Download Chapter 4

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Unified neutral theory of biodiversity wikipedia , lookup

Ecological fitting wikipedia , lookup

Biodiversity action plan wikipedia , lookup

Habitat conservation wikipedia , lookup

Occupancy–abundance relationship wikipedia , lookup

Introduced species wikipedia , lookup

Habitat wikipedia , lookup

Island restoration wikipedia , lookup

Latitudinal gradients in species diversity wikipedia , lookup

Storage effect wikipedia , lookup

Bifrenaria wikipedia , lookup

Transcript
1
The nature of the plant community: a reductionist view
2
3
J. Bastow Wilson
Botany Department, University of Otago, P.O. Box 56, Dunedin, New Zealand.
4
5
Andrew D.Q. Agnew
Institute of Biological Sciences, University of Wales Aberystwyth, SY23 3DA, U.K.
Chapter 4. Mechanisms of coexistence
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
1
2
3
4
5
6
7
8
9
10
11
12
13
Alpha-niche Differentiation .................................................................................................................. 5
1.1 Resources (type of resource and time of availability) ........................................................... 5
1.2 Heterotroph-imposed niches ................................................................................................. 7
1.3 The niche extended by reaction ............................................................................................. 7
Environmental Fluctuation (seasonal, annual and decadal change) ..................................................... 9
Pest Pressure (heterotroph challenges) ............................................................................................... 13
3.1 Pathogens ............................................................................................................................ 14
3.2 Herbivory, general ............................................................................................................... 15
3.3 Herbivory of disseminules and seedlings ............................................................................ 16
3.4 Vegetative herbivory ........................................................................................................... 17
3.5 Pest Pressure conclusions .................................................................................................... 17
Circular Interference Networks .......................................................................................................... 18
Allogenic Disturbance (disrupting growth, mainly mechanically) ..................................................... 24
Interference/dispersal Tradeoff ........................................................................................................... 26
Initial Patch Composition ................................................................................................................... 27
Cyclic Succession: movement of community phases ......................................................................... 27
Equal Chance: neutrality..................................................................................................................... 27
Inertia .................................................................................................................................................. 29
10.1 Temporal Inertia .................................................................................................................. 29
10.2 Spatial Inertia: aggregation ................................................................................................. 30
Coevolution of Similar Interference Ability ....................................................................................... 31
Spatial Mass Effect (vicinism) ............................................................................................................ 31
Conclusion .......................................................................................................................................... 32
Were there no plant species coexistence, there would be no need for this book. However, most
32
plant communities comprise persisting populations of several species (This is a good ‘generalist’
33
statement. I think it should be emphasized more frequently throughout the text. If it wasn’t stated in an
34
earlier chapter, it should be reiterated.. . Populations may increase or decrease through neutral drift or
35
weather fluctuations and species can immigrate or disappear from the local community. However, long-
36
term studies such as the Park Grass experiment (Silvertown 1987) and Bibury (Dunnett et al. 1998)
37
show that the basic tendency is persistence, for example, outbreaks are often followed by a decrease
38
back to the original abundance. This coexistence is the fundamental statement to be made about plant
39
communities, and how it is achieved is the fundamental problem. (good inro to chapter…)
40
Resources (e.g. light and nutrients) are almost always limiting. Competition, and thus
41
interference between individuals and species, is demonstrable in all types of habitat, except immediately
42
after disturbance (chap. 6, sect. 9.3 below; Clements et al. 1929). Interference abilities can never be
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 2 of 34
43
exactly equal, so the result should be the exclusion by interference of all but one species (Gause 1934).
44
Therefore, the amazing thing is not that the species in plant communities show any particular patterns of
45
coexistence, but that they coexist at all. Hutchinson (1941; 1961) asked: "How [is it] possible for a
46
number of species to coexist in a relatively isotrophic or unstructured environment, all competing for the
47
same sorts of materials?". He called it the “Paradox of the Plankton”.” Plants are not plankton and don’t
48
experience an ‘unstructured environment’ so why shouldn’t we expect coexistence?.
49
Monospecific stands of vegetation do exist, i.e. with only one vascular plant species (Plate 4.1).
50
Table 4.1 lists those that we have seen ourselves. They are often at land/water ecotones, in wet places
51
and especially in open water or extreme saline environments. In arid countries, a monotonous vegetation
52
of one halophytic species can dominate the landscape (Zohary 1973). We could generalise that these are
53
habitats where only one species is capable of growth due to a harsh environment, or where the
54
exuberance of one species excludes others by interference, but in some cases it is hard to know whether
55
to credit the extreme habitat or the high interference, e.g. Phragmites communis reedswamps.
56
Table 4.1. Some examples of monospecific stands. We exclude monospecificity in a single stratum or
57
guild of vegetation, such as a tree species or understorey species.
Habitat
Arid Saline
Sand dunes
Marine
submerged
Freshwater
submerged
Freshwater
floating
Freshwater
edge
Tidal/brackish
edge
River edge
58
Climatic zone Exemplar taxa
Reference
Sub-tropical
Halocnemum strobilaceum
Zohary 1973
Temperate
Zygophyllum dumosum
Zohary 1973
Ammophila arenaria
Mediterranean Posidonia oceanica (also Den Den Hartog 1970
other marine Helobeae)
Temperate
Zostera marina
Den Den Hartog 1970
Temperate
Sagittaria sagittifolia
Pieterse and Murphy 1993
Tropical
Podostemon spp.
Meijer 1976
Temperate
Lemna minor
Scunthorpe 1967
Azolla filiculoides
Scunthorpe 1967
Tropical
Eichornia crassipes
Pieterse and Murphy 1993
Temperate
Typha spp
Weisner 1993
Cladium mariscus
Tansley 1939
Tropical
Cyperus papyrus
Lind and Morrison 1974
Temperate
Salicornia spp
Tansley 1939
Subtropical
Avicenna marina
Batanouny 1981
Tropical
Rhizophora mangle
Gilmore and Snedaker 1993
Tropical
Pandanus spp
van Steenis 1981
At the other extreme is vegetation with high species richness. Values depend on the size of the
59
sample units, the life form guild being considered (e.g. often trees alone are recorded in tropical forests)
60
and the recording convention used (rooted or shoot presence, perennially or seasonally visible).
61
Whatever the sampling regime, extraordinarily high diversities can exist. Tropical rain forest is always
62
quoted as an example. Valencia et al. (1994) found 473 species of tree (individuals >5 cm dbh) in 1 ha
63
of Ecuadorian tropical rain forest, while Richards (1996) tabulates other examples with over 100 species
64
in 1 ha in New World and Malesian (not African!) tropical forests, Naveh and Whittaker (1979)
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 3 of 34
65
recorded 179 vascular species in 0.1 ha of a dry shrub/grass community in Israel. Mean species richness
66
of 18.3 per 0.01 m2 has been found in bryophyte carpets in the per-humid West Cape in New Zealand
67
(Steel et al. 2004) and 12.2 species at that scale in limestone grassland on Oeland, Sweden (van der
68
Maarel and Sykes 1993). (for another example, look to the Pine Savannas of Coastal Carolina; high
69
richness on the square meter scale….)
70
Box 4.1: Mechanisms of coexistence.
71
72
73
74
Stabilising mechanisms
Niche-differentiation
1. Alpha-niche Differentiation (type of resource and time of availability)
2. Environmental Fluctuation – season, decadal and gradual change
75
76
77
Balances
3. Heterotroph challenges: Pest Pressure
4. Circular competitive networks
78
79
80
81
82
Escape through movement
5. Allogenic Disturbance – disrupting growth mainly mechanically
6. Competition/dispersal Tradeoff
7. Initial Patch Composition
8. Cyclic Succession: movement of community phases:
83
84
85
86
87
88
89
90
Equalising mechanisms
9. Equal Chance (neutrality)
10. Inertia
Temporal Inertia
Spatial Inertia: aggregation
11. Coevolution of Similar Interference Ability
12. Spatial Mass Effect (vicinism)
Questions about coexistence must be asked at a particular spatial scale. The rainforest tree
91
Swietenia mahogoni (mahogany) occurs in the tropics and Colobanthus quitensis occurs in Antarctica;
92
they cannot be said to coexist. They grow in quite different places and environmental conditions, i.e. in
93
different beta niches. Similarly, Salicornia spp. (glasswort) occur on low-altitude saltmarshes and
94
Androsace spp. are alpines, again they occupy different beta niches. The Paradox of the Plankton as
95
defined by Hutchinson refers to how coexistence can occur in a “relatively isotrophic or unstructured
96
environment”. This scale is difficult to define, because allogenic environmental heterogeneity occurs
97
down to the very finest scales, so all species in a mixture exist as a pattern of abundance. The species’
98
patterns will create further patchiness in resources since species differ in their resource economies: their
99
reaction on the environment. Numerous studies have shown that each species’ individuals affect its soils
100
(section 1.3), which can give autogenic heterogeneity. Also, each individual, by extending over space,
101
must sample a spectrum of resource and environmental qualities. Therefore, rather than specify a
102
particular scale, we state here that we are concerned with mechanisms that allow species to coexist
103
locally, i.e. mechanisms that are not due to imposed habitat heterogeneity within the area considered.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 4 of 34
104
J.B. Wilson (1990) identified 12 distinct mechanisms by which coexistence could be maintained:
105
(1) Alpha-niche Differentiation, (2) Environmental Fluctuation, (3) Pest Pressure, (4) Circular
106
Interference Networks, (5) Allogenic Disturbance, (6) Interference/dispersal Tradeoff, (7) Initial Patch
107
Composition, (8) Cyclic Succession, (9) Equal Chance, (10) Inertia (temporal and spatial),
108
(11) Coevolution of Similar Interference Ability and (12) Spatial Mass Effect. We believe them to be
109
distinct and we know of no new ones, though we have adopted a different arrangement(is the box your
110
arrangement or his?) that we hope brings some new insights.
