Rather than the subdisciplines of succession
and invasion ecology continuing down two parallel tracks,
we think it makes more sense to -take a common and
integrated approach.
REFERENCES

Bartholomew, G.A., "The role of natural history in contemporary


biology". Bioscience 36, 324-329. 1986.
296 Plant Ecology

Christian, J.M., Wilson, S.D., "Long-term ecosystem impacts of an


introduced grass in the northern Great Plains". Ecology 80,2397-
2407.1999.
Connell, J.H., Slatyer, R.O., "Mechanisms of succession in natural
communities and their role in community stability and
organization". Am. Nat. 111, 1119-1144. 1977.
Davis, M.A., Thompson, K., Grime, J.P., Charles 5, "Elton and the
dissociation of invasion ecology from the rest of ecology".
Diversity Distrib. 7, 97-102. 2001.
Glenn-Lewin, D.C., Peet, R.K., Veblen, T.T. (eds.), Plant Succession:
Theory and Prediction. Chapman & Hall, New York. 1992.
Bibliography

Abedon, S. T. 'Phage Ecology", In R. Calendar and S. T. Abedon (eds.),


The Bacteriophages. Oxford University Press, Oxford. 2006.
Ackermann, H.-W., and M. S. DuBow. VirusesofProkllryotes, CRC Press,
Boca Raton, Florida. 1987.
Agrios, George. Plant Pathology. Academic Press. 2005.
Alexopoulos, c.J., Charles W. Mims, M. Blackwell et al., Introductory
Mycology, 4th ed. John Wiley and Sons, Hoboken NJ, 2004.
Archibold, O. W. Ecology of World Vegetation. New York: Springer
Publishing, 1994.
Arneson, PA "Plant disease epidemiology: temporal aspects". Plant
. Health Instructor. American Phytopathological Society. 200t.
Arora, David. (1986). "~ushrooms Demystified: A Comprehensive
Guide to the Fleshy Fungi". 2nd ed. Ten Speed Press.
Barbour, M.G, J.H. Burk, and W.D. Pitts. "Terrestrial Plant Ecology".
Menlo Park: Benjamin Cummings, 1987.
Barea JM, Pozo MJ, Azc6n R, Azc6n-Aguilar C. "Microbial co-operation
in the rhizosphere". ,. Exp. Bot. 56: 1761-1778. 2005.
Bartholomew, G.A., "The role of natural history in contemporary
biology". Bioscience 36, 324-329. 1986.
Baskin, Yvonne. A Plague of Rilts and Rubberuines: The Growing Threat Of
Species Invasions. Island Press, 2003.
Bouma J., R.B. Brown, and P.S.C. Rao. 1982. "Basics of Soil-Water
Relationships -Part I. Soil as a Porous Medium." Soil Science Fact
Sheet SL-37. Florida Cooperative Extension Service. IF AS.
Gainesville, FL.
Breckle, S-W. Walter's Vegetation of the Earth. New York: Springer
Publishing, 2002.
Burdick, Alan. Out of Eden: An Odyssey of Ecological Invasion. Farrar Straus
and Giroux, 2005:
Burrows, C. J. Processes of Vegetsltion Change. Oxford: Routledge Press,
1990.
298 Plant Ecology

Chibani-Chennoufi, S., A. Bruttin, M. L. Dillmann, and H. Briissow.


"Phage-host interaction: an ecological perspective" J. Bacteriol.
186:3677-3686. 2004.
Christian, J.M., Wilson, S.D., "Long-term ecosystem impacts of an
introduced grass in the northern Great Plains". Ecology 80,2397-
2407.1999.
Coates, Peter. American Perceptions of Immigrant and Invasive Species·
Strangers on the Land. University of California Press, 2007.
Connell, J.H., Slatyer, R.O., "Mechanisms of succession in natural
communities and their role in community stability and
organization". Am. Nat. 111, 1119-1144. 1977.
Cronquist, Arthur. An Integrated System of Classification of Flowering
Plants. Columbia Univ. Press, New York. 1981.
Davis, M.A., Thompson, K., Grime, J.P., Charles S. "Elton and the
dissociation of invasion ecology from the rest of ecology".
Diversity Distrib. 7,97-102. 200l.
de ViIliers, M. Windswept: the story of wind and weather. Toronto:
McClelland and Stewart. 2006.
Deacon JW. "Fungal Biology" Malden, MA: Blackwell Publishers. 2005.
Deshpande MY. "Mycopesticide production by fermentation: potential
and challenges.". Crit Rev Microbiol. 25: 229-243. 1999.
Dickenson, C.H. and Pugh, G.J.F. (editors) Biology of Plant Litter
Decomposition. London and New York: Academic Press. 1974.
Dickinson, M . Molecular Plant Pathology. BIOS Scientific Publishers.
2903.
Drenth, A "Fungal epidemics - does spatial structure matter?".. New
Phytologist 163: 4-7. 2004.
Elton, Charles S. The Ecology of Invasions by Animals and Plants. University
of Chicago Press, 2000.
Evans, L. T. (1998). Feeding the Ten Billion - Plants and Population Growth.
Cambridge University Press. Paperback, 247 pages.
Faergi, K. and van der Pijl, L. The Principles of Pollination Ecology. Oxford,
UK: Pergamon Press Ltd. 1966.
Feldmeyer-Christie, (et.al). Modern Approaches In Vegetation Monitoring.
Budapest: Akademiai Kiado, 2005.
George N. Agrios. Plant Pathology, Academic Press. New York. 1997.
Gifford, Ernest M., Adriance S. Foster. Morphology and Evolution of
Vascular Plants. Third edition. Wf{:~ ilnd Company, New
York. 1989. .'
Gleason, H.A. 1926. The individualistic concept of the plant association.
Bulletin of the Torrey Botanical Club, 53:1-20.
Bibliography 299