111
There is a basic distinction between stabilising mechanisms, which contain an increase-when-
112
rare mechanism, and equalising mechanisms, which make the differences between species in
113
replacement rates smaller (Box 4.1; Chesson 2000). Stabilising mechanisms (1 to 8 in Box 4.1) are
114
driven by Alpha-niche Differentiation, Environmental Fluctuation, balances unrelated to niches or
115
escape through movement (Box 4.1). Species abundances are bound to fluctuate and stabilising
116
mechanisms must include negative abundance-dependence to counter this. [For animals, ‘density-
117
dependence’ is often used, but since the concepts of ‘individual’ and ‘density’ are difficult in most
118
plants (chap. 1, sect. 1.1) the more general ‘abundance-dependence’ should be used.] This means that
119
when a species is at lower biomass in the community its plants must have higher fitness in terms of
120
long-term RGR, and its biomass should increase (Chesson in press). Simply, the one necessary and
121
sufficient phenomenon for maintaining a species in a mixture is ‘increase-when-rare’1. The corollary of
122
this is population limitation, i.e. reduced fitness as biomass increases. Either way, a species’ fitness
123
should be inversely related to its abundance. In the short term and in species that reproduce only
124
vegetatively we should consider vegetative RGR (relative growth rate). In the longer term, population
125
growth is the critical question. The increase-when-rare feature is incorporated into the Lotka-Volterra
126
logistic function, as every schoolboy knows. A feature that interferes with this in the real world is the
127
Allee effect, whereby populations cannot recover from very low numbers due to low success in mating.
128
We believe that it is rare in plants because, as we outlined in chapter 1, each individual is effectively a
129
colony and perennials have an indefinite life span. Nevertheless there are examples of the effect in
130
Banksia goodii, a shrub of dry savannah (Lamont et al. 1993), and in the outpollinated annual Clarkia
131
concinna in California (Groom 1998). Species with obligate outcrossing and/or scattered distributions
132
and/or monocarpic reproduction and/or specialised pollen vectors would be more liable to it.
133
True coexistence must be through a stabilising mechanism. Equalising processes do not contain
134
an increase-when-rare mechanism, but are ways in which species may persist for a time in unstable local
135
coexistence, slowing exclusion by interference. Moreover, if the difference between the interference
136
abilities of species is large, even the presence of a stabilising mechanism may not prevent exclusion by
137
interference and in this situation an equalising mechanism might reduce the difference in interference
138
ability between two species so that the stabilising mechanism is able to cause coexistence (Chesson
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 5 of 34
139
2000). Ultimately every plant has established itself by a process that can be explained by its tolerances
140
(its niche) and the environmental conditions prevailing during its ecesis, but the reasons for a species'
141
presence in a particular spot are usually obscure. For example, a tree may persist for so long that the
142
local soil/geomorphological conditions that allowed it to establish have since changed. It is therefore
143
tempting to suggest that each individual's presence owes as much to chance as to ecological
144
differentiation and a theory of equal chance has been proposed to explain species mixtures, as we
145
discuss below. Again, species may persist temporarily through inertia of individuals or populations in
146
time or aggregation in space. Any process causing similarity in interference ability is also equalising.
147
Lastly, and rather hesitantly, we include the Spatial Mass Effect as an equalising mechanism.
148
1 Alpha-niche Differentiation
149
We discussed alpha niches in chapter 1 (sect. 4.1). It has been pointed out that coexistence by
150
Alpha-niche Differentiation is impossible to disprove. Each species must by definition occupy a
151
different niche (chap. 1, sect. 2.1). Moreover, by reaction it uniquely constructs part of its niche. The
152
other side of the coin is that if redundancy really occurs, i.e. there are coexisting species that do not
153
differ in alpha niche, they will not coexist by this mechanism and some of the 11 other mechanisms
154
mentioned below must account for their presence. The ‘increase when rare’ element occurs here because
155
when a species is rare the resource that it particularly takes up and requires will be present in greater
156
abundance, that is, the niche is not fully occupied, though interferenceinterferencei would use another
157
word here as this means something specific in your text with this process could come through luxury
158
uptake of nutrients (Lipson et al. 1996). The population limitation on the other hand is due to full niche
159
occupancy, i.e. full use of its resource. The degree of niche separation required will increase with the
160
difference in interference ability between the species; if one species is a very strong competitor another
161
species will be able to coexist only if it is occupying a completely different niche. However, in contrast
162
to some stabilising mechanisms of coexistence, if the niche differentiation between two species is strong
163
enough, they can always coexist.
164
1.1 Resources (type of resource and time of availability)
165
Tilman (Titman 1976) demonstrated that coexistence was possible between two algae limited by
166
different nutrients, P and Si, and concluded that the number of species able to coexist is equal to the
167
number of resources and then only if each species is limited by a different resource. He confirmed this
168
in modelling (Tilman 1977), and there appears to have been no contradiction. Vance (1984) claimed to
169
show that two species can coexist on one limiting resource, but only “if each species interferes less with
170
resource acquisition by the other than with resource acquisition by itself”, which with pure competition
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 6 of 34
171
must mean niche differentiation (e.g. the one resource is water, but it is taken up from different soil
172
strata).
173
The primary resource requirements of most embryophytes are similar (light, water, CO2, N, P, K,
174
minor elements, sometimes pollination and dispersal). The concept of a resource gradient as niche
175
differentiation is simple for seed sizes as a resource for birds, but it applies less readily to plants, for
176
which most of the resources are discrete requirements. For example one species cannot require a low-
177
concentration type of P and another a high-concentration type, and the two cannot occur simultaneously
178
anyway. However, in other cases such as soil resources at different depths and pollinator service during
179
the season (section 2), the separation and specialisation of species along gradients are important
180
mechanisms of coexistence (Fig. 4.1; MacArthur and Levins 1967). An important question is how much
181
separation is needed, but in spite of the calculations of MacArthur and Levins this remains unanswered
182
for the real world. Even the existence of such niche limitation has been controversial and difficult to
183
prove (chapter 5).
184
Fig. 4.1. The MacArthur and Levins (1967) concept of niche separation along a gradient.
Fig. 4.1. The MacArthur and Levins (1967) concept of niche separation along a gradient.
185
The above-ground structure of a plant is a light-capture mechanism. Therefore gross
186
characteristics of plant form have great relevance. Consider tree size and shape in rain forest, where
187
light is the prime resource; Kohyama (1992) and Akashi et al. (2003) concluded from modelling that
188
short species, with high seedling recruitment but with height-limited growth, could coexist with taller
189
species with lower recruitment rates. This is basically niche differentiation based on canopy strata.
190
Kohyama (1993) used the model to show that stable coexistence resulted without requiring a stand
191
mosaic and Yokozawa et al. (1996) demonstrated that two canopy shapes, conical and spheroidal, could
192
interact with different speeds of recruitment to give situations that allowed a diverse canopy flora. These
193
are subtle and devious ways in which resource differentiation takes place in this famous biome.
194
Stratification below ground, i.e. in rooting depth, is another important niche gradient (chap. 1, sect. 4.3),
195
especially when there is both precipitation and an accessible water table.
196
Two types of temporal gradient can be seen: (1) If growth is triggered by the resource itself,
197
species can differ in their speed of reaction to resource availability, their opportunism. Opportunistic
198
species react fast to resource availability, for example production of surface roots of succulents,
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 7 of 34
199
ephemeral leaves like Grewia spp. (cross berry) that have a leaf flush after every rain in African summer
200
deciduous bushland, and seasonally produced leaves that must survive periods of resource starvation not
201
an example of opportunism. (2) More commonly seasonal separation of species' growth patterns is
202
controlled not by the resource itself, but by signals such as daylength and temperature. This causes
203
regular seasonal phenological separation of species’ activity. Here, the mechanism overlaps with the
204
Storage effect (section 2).
205
1.2 Heterotroph-imposed niches
206
Pollination and dispersal can be switch mediators (chap. 3, sect. 5.4.G), but also means of niche
207
differentiation. Pollinators come in many sizes and specialisations: insects, birds, mammals and even
208
reptiles. Among insect pollinators there is huge variation in characteristics and their interplay with
209
plants can be rich and complex. There are robbers, mimics, rewards, guides and warnings discuss
210
further or drop this sentence. The pollination niche is liable to the Allee effect, both for self-
211
incompatible plants when there is no mate in the neighbourhood and for those specialised to particular
212
pollinating insects when the plant population is not large enough to attract the pollinator true, but
213
relevant to spp. Coexistence?. Dispersal tends to be less specialised, without an equivalent to the close
214
relation between flower morphology and pollinator morphology seen with some insects and birds, but
215
differences in fruiting times could reduce competition for dispersers. An Allee effect is possible in
216
dispersal if the population is too small to attract more specialised dispersers. Allee effects can neutralise
217
an increase-when-rare mechanism.
218
Vascular plants could occupy different niches by associating with different mycorrhizal fungi.
219
However, specificity within the two main groups of fungi (VAM, ecto, ericoid/epacrid, etc.) is
220
quantitative, in terms of efficacy rather than in absolute ability to colonise the roots. Moreover, the
221
effect on the higher plant with all types is on availability of soil nutrients (especially P) and water, and
222
the loss is in carbon. We conclude that niche diversification through mycorrhizae is unlikelytrue or just
223
unstudied? There are no references presented either way, just stated as fact.
224
1.3 The niche extended by reaction
225
The alpha niche is not a pre-existing box into which a species has to fit. We have emphasised
226
that the individuals of a species react on their environment, changing it and to a lesser or greater extent
227
constructing their own niche. Ramets of a species always show some sort of density pattern (chap. 3,
228
sect. 1) and this pattern must cause patchiness in the micro-environment of the habitat. Litter production
229
is often the basis for nutrient heterogeneity, but plant morphology and root growth can also be the cause
230
(Vogt et al. 1995). Habitats, therefore, must always be patchy in resource availability and physical
231
environmental actors and the size of the patches depends on the pattern of densities of each species.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 8 of 34
232
There is evidence of this, particularly from forests where the sheer size of the trees makes their patches
233
large and easy to sample. For example, Pelletier et al. (1999) examined a mixed-species forest in
234
Québec, Canada, and found using ordination that forest-floor soil was different beneath different
235
species. For example, soil [Ca] was low below Fagus grandifolia (American beech) and they concluded
236
that it reduced the soil [Ca]. In most such observational studies there is a chicken-and-egg problem, that
237
perhaps the soil differences are determining which species grows at a point, not the reverse. However,
238
Pelletier et al. went two steps further: (a) they used spatial statistics to remove spatial correlations, so
239
that as far as possible they were examining the effects of individual trees, and (b) they offered evidence
240
that F. grandifolia produces litter which, from its Ca, lignin, polyphenol and tannin contents, was likely
241
to reduce soil [Ca]. The study of Ehrenfeld et al. (2001) went that step further in another way. They
242
found higher pH below two exotic species in a deciduous forest in New Jersey, USA, than beneath the
243
native Vaccinium spp., but they also grew the species in the greenhouse on field soil and found pH
244
differences in the same direction. The main problem for interpretation here is that the two exotics
245
(planted together) may have raised the pH more because of their greater growth. For another example,
246
you might look at a natural plant community found in the Southern Appalachians, the acidic cove forest
247
(see Schafle and Weakley 1998 for more info.).
248
Presumably plants can have the same effects (your prev, examples were also plants), but there
249
are few published studies. This argument is that every plant community must show pattern associated
250
with each of its constituent species, but there are other biotic forces making habitats more
251
heterogeneous. Litter is not only a significant niche factor, but one especially liable to cause change by
252
reaction.
253
The necessary presence of patches in a community is important to every species’ resource
254
foraging strategy. Large resource patches are best exploited by stoloniferous herbs (Wijesinghe and
255
Hutchings 1997). Jackson and Caldwell (1996) modelled the effect of root plasticity on the uptake of
256
nutrients from heterogenous versus homogenous environments and found that plasticity was
257
theoretically highly advantageous in sagebrush steppe conditions. In the same habitat the effect of
258
patchy soil nutrients differentially affected Agropyron desertorum (wheatgrass) and Artemisia tridentata
259
(big sagebrush), but only under shade conditions where carbon gain was reduced (Cui & Caldwell,
260
1997). Here A. desertorum outcompeted A. tridentata because, although its root proliferation was
261
reduced, its efficiency of P uptake was not and it could exploit rich patches. A. tridentata on the other
262
hand lost efficiency of P uptake and could not make use of patches of higher P availability.
263
Of course, the above-ground environment is altered too. Canopy trees create a
264
shade/temperature/humidity niche for understorey plants, that might not be able to survive without those
265
modifications. Many manifestly cannot do so in mixture with other species, since they are restricted to
266
forest understories. Trees also create the niche for climbers and epiphytes via their support, and they
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 9 of 34
267
create a niche for many epiphytes by water and nutrient stemflow. Parasite plants obviously occupy a
268
niche that would not exist without other plants; the term ‘niche construction’ is very appropriate here.
269
2 Environmental Fluctuation (seasonal, annual and decadal change)
270
Species can separate along annual and other changes in the environment, predictable and
271
unpredictable. We cannot talk of coexistence caused by millenial-scale fluctuation unless the plants are
272
very long-lived (say 500+ yr), because exclusion by interference may happen before the environmental-
273
fluctuation coexistence mechanism can operate. We discuss flowering and fruiting niche gradients,
274
mediated by pollinators and dispersers, in chapter 5 (sect. 6.2) and species can also separate along niche
275
axes of vegetative phenology. An excellent example is the vernal ground flora of deciduous forest.
276
Fargione and Tilman (2005b) found evidence that vegetative phenological niche differentiation added to
277
rooting-depth differences in facilitating the coexistence of species at Cedar Creek with the dominant
278
grass Schizachyrium scoparium (bluestem). Separation in flowering times will reduce competition for
279
pollinators, giving coexistence based on niche differentiation.
280
Environmental fluctuation can cause coexistence if it be on scales shorter than this but long
281
enough for there to be feedback on resources. The fluctuation can be one that affects vegetative growth,
282
for example the vernal flora of forests and spring ephemerals of semi-arid areas, though there is often an
283
accompanying fluctuation in reproduction.
284
As will be clear below, there has to be an interaction between growth and resource supply for
285
environmental fluctuation to cause coexistence. There has been confusion about this. Many authors have
286
claimed that simple variation in the environment and therefore in demographic parameters would allow
287
long-term coexistence. For example, Gigon (1997) wrote: “The fluctuations and their interferences mean
288
that no species encounters optimal growth conditions for a prolonged period of time. Therefore no species
289
can outcompete the others. Fluctuations are thus decisive for the coexistence of species”. Coexistence
290
cannot happen this way. For coexistence, the long-term growth rate of each species has to be RGR = 0.0
291
(r = 0.0, λ = 1.0). The long-term growth rate for a species is the arithmetic average of RGR (the
292
geometric average of λ) in each period. Variation in growth rate will not make it more likely that long-
293
term r is 0.0, in fact a value of exactly 0.0 due to such averaging is infinitely unlikely. Whilst it is true
294
that environmental variation can cause coexistence it can also promote exclusion by interference or have
295
no effect at all on coexistence/exclusion, depending on the biological response of the species to the
296
environment and to competition (Chesson 1990). There are only two ways in which temporal variation
297
can lead to the coexistence of two species: Relative non-linearity and Sub-additivity storage effect.
298
These can cause coexistence only if the interference unbalance between the two species is not too great,
299
and equalising mechanisms can contribute to this.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 10 of 34
300
Relative non-linearity means that two species respond differently to levels of a resource for
301
which they are competing, and moreover that they respond by differently-shaped relations (Chesson in
302
press). For example, the three scenarios in Fig. 4.2 count as different shapes you only discuss the 3rd
303
graph below. The way to test for Relative non-linearity of shapes is to plot the values of RGR of one
304
species at each level of resource R against the values of the other: if the result is anything but a straight
305
line, the species are relatively non-linear.
C
D
RGR
A
F
B
Rmean
Resource level [R]
Rmean
Resource level [R]
Rmean
Resource level [R]
E
Fig. 4.2: Pairs of two species showing relative non-linearity.
306
307
Take the third
308
graph.
C
D
F
B
E
309
310
Take the third graph. If [R], the level of resource R, is constant at the mean value (Rmean), species E has
311
a higher RGR than F. However, if there be environmental fluctuation depends on the degree of
312
fluctuation around the mean, the mean growth rate of F would be higher. Thus, low fluctuation in [R]
313
advantages E, high fluctuation advantages F. The reason this matters is that at low [R] species E grows
314
faster than F and therefore depletes the resources, at high [R] it grows little more than at Rmean, and
315
leaves much of the R unutilised. Both ways, when species E is in the majority it exacerbates the
316
fluctuations in [R]. Conversely when it is in the minority, fluctuation in [R] is lower, which favours it:
317
increase when rare is achieved.
318
In contrast, species F grows little at low [R], and will hardly deplete R. At high [R] it grows
319
disproportionately fast, absorbing R and therefore reducing [R]. Both ways, when species F is in the
320
majority it damps down the fluctuations in [R]. Conversely when it is in the minority, fluctuation in [R]
321
is higher, which favours it: increase when rare is achieved.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 11 of 34
322
323
The second way that environmental fluctuations can cause coexistence is the Storage effect
(Chesson in press. There are four requirements for the Storage effect to operate:
324
1. The species must be competing for a resource.
325
2. They must be affected by an environmental (i.e. non-resource) factor, and respond differently
326
327
to it.
3. There must be covariance between the environmental factor and the intensity of competition.
328
We would expect this, because when the plants are denser and/or larger, competition will
329
be more intense. That is, in ‘favourable’ conditions competition will be greater.
330
4. There must be subadditivity (= buffering, = an interaction between environment and
331
competition). That is to say, when environmental conditions are favourable to growth the
332
effect of competition on RGR is greater. So, whilst ‘3’ refers to the intensity of
333
competition, ‘4’ refers to the effect of competition.
334
In years (or other periods) when the environment is favourable for a species, if it is in the majority and
335
therefore competing against itself it cannot take much advantage of the favourable conditions because it
336
is competing against itself at high biomass: X in Fig. 4.3.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 12 of 34
Competitive intensity
X
Low RGR
High RGR
Environmental favourability
Fig. 4.3: The effect of competitive intensity and environmental favourability on RGR. For ‘X’,
see the text.
337
Competitive intensity
X
Low RGR
High RGR
Environmental favourability
Fig. 4.3: The effect of competitive intensity and environmental favourability on RGR. For ‘X’,
see the text.
338
339
Chesson’s (1994) calculations indicate that the Storage effect is a considerably stronger force than
340
Relative non-linearity. Your explanation of the storage effect and fig 4.3 is not clear (though I
341
understand what the storage effect is)
342
These mechanisms clarify that the timescale on which environmental fluctuation can cause
343
coexistence is set by the timescale on which resource depletion can occur. Light intensity can change
344
instantaneously, it cannot be stored from one second to another and its effects in producing
345
photosynthate are quite short-term, so within-day fluctuation could suffice. Water depletion could occur
346
over a few days, and nutrient depletion over a few months. Soil nutrients often become more available
347
in the spring due to mineralisation, but slow uptake over winter, and are depleted during the period of
348
active growth, so the Storage effect can operate on within-season or between year variation in nutrient
349
use (reword, runs on ). For neither Non-linearity or the Storage effect do the species need to differ in the
350
resources they use. However, they do use them at different times. Seasonal differences in resource use
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 13 of 34
351
can be seen either as Alpha-niche Differentiation or as the Storage effect. (this first mechanism is
352
wanting of a better conclusion paragraph)
353
3 Pest Pressure (heterotroph challenges)
354
Both pathogens and herbivores (from insects to large mammal herbivores) have the potential to
355
give an increase-when-rare process (we use ‘pest’ to cover both pathogens and herbivores). For this,
356
three conditions are required.
357
358
1. Impact: the pests involved must have a significant impact on the growth and/or survival of the
plant species, i.e. their fitness must be reduced.
359
2. Specificity: The pests involved must be to some degree specific to the plant species. It could be
360
sufficient for the species with lowest interference ability to have no specific pests/diseases, but
361
to benefit when the others are suppressed by pests/diseases.
362
3. Abundance-dependence: The challenge from pests must be less on a sparse than on an abundant
363
species. This represents an abundance-dependent effect. The requirement is for a lower impact
364
on the growth and reproduction of sparse species, but this will presumably be through reduced
365
infection.