Glenn-Lewin, D.C., Peet, R.K., Veblen, T.T. (eds.), Plant Succession: Theory
and Prediction. Chapman & Hall, New York. 1992.
Glime, Janice M., Bryophyte Ecology, Volume 1. Physiological Ecology.
Ebook sponsored by Michigan Technological University and the
International Association of Bryologists. 2007.
Goyal, S. M., C. P. Gerba, and G. Bitton. Phage Ecology. CRC Press, Boca
Raton, Florida. 1987.
Gray, J. "The Microfossil Record of Early Land Plants: Advances in
Understanding of Early Terrestrialization, 1970-1984".
Philosophical Transactions of the Royal Society of London. Series
B, Biological Sciences (1934-1990) 309 (1138): 167-195. 1985.
Grime, J.P. 1987. Plant strategies and vegetation processes. Wiley
Interscience, New York NY.
Janick, Jules. Horticultural Science. Sar;t Francisco: W.H. Freeman, 1979.
Kabat, P., et al. (eds). Vegetation, Water, Humans and the Climate: A New
Perspective on an Interactive System. Heidelberg: Springer-Verlag
2004.
Kenrick, Paul & Crane, Peter R. The Origin and Early Diversification of Land
Plants: A Cladistic Study. Washington, D. c.: Smithsonian
Institution Press. 1997.
Lindeman, R.L. 1942. 'The trophic-dynamic aspect of ecology". Ecology
23: 399-418.
Lockwood, Julie; Martha Hoopes, Michael Marchetti Invasion Ecology.
Blackwell Publishing, 2006.
Lord, Thomas R. Ferns and Fern Allies of Pennsylvania. Indiana, PA:
Pinelands Press. 2006.
Macarthur, R.H. and E.O. Wilson. The theory of Island Biogeography.
Princeton: Princeton University Press. 1967
Madden, Laurence; Gareth Hughes, Frank Van Den Bosch. Study of Plant
Disease Epidemics. American Phytopathological Society. 2007.
McNeeley, Jeffrey A. The Great Reshuffling: Human Dimensions Of Invasive
Alien Species. World Conservation Union (IUCN), 2001.
Merva G.E. Physio~ngineering principles. 1975. The AVI Publishing
Company, Inc. Westport, CT.
Milton Zaitlin. Discoveries in Plant Biology, New York 14853, USA. Pp.:
105-110. 1998.
Moran, Robbin C. A Natural History of Ferns. Portland, OR: Timber Press.
2004.
Mueller-Dombois, D., and H. Ellenberg. Aims and Methods of Vegetation
Ecology. The Blackburn Press, 2003.
Patten, B.C. 1959. "An Introduction to the Cybernetics of the Ecosystem:
The Trophic-Dynamic Aspect". Ecology 40, no. 2.: 221-231.
300 Plant Ecology