366
(3.1-3.3 mechansims don’t follow all of the above requiredIf these three conditions… are all three still
367
considered coexistence mechanisms?)If these three conditions metobtain, when any one of the plant
368
species in the mixture becomes more abundant, the host-specific pest (Condition 2) will move more
369
rapidly amongst its host population and the degree of infestation will increase (Condition 3). This will
370
reduce its fitness (Condition 1). This will not directly impact other species, or will do so only to a lesser
371
extent (Condition 2). Conversely, when a species becomes sparse, infestation by its specific pests will
372
decrease and its fitness increase relative to its fitness when more abundant, giving the increase-when-
373
rare effect. Again, the strength of conditions 1-3 necessary for coexistence depends on the degree of
374
difference in interference ability, and equalising mechanisms of coexistence can allow Pest Pressure to
375
result in coexistence when it otherwise would not.
376
Condition ‘1’ can often be met, since pests have various effects on plant production. Basically,
377
the pest organism must have a carbon requirement, which is almost bound to result in lower production
378
and fitness for the plant. Many pests are quite specific to a species or group of plant species, meeting
379
condition 2. Ecologists tend more to question condition ‘3’ because it is less obvious how abundance-
380
dependence could operate. Possible mechanisms (Boudreau and Mundt 1997) are via: (a) a decreased
381
abundance of palatable / susceptible plants, which inhibits the dispersal of herbivores, disease-spores or
382
disease-vectors, (b) the flypaper effect for disease spores and possibly for insect pests, virus vectors and
383
hence for the viruses they carry, in which the pest is caught by a passive surface, (c) alteration in the air
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 14 of 34
384
flow and microclimate, (d) chemicals from associated species that repel herbivorous insects, and (e)
385
promotion by an associated species of natural enemies of the herbivorous insects, i.e. their predators.
386
We shall discuss diseases and herbivory separately since they act in different ways with different
387
dynamics. However, there is sometimes evidence for abundance-dependent mortality or a reduction in
388
growth that is circumstantial evidence for the process and therefore for the Pest Pressure mechanism,
389
that cannot be attributed to a particular pest. For example, Packer and Clay (2000) examined the
390
distribution of seedlings of Prunus serotina (black cherry). The greatest number of seeds germinated
391
quite close to their parent tree, 5-10 m. However, 4 months later and thereafter up to 28 months,
392
seedling survival was higher the greater the distance from the parent tree, up to the furthest distance
393
monitored (30 m). This is not always the pattern. Dalling et al. (1998) in tropical rainforest on Barro
394
Colorado Island found that seedlings tended to be denser nearer to an adult of the same species. For
395
temperate forests Houle (1992) found the seedling mortality of Acer saccharum (sugar maple) in an
396
Eastern American forest was not abundance-dependent; there was no particular spatial relation between
397
trees and seedlings. Hyatt et al. (2003), in a thorough review of the literature, found no evidence for an
398
effect of distance from conspecifics on seed survival in either temperate or tropical communities, but
399
there was a tendency for seedlings to show higher survival at distance, with hints that this occurred
400
especially in the tropical forests. This matches the conclusions of Wright (2002) who, with a rather
401
different review approach, found considerable evidence of low growth performance of saplings near
402
conspecific adults. When effects like this are found, they are assumed to be because of Pest Pressure,
403
though Wright discusses other explanations. This kind of pattern recalls the ‘Janzen-Connell’ hypothesis
404
that whilst the greatest density of fruit will be dispersed to near the parent plant, pests will have the
405
greatest impact there, so the maximal regeneration will occur at an intermediate distance from the
406
parent. This would be expected to be an important abundance-dependent method of species'
407
maintenance in diverse, stable communities of trees, specifically tropical rain forest, but it seems that it
408
is far from universal.
409
3.1 Pathogens
410
Pathogens act in the soil, in the plant systemically and in the plant’s photosynthetic and
411
reproductive systems. The impact of fungal pathogens can be considerable. Mihail et al. (1998) found
412
that in a greenhouse experiment with the annual legume Kummerowia stipulacea (Korean clover) the
413
fungus Rhizoctonia solani caused mortality that reduced plant density by 40 % whilst the fungus
414
Pythium irregulare reduced density by 80 %. C.E. Mitchell (2003) found that in an oldfield grassland at
415
Cedar Creek 8.9 % of the leaf area was infected by fungal pathogens, which decreased root production
416
by 25 % by decreasing leaf life, while herbivores had no effect.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 15 of 34
417
418
The species-specificity requirement is less easy to meet. Gilbert (2002) concludes that the
evidence so far indicates that in natural communities that most fungi infect a large number of hosts.
419
However, the abundance-dependence criterion can often be met. Most ephemeral pathogens are
420
transmitted aerially including rusts and smuts which affect leaves, stems and floral parts. Abundance of
421
the host plant affects the pathogen's population and its persistence, and the transmission of specialist
422
pathogens can be highly sensitive to the identity of other host species in the community (Boudreau and
423
Mundt 1997). An example is the Ustilago violacea smut on Silene alba (≡ S. latifolia; white campion),
424
for which Thrall and Jarosz (1994) experimentally compared the behaviour of the host and pathogen
425
populations to theoretical models. The match was good and showed that both density dependence and
426
frequency dependence occurred. An excellent study by Burdon et al. (1992) described the mortality of
427
Pinus sylvestris (Scots pine) caused by the snow blight fungus Phacidium infestans as being mostly
428
abundance-dependent, with greater mortality in subsites where the host had been denser the previous
429
year. This abundance-dependence has sometimes been shown to lead to a lower pathogen load in
430
mixtures, which must imply some host specificity. C.E. Mitchell et al. (2002) examined 147 plots in an
431
experiment at Cedar Creek established by Tilman and co-workers, sown and weeded to species richness
432
from 1 to 24 species. The percentage of each leaf visibly infected was guessed, using calibrated cards as
433
a guide. Infection dropped as species richness increased, the 24-species plots having only 37 % the
434
foliar fungal pathogen load of the mean monoculture (though more than the least-infected monoculture).
435
Similarly, C.E. Mitchell et al. (2003) analysed another Cedar Creek experiment sown and weeded to 1
436
to 16 species, and the pathogen load in the 16-species plots was only 34 % of that in the mean of
437
monocultures. In chapter 3, sections 6 and 7.5 we described the ‘selection’ artefact in overyield and
438
invasion-resistance experiments. A similar artefact would be possible here if the species less susceptible
439
to disease had thereby an interference advantage and increased its proportion in the mixture, so that the
440
mixture had a lower mean pathogen susceptibility and thus a lower pathogen load. However, C.E.
441
Mitchell et al. (2002; 2003) present evidence that this is not the cause of the effect they found. C.E.
442
Mitchell and Power (2006) conclude that “the transmission of specialist pathogens can be highly
443
sensitive to the identity of other host species in the community”.
444
Similar effects could be caused by below-ground pathogens. Bever (2003) modelled this, but
445
concluded that there is no evidence for it yet.
446
3.2 Herbivory, general (This section does not convince me that herbivory can
447
clearly act as a mechanism of coexistence. How can a plant species be selected
448
based on this mechanism?
449
450
Herbivores come in all sizes, specialisations and guilds. On vegetative parts there are leaf eaters,
stem borers and root eaters. On reproductive systems there are flower exploiters, frugivores and
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 16 of 34
451
granivores. Plant species are variously adapted to herbivory, with chemical and physical defences, life
452
histories and growth patterns that have evolved seemingly to deal with the challenges. The potential
453
mechanism for coexistence via herbivory is similar to that for pathogens (chap. 2, sects. 7.3, 7.4).
454
However, whilst pathogens often reduce the functional efficiency of plant parts, most herbivores simply
455
remove plant material so that the plant needs to regrow to replace tissue and thus its resource base.
456
Obviously a great variety of relationships can be expected between plant species and their herbivores,
457
including symbiotic ones, such as Tegeticula spp. and Parategeticula spp. (yucca moths; James et al.
458
1993).does this fit in with a sect about herbivory?. In many of these systems herbivores exploit plant
459
populations in an abundance-dependent way. Grover (1994) modelled this and used the keystone
460
concept (chap. 5, sect. 11 below) to suggest that a controlling herbivore is one that holds down the
461
abundance of a potentially-dominating plant species and thus allows subordinate species to survive. We
462
may distinguish between abundance-dependent culling of seeds and seedlings and wholesale removal of
463
plant material, i.e. vegetative herbivory.
464
3.3 Herbivory of disseminules and seedlings
465
Seeds and seedlings are a rich nutritional resource and are heavily predated. Maron and Gardner
466
(2000) showed by modelling that herbivores can control adult population abundance by limiting the
467
seed input to the seedbank. Such limitation seems to be widespread. It occurs also via vegetative
468
disseminules. Thus, the ‘impact’ requirement of the Pest Pressure mechanism can be met by
469
disseminule/seedling herbivory.
470
Disseminule herbivory can often be abundance-dependent. Cygnus bewickii (swans) eat the
471
turions (disseminules, fleshy buds) of Potamogeton pectinatus (pondweed) in the autumn. Jonzen et al.
472
(2002) demonstrated clear abundance-dependent control of the P. pectinatus in which the denser patches
473
of turions were exploited, reducing their density, while areas of low turion densities were unexploited
474
and here the plant density subsequently increased. Edwards and Crawley (1999) examined four species
475
of British meadows and found that granivory by rodents was abundance-dependent, but its effects on
476
adult densities differed. Densities of species with bigger seeds (Arrhenatherum elatius, oat grass; and
477
Centaurea nigra, knapweed) appeared to be reduced, but in the smaller seeded Rumex acetosa (sorrel)
478
and Festuca rubra (fecundity?) survival increased to compensate for seed predation, with no overall
479
effect on plant density. Again, Ehrlen (1996) found that in Lathyrus vernus (spring pea), although seed
480
predation by a Bruchidae beetle was correlated with seed density in small plots and with inflorescence
481
size, this had no consistent effect on plant population recruitment. Thus, even the occurrence of
482
abundance-dependent seed predation is no guarantee that it will control the population and hence
483
contribute to species coexistence.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 17 of 34
484
Another limitation of disseminule/seedling herbivory as a mechanism of coexistence may be that
485
most mammal granivores are not specific to one species. There is huge literature assuming that bird
486
granivores are restricted by beak size to a particular range of fruit/seed sizes, but not to one species.
487
However, species of Bruchid beetle are generally restricted to the seeds of one or a few species of
488
Fabaceae (legumes). Seedlings are presumably eaten by invertebrates, but they are likely to be quite
489
generalist.
490
3.4 Vegetative herbivory
491
Herbivory on vegetative parts can be considerable. Vertebrate herbivores (Jones 1933) are
492
involved, and insects both above- and below-ground (Brown and Gange 1989). The selectivity of
493
herbivores varies widely. Many large non-ruminant animals such as Loxodonta africana (African
494
elephant), even though they have preferences, will readily eat a wide range of species. More
495
importantly, some such as Equus spp. (horses) often graze finely-patterned vegetation at a relatively
496
coarse scale, necessarily taking in species with a range of palatability. Other ungulates such as sheep
497
and cattle are more selective, some tiny flower gall wasps are confined to one or a few plant species,
498
and some lepidopterans feed on only one species such as Tyria jacobaeae (cinnabar moth) on Senecio
499
jacobaea (ragwort): Plate 4.2.
500
There are cases when coexistence can be attributed rather clearly to insect vegetative herbivory.
501
For example, Carson and Root (2000) found that periodic plagues of folivorous chrysomelid beetles
502
checked populations of the dominant Solidago altissima (goldenrod) in an oldfield in New York state,
503
USA, and were responsible for the diversity and successional rates. The effect of vertebrate grazing in
504
increasing species diversity is well known, as shown by the rabbit exclosures erected by Tansley and
505
Adamson (1925). However, the effect there is surely that when the sward is higher, light competition is
506
more important, and especially there is more opportunity for the feedback between the outcome of
507
competition and interference ability to occur (chap. 2, sect. 2.3), allowing exclusion by
508
interference.evidence? This has nothing to do with any mechanism of coexistence between plant
509
species why not, seems to be a stabilizing mechanism to me.
510
3.5 Pest Pressure conclusions
511
The Pest Pressure effect seems most likely to operate via diseases. However, there can be
512
complex interactions. An example is seen on Netherlands sand dunes, on which there are relatively
513
uniform soils and clear successional sequences. The growth of the pioneer Ammophila arenaria
514
(marram grass) is impacted by nematodes and pathogens (Van der Stoel et al. 2002). The nematodes
515
reduce growth only early in succession and early in the season, and are not significant in the marked
516
decline in vigour commonly seen in older stands of A. arenaria. However, there is apparently a
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 18 of 34
517
synergistic effect between fungi, and between fungi and nematodes, which reduces A. arenaria’s growth
518
late or early in succession?, and this is important in mixture with Festuca rubra, the next dominant in
519
the succession (de Rooij-van der Groes 1995; Van der Putten and Peters 1997). Still Later in the
520
succession pathway , in nutrient poor grasslands, there is often a mosaic of Festuca rubra and Carex
521
arenaria (sand sedge). Here Olff et al. (2000) discovered that each species had phases of increased and
522
decreased vigour, replacing each other, and that this process was associated with pest phases,
523
particularly the plant-feeding nematodes. Each species seems to be affected by different groups of pests,
524
leading to the changing mosaic aspect of the vegetation, which might look superficially like cyclic
525
succession (chap. 3, sect. 4).
526
4 Circular Interference Networks
527
Interference relations between a set of species are said to be transitive if the species can be
species A
species B
species C
Fig. 4.4. A circular competitive network between three species.
species A
species C
species B
Fig. 4.4. A circular competitive network between three species.
528
arranged in a pecking order, such that a species higher in the order can always competitively exclude
529
one lower down. An alternateThe opposite situation is the existence of circular interference networks
530
(Fig. 4.4). If such networks exist, they would contain an increase-when-rare mechanism: as species A
531
starts to displace species B, species C increases because it has high interference ability against A, but
532
then it in turn is replaced by B, completing the cycle.
533
Simple questions are not always neatly answerable. First, we note that the question can be asked
534
only in one environment, for competitive abilities will change with the environment (Keddy et al. 2000;
535
Fynn et al. 2005). Clearly they must; that is the main reason there is different vegetation in different
536
places. Second, the species that dominates the mixture will be the one with the higher relative growth
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 19 of 34
537
rate, but as interference proceeds the proportions of the species will change and as a result the relative
538
RGRs of the two species may change. Therefore, the eventual result must be judged in terms of
539
exclusion by interference (often loosely referred to as ‘competitive exclusion’). Yet we know that for a
540
variety of reasons have been presented in this chapter for why(Chapter 4) exclusion by interference does
541
not always occur. For these cases, the question of transitivity cannot be asked.
Several studies have determined interference ability by comparing of species’ performances in
542
543
mixture with those in monoculture. Connolly (1997) pointed logical flaws in this. Correction can be
544
made for the “size-bias”, but the basic error has been comparison with a monoculture. We are tempted
545
to conclude that if species A grows more slowly in mixture than in its monoculture whilst species B
546
grows faster in mixture than in its monoculture, B has the higher interference ability. Yet Connolly’s
547
table (4.2), over the undefined period of his artificial data and assuming a starting biomass of 1, gives an
548
example where A does worse in mixture than in monoculture, and B does better in mixture than in
549
monoculture. Yet A has the faster growth rate in mixture (loge 2.77 – loge 1 = 1.02) than Species B (loge
550
2.71 – loge 1 = 1.00) and will come to exclude its competitor from the mixture (subject to the conditions
551
mentioned above). If B goes extinct it can hardly be said to have had the higher interference ability.
552
How does this demonstrate a circular interference network?
Table 4.2. Which species has the higher interference ability? The
starting biomass for both species was 1.00
Species
Biomass in
Biomass
monoculture
in mixture
A
3.00
decrease
2.77 Winner in mixture
B
2.64
increase
2.71
553
It turns out that what is essential in designing such an experiment is not the monocultures, as
554
many people had thought, but two harvests so RGR can be calculated. This invalidates almost all the
555
studies of transitivity done so far. All we have to do is to wait, perhaps for close to infinite time, and see
556
which species has the higher growth rate as the mixture approaches one of themawkward way to say
557
this. This is coming to be one of those community ecology questions that are impossible to answer.
558
At the moment, it is interesting to look at the imperfect evidence available. Buss and Jackson
559
(1979) claimed several competitive cycles for coral reef sedentary organisms, as seen in static evidence
560
for overtopping. Likewise, Russ (1982) claimed non-transitive relations between species in the
561
overgrowth of sedentary marine organisms observed colonising experimental plastic sheets in the sea in
562
Australia, though no cycle can be made out of his results.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 20 of 34
563
Turning to pure plant work, Mouquet et al. (2004) grew eight meadow herbs species in
564
replacement mixture in all possible pairs. Using relative yield (RYi,j = biomass of species i when
565
growing with species j / biomass of i in monoculture), if the species form a transitive hierarchy it should
566
be possible to arrange them so if species i is further up the hierarchy than species j, and RYi,j-RYj,I is
567
always positive. In his experiment, at both low and high density, it almost is, and with a very similar
568
order (Table 4.3).
569
Table 4.3. Competitive hierarchy from Mouquet et al. (2004), strong competitors at the top
High density
Holcus lanatus
Rumex acetosella
Cerastium glomeratum
Anthoxanthum odoratum
Festuca rubra
Arabidopsis thaliana
Lamium pupureum
Veronica arvensis
Low density
Holcus lanatus
Rumex acetosella
Cerastium glomeratum
Anthoxanthum odoratum
Festuca rubra
Lamium pupureum
Arabidopsis thaliana
Veronica arvensis
570
571
At each density, there is one negative RY1,2-RY2,1 indicating a conflict with the hierarchy, it is between
572
species not contiguous in the hierarchy, but it is of size -0.05 andor -0.06 respectively which is clearly
573
within the experimental error.
574
A study that returned a clear answer to the question of transitivity is that of Roxburgh and
575
Wilson (2000a). It relates to a real community, since the seven species used in the interference
576
experiment were taken from that community, the University of Otago Botany Lawn, grown in lawn soil
577
in boxes placed near the lawn. The use of 10 replicates in careful experimental conditions allowed
578
significance tests. The seven species could be arranged in a hierarchy to which all significant
579
competitive relations conformed, i.e. if species X is higher in the hierarchy and species Y lower, then
580
the suppressive effect of X on Y is greater than that of Y on X (Fig. 4.5). In fact, relations between all
581
pairs of species, significant or not, were compatible with the hierarchy.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 21 of 34
582
583
Figure from most competitive to least? Adding an arrow to indicate this would be nice. Explain this
584
figure more. Why no arrows from Holcus to trifolium, Holcus to ranunculus, etc?
585
Fig. 4.5: Competitive relations in seven species from the University of Otago Botany Lawn. From
586
Roxburgh and Wilson (2000a).
587
The experimental design of Keddy et al. (1998) comprised planting a number of ‘wetland’
588
species into a number of swards of wetland species. They report results for 18 species planted into five
589
swards. The 18 species tended to respond similarly to different swards, e.g. Kendal’s coefficient of
590
concordance took a rank of 0.7 (1.0 = complete agreement as to which target suffered more/less), highly
591
significant. Some of the variation in invader/sward? Explain this study more combinations could be due
592
to experimental error (no replication was possible), but some results are impressive, e.g. the rank of
593
Carex crinita (sedge) varied only from 14 to 17 across the 5 swards (18=suppressed most), and Lythrum
594
salicaria (purple loosestrife) varied from 4 to 7 (1= suppressed least).
595
In a different approach, Silvertown et al. (1992) used data from an experiment where several
596
species had been planted in adjacent hexagons, and invasion between hexagons recorded. Examining the
597
difference between the invasion of Species A into Species B and that of Species B into Species A,
598
replacement rates could be calculated. A pecking order can be formed from these results (Fig. 4.6), with
599
no discrepancies (though L. perenne (ryegrass) and C. cristatus (dog’s tail) could equally well exchange
600
positions). There are qualitative discrepancies. Since H. lanatus (Yorkshire fog) can invade P. trivialis
601
(A) and P. trivialis can strongly invade L. perenne (B), the expectation would be that H. lanatus would
602
be able to invade L. perenne even more strongly, but in fact their invasion rates are exactly balanced
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 22 of 34
603
(C). Moreover, although the species A. stolonifera (creeping bent) at the top of the order can invade C.