Paul, J. H., and C. A. Kellogg. "Ecology of bacteriophages in nature",


p. 211-246. In C. J. Hurst (ed.), Viml Ecology. Academic Press, San
Diego. 2000.
Perotto S, Bonfante P. "Bacterial associations with mycorrhizal fungi:
close and distant friends in the rhizosphere.". Trends Microbiol. 5:
496-501. 1997
Phillipson, J. Ecological Energenics. London and Beccles. William Clowes
and Sons. 1966.
Pielou, E.C. The Energy of Nature. Chicago: The University Of Chicago
Press. 2001.
Raven, J.A.; Edwards, D."Roots: evolutionary origins and
biogeochemical significance" . Journal of Experimental Botany 52
(90001): 381-401. 2001.
Raven, Peter H., Evert, Ray F., & Eichhorn, Susan E. Biology of Plants (7th
ed.). New York: W. H. Freeman and Company. 2005.
Ridley, H.N. The dispersal of plants throughout the world. Ashford (UK): L.
Reeve. 1930.
Robert Ulanowicz (1997). Ecology, the Ascendant Perspective. Columbia
Univ. Press.
Schofield, W. B. Introduction to Bryology. New York: Macmillan. 1985.
Shultz, T.R; Line, R.F "High-Temperature, Adult-Plant Resistance to
Wheat Stripe Rust and Effects on Yield Components". Agronomy
Journal 84: 170-175. American Society of Agronomy. E92.
Simpson, M.G. Plant Systematics. Elsevier Academic Press. 2006.
Stewart W. N. & Rothwell G. W. Paleobotany and the Evolution of Plants.
Cambridge Univ. Press, NY, USA. 521pp.
Tansley, A.G. 1935. "The use and abuse of vegetational concepts and
terms". Ecology 16: 284-307.
Taylor, Thomas N. & Taylor, Edith L. The Biology and Evolution of Fossil
Plants. Englewood Cliffs, NJ: Prentice Hall. 1993.
Terrill, Ceiridwen. Unnatural Landscapes: Tracking Invasive Species.
University of Arizona Press, 2007.
Thomas B. and Vince-Prue D., Photoperiodism in plants, Academic Press,
1997.
Van Der Maarel, E. Vegetation Ecology. Oxford: Blackwell Publishers,
2004.
Van Driesche, Jason; Roy Van Driesche Nature Out of Place: Biological
Invasions In The Global Age. Island Press, 2004.
Vankat, J. L. The Natural Vegetation of North America. Krieger Publishing
Co., 1992.
Watson, E. V. The Structure and Life of Bryophytes. London: Hutchinson
University Library. 1971.
Index

Alluvial soils 44 Fine-textured soils 40


Anchoring rhizoids 134 Fossillartd plants 5
Anti-bacterial chemotherapy Free-living organism 146
109
Aquatic environments 111 Gametophyte 126
Aquatic organisms 186 Glyceraldehyde 3-Phosphate
(G3P) 62
Bacterial microcolonies 241 Gravitational water 53
~cterial phenotypes 243
Biogeochemical cycles 9 Heterotrophs 58
Bisexual gametophytes 127
Bryophytes 130 Invasion biology 275

Chemosynthetic organisms 58 Land plants. 1


Coarse-textured soils 39 Latitudinal diversity gradient
Continental drift theory 7 190
Crassulacean Acid Metabolism Late Embryogenesis Abundant
(CAM) 63 (LEA) 9
Lepidodendron 144
Devonian plant 138 Life zones 189
Dikaryotic mycelium 116 Linnaeus' system 1
Liverwort sporophytes 134
Ecosystem ecology 229 Low-resource environments
Environmental microbiology 285
230 Lypopolysaccharide (LPS) 245
Epiphytes 181
Ethnomycology 109 Medium-textured soils 39
Evapotranspiration 47 Moisture Index (Ml) 291
Molecular model systems 232
Field Capacity (FC) 53 Motile sperm 131
302 Plant Ecology

Multicellular eukaryotes 108 Phototropism 55


Mycorrhizal networks 110 Physiological ecology 231
PhysiolOgical mechanisms 117
Net Primary Productivity Plant life cycle 126
(NPP) 196 Plant nutrient ions 39
Non-photosynthetic ancestors Plantae sensu lato 2
2 Plant·pathogenic fungi 111
Northern hemisphere Polypodiopsida 145
peatlands 135 Potential Evapotranspiration
Nurse logs 74 (PE) 291
Nutrient storage capacity 38 Psychotropic compounds 108
Purple nonsulfur bacteria 60
Oxygenic photosynthesis 61
Relative Growth Rate (RGR)
Permanent Wilting Point 285
(PWP) 53 Relative Humidity (RH) 68
Permian-triassic boundary 120
Phage community ecology 233 Secondaly minerals 39
Phage community-ecology Soil-plant-atmbSphere con-
theory 245 tinuum 49
Phage ecology 229 Sporophyte growth 133
Phage ecosystem ecology 234 Successional communities 192
Phage organismal ecology 238
Phage population ecology 233 Taiga forests 185
Phage-ecological interactions Temperate grasslands 183
227 Temperate-phage mutation
Phage-resistant bacterial 237
phenotype 244 Terrestrial ecosystems 109
Phosphoenolpyruvate (PEP) Triangular cross-section 73
63
Phosphoglyceraldehyde Vascular plant 137
(PGAL) 62 Vrridiplantae 1
Photosynthetic plants 4
Photosynthetic systems 60 Woody plants 74

Vous aimerez peut-être aussi