604
cristatus at the bottom, the rate of replacement is less than for other pairs (D).
Agrostis stolonifera
Holcus lanatus
A
Poa trivialis
D
B
C
Lolium perenne
Cynosurus cristatus
605
Agrostis stolonifera
Holcus lanatus
A
Poa trivialis
D
B
C
Lolium perenne
Cynosurus cristatus
606
607
Key:
Strong (> 0.2) difference in invasion rates
608
Weak-moderate difference in invasion rates
609
Invasion rates equal (i.e. no net invasion)
610
Fig. 4.6. The competitive hierarchy from invasion rates in data of Silvertown et al. (1992).
611
In a similar experiment Silvertown et al. (1994) used only four species, so there was less
612
opportunity for intransitivity, but in any case there was none in any of the four grazing treatments (Table
613
4.4).
614
Table 4.4: Competitive hierarchy of four species in four treatments in Silvertown et al. (1994).
615
Summer sward
grazing height
Winter and
spring
Invasion ability: greater → lesser
3 cm
Grazed
Lolium perenne → Festuca rubra → Schedonorus phoenix → Poa pratensis
3 cm
Ungrazed
Festuca rubra → Lolium perenne → Poa pratensis → Schedonorus phoenix
9 cm
Grazed
Festuca rubra → Lolium perenne → Schedonorus phoenix → Poa pratensis
9 cm
Ungrazed
Lolium perenne → Festuca rubra → Poa pratensis → Schedonorus phoenix
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 23 of 34
(1)
(2)
A
A is taller
than B and
shades B
out
C produces an
allelopathic
chemical,
toxic to A
C
C is shadetolerant, and
scavenges
nitrogen
C
616
A
o
Tree A is taller
than shrub B
and
shades B
out
grass C lowers the
temperature, and
suppresses
seedlings of A
B
B is taller
than C
(3)
A
B
C
shrub B shades out grass
C, and is not affected by
lower temperature
(4)
A is taller
than B and
shades B
out
B is taller than C,
and fixes N
A
A is taller
than B and
shades B
out
C with A is taller
than it, and
shades it out
B
C
B
B with C is taller than it,
and shades C out
.
(1)
A is taller
than B and
shades B
out
C produces an
allelopathic
chemical,
toxic to A
C
C is shadetolerant, and
scavenges
nitrogen
A
o
A
Tree A is taller
than shrub B
and
shades B
out
grass C lowers the
temperature, and
suppresses
seedlings of A
B
B is taller
than C
(3)
C
(2)
A
B
C
shrub B shades out grass
C, and is not affected by
lower temperature
(4)
A is taller
than B and
shades B
out
B is taller than C,
and fixes N
B
A
C with A is taller
than it, and
shades it out
C
A is taller
than B and
shades B
out
B
B with C is taller than it,
and shades C out
617
618
619
620
Fig. 4.7. Possible causes of intransitivity between three species: A, B and C.
It’s interesting to wonder what ecological processes would give rise to intransitivity (Fig. 4.7). In
scenario ‘1’, we use an allelopathic chemical produced only by C and toxic only to A. This works, but
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 24 of 34
621
species-specific allelopathy is rather like Getafix’s magic potions in the Asterix books??: it can
622
perform/explain any wonder. Scenario ‘2’ is similar, except that the third factor is lower temperature
623
(Ball et al. 2002) rather than a toxin. In ‘3’, we have to ask why C can suppress A; presumably the
624
shade-tolerance of C minimises the competition for light, so competition for N becomes important, and
625
C has the lower Tilman R* (chapt. 6, sect. 7.1). Why cannot C suppress B? Perhaps because it is shorter
626
and so cannot compete for light, and its low R* for N does not help because B can fix N. Does this
627
work? Probably. In all three cases, not all pairs are interfering using the same resource/factor. Could we
628
envisage a 3-species solution using competition for light (‘4’)? How can there be heights of A>B, B>C
629
and C>A? Differential plasticity allows such magic: in this case probably by red:far-red effects (chap. 2,
630
sect. 2.6). However, we are again introducing a second factor: light spectrum in addition to light
631
intensity. All this is rather convoluted, which suggests that intransitivity will not be the norm. Move
632
this theoretical discussion to before the examples
633
The evidence is that circular interference networks are uncommon. They have not been observed
634
in plants. In retrospect we should have expected that, because we had not thought what mechanisms
635
would cause them, and such mechanisms are difficult to envisage. This is almost certainly not an
636
important mechanism of coexistence.
637
5 Allogenic Disturbance (disrupting growth, mainly mechanically)
638
Disturbance can have the same effect as climate variation (Roxburgh et al. 2004), but the true
639
Intermediate-timescale disturbance mechanism is a patch mechanism: within an area there are patches
640
of different time since disturbance, with different suites of species (J.B. Wilson 1994a). This gives a
641
successional mosaic. Whether this comprises coexistence depends on the scale at which the system is
642
viewed. Coexistence is seen only when considering a scale that is larger than the size of a disturbance
643
patch, so that it includes patches of differing time since disturbance – newly-disturbed versus recovered.
644
The different patch types are different beta niches, but on a small scale. A, and a species specialising in
645
a particular patch type will increase when rare because it will have more of its specific resources
646
available. In a sense, Allogenic Disturbance should not be counted as a mechanism of coexistence; we
647
do so here because it is so frequently seen as one, because of the impossibility of defining the target
648
scale, and because disturbances occur on all scales so that however small a scale we examine there will
649
still be disturbances within it. The disturbance does need to be sufficiently frequent that each patch will
650
usually include patches at various stages of recovery explain, or the mechanism will not operate.
651
All types of allogenic disturbances happen and it is not always easy to separate disturbance from
652
climatic stress, i.e. the environmental fluctuation discussed above (section 2). The characteristics of
653
disturbance are: (1) Plants not only fail to reproduce but are killed, at least above ground. (2) Most
654
species are killed, not just those that cannot tolerate a particular stress. (3) The event is sudden. (4) The
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 25 of 34
655
environmental effect is temporary, i.e. it is a pulse perturbation, so the original species can re-establish
656
the composition of the patch. However, the real difference is that Allogenic Disturbance is a between-
657
patch mechanism. Disturbance is common, creating gaps over the landscape at a range of scales from
658
meteor hits (many km2) to worm casts (about 0.03 m). Fossorial rodents, ants and termites act at scales
659
which can be important for individual plants. A good example is that of McGinley et al.’s (1994)
660
description of enriched harvester ant mounds in western Texas. It is possible that much of the variation
661
seen in communities is due to old disturbances, where vegetation cover has been regained, and obvious
662
pioneers have been eliminated, but differences in species composition remain. Perhaps we do not realise
663
this because ecologists fail to recognise mid-succession species as being such (Veblen and Stewart
664
1982).
665
The Allogenic Disturbance mechanism assumes that there are distinct pioneer and climax
666
species, i.e. r and K, R and C-S. However, we are talking of secondary succession, and cannot assume
667
this. Peterson and Pickett (1995) found that after windthrow disturbance in a North American conifer /
668
deciduous forest some species regenerated by the germination of seed and some from already-present
669
seedlings, but pioneer shade-intolerant species were sparse, apparently due to a lack of propagule input.
670
Autosuccession, in which the climax species immediately re-establishes after a disturbance, is known
671
from mesic areas such as after windthrow in temperate Nothofagus rainforest in New Zealand
672
(Cockayne 1926), but it is specially found under environmental stress, as predicted by C-S-R theory
673
(chap. 6, sect. 6.7 below; J.B. Wilson and Lee 2000). We might expect that the greater species richness
674
in tropical rain forests would include a good number of gap specialists. Indeed, Hubbell (2005)
675
demonstrated for Barro Colorado Island tropical rainforest a close negative correlation among species
676
between survival rate in shade and growth rate in full light (in gaps), though admitting there were rather
677
few gap species and their abundance was low. Wright et al. (2003) found that there was a continuous
678
distribution of gap-colonising species and those that avoided gaps in Barro Colorado Island, Panama,
679
but that the majority were rather indiscriminate. Similarly, Lieberman et al. (1995) found that 87 % of
680
the tree species in Costa Rican tropical forest had no significant canopy-gap / matrix specialization.
681
Poorter et al. (2005) found that only one of 47 species in a Liberian tropical rainforest was a shade
682
species for its whole life, and only one a light species for its whole life. It is clear that most species are
683
intermediate in this respect. This suggests that Allogenic Disturbance may not be an important
684
mechanism of coexistence in the very biome where we tend to envisage it. Yet in temperate forests there
685
may be greater opportunities for it to increase species richness: Poulson and Platt (1996) demonstrated
686
in Michigan that the size of the gap affects the species re-establishing, such that single treefalls favoured
687
Fagus grandifolia (American beech) but multiple fall gaps favoured Acer saccarum (sugar maple). This
688
is not a question of gap versus non-gap, but also of differences between different sorts of gap.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 26 of 34
689
Gaps are by no means restricted to forests. Grubb (1982) suggested that in roadside communities
690
around Cambridge, England, the climax dominant amongst the grasses was Arrhenatherum elatius (oat
691
grass), but Dactylis glomerata (cocksfoot) and Plantago lanceolata (ribwort plantain) retained their
692
place in the community by being the first to invade small gaps.
693
6 Interference/dispersal Tradeoff
694
This concept originated simultaneously with Skellam (1951) and Hutchinson (1951). It has been
695
known under a variety of names (J.B. Wilson 1990), including ‘Life History Differences’ and the
696
endearing if not entirely accurate ‘Musical Chairs’ (Crawley 1986). Consider a model in which two
697
annual species occupy single-plant safe sites. Species C is the better competitor, and eliminates the
698
weaker competitor D if it reaches a site, but it has less efficient reproduction/dispersal than species D
699
and therefore fails to reach some sites. Species D has better dispersal and is therefore available to
700
colonise most of the sites that C has not reached. If C becomes sparse, there are many empty sites for its
701
offspring to occupy and its population growth rate increases; similarly if D becomes sparse, there are
702
many sites left over by C for it to occupy. This is increase-when-rare. The mechanism can be
703
distinguished from (‘1’) Niche differentiation in that no differences between species in resource use are
704
required. It can be distinguished from (‘5’) Allogenic Disturbance in that: (a) the gaps are caused by
705
monocarpic or seasonal death, not necessarily by external disturbance, though that is possible, and (b)
706
species C is limited only by dispersal, not by its ability to tolerate the environment of the gap. It can be
707
distinguished from (‘9’) Equal Chance in that, though there is a random element, it acts via dispersal;
708
the interference abilities of the two species are very different.
709
There have been many mathematical models of the mechanism, e.g. Levins and Culver (1971),
710
Nee and May (1992) and Tilman (1994). There is an assumption of a negative correlation, due to a
711
trade-off, between interference ability and dispersal ability, but Ehrlén and van Groenendael (1998)
712
surveyed the literature and found that this was un?common. Turnbull et al. (1999) demonstrated the
713
mechanism experimentally by sowing seven species from a limestone grassland, ranging from a seed
714
mass of 0.013 to 0.16, back into that grassland. When the seeds were sown at a high density, 83 % of the
715
resulting plants were from the three species with the largest seeds, but when a low density was sown this
716
percentage was reduced to 49 %. This is entirely compatible with the Interference/dispersal Tradeoff
717
mechanism: when there were enough seeds to reach almost all microsites the three big-seeded, strong
718
competitors occupied them, but when fewer seeds were sown there were microsites not occupied by the
719
big three, which the light seeded, probably well-dispersed species could occupy. The unlikely
720
Interference/dispersal Tradeoff theory is proved my intro bio students would loose points for a statement
721
like this in their lab reports.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 27 of 34
722
723
7 Initial Patch Composition
The coexistence model of Levin (1974) is that two species occupy small, transient patches. Some
724
patches will by chance have more individuals of one species than the other. The species in the majority
725
will suppress the other in that patch if intra-specific interference is less than inter-specific interference.
726
The latter condition is beloved of ecological modellers, but it seems unlikely in the real world. It would
727
be possible with mutual species-specific allelopathy: the Getafix potions? of community ecology. We do
728
not believe this model can apply to plants (or at all). Examples? Evidence?
729
8 Cyclic Succession: movement of community phases
730
This topic was covered in chapter 3, section 4. The increase-when-rare mechanism is similar to
731
that of (‘4’) Circular Interference Networks. The latter are between individual species whereas cyclic
732
succession involves the whole community, though in many of Watt’s (1947) examples the community
733
comprises one species. Cyclic succession involves reaction, but then interference also involves
734
environmental modification, be it more temporary. There could be cyclic succession between just two
735
phases, whereas there cannot logically be a circular interference network with fewer than three species.
736
Autoallelopathy, for example, would satisfy our criterion for increase-when-rare because a
737
species that is sparse finds less of its self-toxin in the soil, but there must be some kind of reaction
738
involved, giving abundance-dependence. A mosaic arises because a phase of the cycle that is replaced at
739
one point appears elsewhere; we therefore count it as a mechanism that uses movement to escape
740
exclusion by interference. This also means that the mechanism is scale-dependent: the scale examined
741
has to be one that includes patches of the mosaic in different phases.
742
9 Equal Chance: neutrality
743
It is a longstanding idea that there is an element of chance in which species occurs at a spot
744
(Lippmaa 1939). This has been especially invoked for tropical rain forests (e.g. Schulz 1960; Hubbell
745
and Foster 1986). Sale (1977) described it as a ‘lottery’ and Connell (1978) formally put it forward as a
746
mechanism of coexistence, the ‘Equal Chance’ mechanism. In this section, when we speak of chance,
747
we refer to processes such as dispersal that are so complex as to be unpredictable in practice, combined
748
with equally unpredictable climatic and catastrophic disturbance events. It is impossible to prove the
749
operation of chance, but some have implicated it.
750
Equal Chance means that any one of a number of species is equally likely to occupy and
751
pre-empt by reaction a particular microsite. One cause would be that the probability of a disseminule
752
reaching a site is proportional to its abundance this is equal chance?. Then, dispersal would determine
753
which species occupies a particular site (Schulz 1960). In New Zealand, Veblen and Stewart (1980)
754
used this as an explanation for the colonisation of canopy gaps by either Dacrydium cupressinum
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 28 of 34
755
(rimu), Weinmannia racemosa (kamahi) or Metrosideros umbellata (southern rata) as a function of
756
seed/seedling availability, mast seeding and the ability of many New Zealand tree species to remain as
757
suppressed seedlings. Equal interference abilities are likely to be invoked: the ‘Equivalence of
758
Competitors’ concept of Goldberg and Werner (1983) that often interference is intense, but many
759
species are similar in their interference ability. Sometimes the outcome of interference can be
760
established at the seedling stage, since when competition is for light and therefore cumulative, the first
761
plant to establish will exclude others, a type of inertia (chap. 2, sect. 2.3). This again invokes random
762
dispersal. Alternatively, both the probabilities of ecesis and their interference abilities might be different
763
between species, but the two balance.
764
The Equal Chance concept would result in variation in the species composition of communities
765
which it was impossible to correlate with any environmental factor, present or past. Some have used this
766
kind of negative result as evidence of chance. McCune and Allen (1985) in forests in Montana, USA,
767
R.B. Allen and Peet (1990) in forests in Colorado, USA, and Kazmierczak et al. (1995) in kettle-holes in
768
Poland found only weak correlation between species composition and the environment and invoked
769
chance. In such work, the weak correlation could be because there were important environmental factors
770
that had not been measured, or because some factors measured gave a non-linear response that the
771
analysis could not cope with. The Equal Chance hypothesis, used as an excuse for failing to find
772
vegetation/environment correlations is the last resort of the scoundrel (ouch, that’s pretty harsh. Are
773
you sure you are presenting McCune and Allen (1985) and R.B. Allen and Peet (1990) fairly and
774
accurately?. Lavorel and Lebreton (1992) compared the composition of the vegetation with that of the
775
seed pool in fields from southern France, and took the similarity as evidence of a random draw from the
776
seed pool. This is also doubtful evidence; it could equally well be caused by determinism.
777
The most well-known invocation of chance is the Island Biogeography model of MacArthur and
778
Wilson (1963), based on probabilistic immigration and extinction. However, Kelly et al. (1989) and
779
Tangney et al. (1990) could find little evidence for its operation in Lake Manapouri islands. On islands,
780
a direct test of determinacy v. chance (assuming that incidence functions are not important) is available
781
in a test for nesting. On the other hand, J.B. Wilson (1988d) found plant species nesting among these
782
islands to be significant, but far from complete. His analyses pointed to habitat control rather than
783
chance, at least for native species. J.B. Wilson et al. (1992a) sampled the algal flora of intertidal rock
784
pools, selected for habitat uniformity within a limited area, and analysed as virtual islands. The
785
distribution of species agreed closely with that expected at random, whether examined by the
786
distribution of associations, by nesting, by chequerboarding or by incidence functions. The simplest
787
explanation is that differences in specific composition between the pools are caused by chance, but that
788
is no proof; it is a minimalist default. The best example of chance – no difference between species if one
789
can ever have an example of no difference – is from Munday (2004), who investigated two small
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 29 of 34
790
congeneric coral-reef fish species, where there was evidence for interference, in field removal
791
experiments, and in lab colonisation. However, in none of these experiments, nor in field distribution,
792
was there any evidence of niche differentiation.
793
Even Equal Chance’s strongest advocates have been equivocal. Hubbell (2005), having
794
emphasised differences in niche between Barro Colorado Island tropical rain forest species, eventually
795
attributed coexistence to dispersal and recruitment limitation. This is in effect a resort to Equal Chance.
796
However, he immediately discussed negative abundance-dependence, which is stabilising, not
797
equalising. The Equal Chance mechanism is the equalising mechanism par excellence, and should be
798
seen as no more than that.
799
800
801
What about neutral models ala Hubbel?
10 Inertia
Inertia is another type of equalising mechanism, slowing exclusion by interference and possibly
802
allowing stabilising mechanisms to operate.
803
10.1 Temporal Inertia
804
Temporal Inertia (Cowles 1901) can be an individual or population effect, effectively the same
805
because the ‘individual’ concept is not meaningful for plants. Trees stand where they stand and cause
806
inertia. If, when a tree fell over, there were a tendency for the niche it had constructed to favour its own
807
juveniles in the canopy gap created, this would represent a small-scale dispersal switch. This could
808
constitute inertia, slowing the ingress of a superior competitor. We mentioned above that Dalling et al.
809
(1998) found seedlings on Barro Colorado Island to be denser near to a conspecific adult. Species
810
differed little in the correlation of their growth rate with light intensity and they declared that differential
811
responses to soil and topography were rare. This left them to speculate that there was dispersal
812
limitation, as supported by the correlation between parent and juvenile being weaker for species with
813
small disseminules. This is inertia due to dispersal limitation. Annuals are a conspicuous life form in
814
arid climates where the rainfall is highly erratic and form a long-lived seed bank. This seed bank gives
815
inertia as well as a storage effect. A population with a seed bank has similarities to a tree, except that
816
this is multi-generation inertia. The decade-long dominance phases, with smooth increases and
817
decreases, that Watt (1981) found in the Breckland may owe something to inertia. A more active
818
mechanism of inertia would be delay in vegetation change caused by a switch (chap. 3, sect. 5.3). Inertia
819
caused by the length of life of trees must be common. Abrams and Scott (1989) describe a situation
820
where, until disturbance occurs, early-successional trees dominate the canopy with young plants of later
821
successional stages beneath them, and their model diagram shows the high species richness resulting at
822
this stage.
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 30 of 34
823
These situations are obvious in dryland vegetation where the rainfall is erratic and temperatures
824
are high. For many succulents, sufficient rainfall for the development of surface root hairs is itself a rare
825
event, yet by means of massive storage, CAM photosynthesis and often a high albedo, they can
826
withstand years of unavailable moisture. The establishment of such succulents can be a very rare event,
827
as exemplified by Agave macroacantha in Mexico (Arizaga and Ezcurra 2002). Minor rain events may
828
allow annuals to grow, while deep-rooted shrubs and trees maintain contact with a deep water resource
829
giving the appearance of a plant community. Clarke (2002) described a similar situation with woody
830
dryland vegetation in southwestern Australia, where no natural recruitment of shrubs was observed over
831
five years. However, the rare event required to cause the state change could be a different grazing
832
regime, as Prins and van der Jeugd (1993) found in Tanzania. Two pandemics in the herbivores in 1880
833
(rinderpest) and 1961 (anthrax) temporarily reduced browsing and allowed even-aged stands of Acacia
834
tortilis (umbrella thorn) to establish. These are now a conspicuous and apparently integral part of the
835
vegetation of national parks in the area, yet are present through inertia, not as maintained populations.
836
Here, the state change was anthropogenic, but a similar situation could occur naturally. Inertia may not
837
apply to all the species in a community, since many contain species that differ markedly in survival and
838
establishment probabilities. Extremely long-lived individuals of slow growth exist alongside perennials
839
with lifespans shorter by at least one order of magnitude. The long-lived individuals can establish only
840
during a rare event, which could be a disturbance such as flood, a 1/100 yr wet season. The probability
841
of such an event occurring in any one year is very low and does not change from year to year. Thus their
842
occurrence is stochastic yet within the time scale of very long-lived plants.
843
We still ask what the original coexistence was due to: if there is no coexistence, inertia cannot
844
prolong it.
845
10.2 Spatial Inertia: aggregation
846
Spatial aggregation of the plants of a species also gives inertia, delaying exclusion by
847
interference since it occurs only at patch boundaries. Presumably the aggregation was established due to
848
dispersal processes, an ‘ecological founder effect’. Stoll and Prati (2001) demonstrated beautifully the
849
slowing of exclusion by interference by experimental aggregation. Amongst four annuals they found
850
that the species with least interference ability (Cardamine hirsuta, bitter-cress) decreased over the
851
experiment to 6 % of the monoculture in a random arrangement but only to 26 % in an aggregated
852
arrangement. The species with lowest interference ability (Stellaria media, chickweed) increased its
853
biomass to 324 % of the monoculture in the random but to only 239 % in the aggregated1. This would
854
be a most potent mechanism for delaying exclusion by interference of a subservient species. Rebele
855
(2000) found a similar, but very slight, effect in an outdoor mesocosm experiment using mixtures of
856
Calamagrostis epigejos (reed) and Solidago canadensis (goldenrod).
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 31 of 34
857
Thórhallsdóttir (1990) had planted outdoors a hexagonal grid of adjacent plots. Each plot
858
contained one of five meadow species: Agrostis stolonifera (creeping bent), Holcus lanatus (Yorkshire
859
fog), Cynosurus cristatus, Poa trivialis (meadow grass), Lolium perenne (ryegrass) and Trifolium
860
repens (white clover). Silvertown et al. (1992) ran simulations to see in retrospect what effect
861
aggregation would have, given the invasion rates that Thórhallsdóttir found between the pairs of grass
862
species. After 50 time periods when the species were intermixed in a random pattern, the weakest
863
competitor Lolium perenne had almost disappeared (reduced from 20 % to 1 %), but with the species
864
‘planted’ in bands, depending on the order of the species in the bands, it decreased only to 9 %, stayed
865
at 20 % or even increased slightly to 21 %.
866
Aggregation might also delay exclusion by interference via effects on herbivory (Parmesan
867
2000), fire spread (Hochberg et al. 1994) and other environmental factors.
868
11 Coevolution of Similar Interference Ability
869
Aarssen (1983) suggested that in a mixture of two species stronger selection pressure on the one
870
with lower interference ability would cause it to become the stronger competitor of the two, “Superiority
871
in competition therefore alternates between … members of the two populations”. He later (1989)
872
produced some evidenced for this: over two generations the interference ability of Senecio vulgaris
873
(groundsel) increased relative to a standard genotype of Phleum pratense (Timothy grass) with which it
874
was growing. Selection can result in small-scale genetic change in populations, as apparently occurred
875
in Trifolium repens (white clover) associated with different ecotypes of Lolium perenne (ryegrass) in the
876
pastures that Lüscher et al. (1992) investigated. However, neither this study nor that of McNeilly and
877
Roose (1996) could find evidence of co-adaptation between neighbouring ecotypes of associated L.
878
perenne. Eventual ecotypic evolution in response to neighbours would be expected, and has
879
occasionally been demonstrated (Martin and Harding 1981). However, Aarssen’s proposal is
880
unbelievable because it involves continual increases in interference ability, as Aarssen (1985) has since
881
concluded. The plastic response to interference (chap. 2, sect. 2.2) can also give a buffering effect.
882
12 Spatial Mass Effect (vicinism)
883
The Spatial Mass Effect refers to the maintenance of a population of a species by constant
884
immigration into a patch where the species cannot otherwise maintain itself (Zonneveld 1995). It has
885
been called the sink effect. The immigration could be by seeds, or in theory by rhizomes or stolons.
886
Populus tremuloides and related species (aspen) produce root suckers (Barnes 1966) and these can
887
appear beyond the canopy of the tree where there is no chance that they will survive to be self-
888
supporting, let alone sexually reproductive, for example in a lawn. Seed immigration is the most
889
common but difficult to demonstrate. It is difficult enough to monitor occasional seeds blowing in, and
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 32 of 34
890
even more difficult to demonstrate that the population into which they are blowing would have RGR <
891
0.0 without that subsidy. Snyder and Chesson (2004) have applied the concepts of the ‘storage effect’
892
and non-linear dynamics to coexistence between species that have different tradeoffs of interference
893
versus fecundity+dispersal. Their model has Spatial Mass Effect, though also elements of (‘6’)
894
Interference/dispersal Tradeoff. The effect clearly maintains populations that are not susceptible to
895
considerations of abundance-dependence or increase when rare, the stabilising mechanisms we require
896
here, yet it can maintain coexistence indefinitely.
897
The Spatial Mass Effect has rarely been quantified. Kunin (1998) examined boundaries between
898
plots with different fertiliser treatment in the 150-year old Park Grass Experiment. There was a very
899
sharp pH change, within 50 cm of the boundary. Although there were many exceptions, the majority of
900
plots examined (34 out of 51 non-zero, 2-tailed p = 0.024) showed higher species richness towards the
901
boundary. The effect was seen especially where the two adjacent plots differed more in species
902
composition. The Spatial Mass Effect can be seen clearly in extreme cases where the recipient (sink)
903
population does not reproduce at all, like the 13 species of angiosperm that grow in the Lost World
904
Cavern, northern North Island, NZ, without any of them ever setting seed (de Lange and Stockley
905
1987). Studying an Argentinian steppe with the (palatable) grass Bromus pictus amongst tussocks of
906
unpalatable grasses Stipa spp. and Poa ligularis, Oesterheld and Oyarzábal (2004) found more B. pictus
907
in the upwind part of a grazing exclosure, showing that a seed subsidy was arriving from the grazed
908
area. The tussocks outcompeted the B. pictus when ungrazed, reducing the local seed output in the
909
exclosure. This situation may be the commonest way in which the spatial mass effect operates to
910
maintain species metapopulations.
911
13 Conclusion
912
We believe our review covers all the mechanisms by which species can coexist in stable
913
mixtures. Chesson’s terminology of stabilising versus equalising mechanisms is useful and important. It
914
has focussed attention on the fact that some proposed mechanisms of ‘coexistence’ do not, in fact, cause
915
long-term coexistence. It has also highlighted what few had recognised, that even though the equalising
916
mechanisms cannot on their own cause stable coexistence between two species, they can reduce the
917
difference in interference ability between species to the extent that a stabilising mechanism can operate.
918
We must speculate on the importance of each mechanism in order to build up in our minds a
919
vision of the plant community. The overwhelming reason for species coexistence is Alpha-niche
920
Differentiation. Environmental Fluctuation is probably important. It can be seen as niche differentiation
921
in time, but with special restrictions on when it can operate. In seasonal climates, local environment can
922
vary enormously, both stochastically and predictably. Each of the ecosystem attributes enumerated by
923
Reichle et al. (1975; see also chap. 1, sect. 1 above) must change during the year: the energy base
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 33 of 34
924
(affected by irradiance), the reservoir of energy, nutrient cycling (through mineralisation rates) and rate
925
regulation (temperature, water availability, herbivory). It seems that the available states of these
926
variates, factorially combined, should allow for the coexistence of a very large number of plant species.
927
Pest Pressure may be important; Gillett (1962) suggested that it is the major mechanism, but that
928
remains to be proved. The Spatial Mass Effect must be very common. Disturbance is clearly common,
929
and has potential to allow co-existence; surely all communities are successional mosaics. We earlier
930
discussed autogenic disturbance. It could have been listed as a separate mechanism here, it could have
931
been merged with Allogenic Disturbance since many disturbances are partly allogenic and partly
932
autogenic, or it could have been included with Cyclic Succession since it will often be a component.
933
Other mechanisms are probably of more minor importance. For example, circular interference
934
networks are an attractive idea, but remain undiscovered. Cyclic succession is more believable and from
935
time to time fashionable, but seldom observed and even more rarely proven to take place.
936
Autoallelopathy may be widespread but the soil is an intractable and infinitely complex medium where
937
clear chemical pathways and effects are difficult to prove.
938
Based on the evidence derived from the present literature, we list the mechanisms below in
939
increasing order of importance: This is not a very effective way of ordering these mechanisms…how
940
about grouping into minimal, moderate and strong influencial force? Or, maybe present them in
941
increasing/decreasing order of importance in the text.
942
Initial Patch Composition (7)
943
Co-Evolution of Similar Interference Ability (11)
944
Equal Chance (9)
945
Circular Interference Networks (4)
946
Cyclic Succession (8)
947
Temporal and Spatial Inertia (10)
948
Interference/Dispersal Tradeoffs (6)
949
Allogenic Disturbance (5)
950
Spatial Mass Effect (12)
951
Pest Pressure (3)
952
Environmental Fluctuation (2)
953
Alpha-niche Differentiation (1)
954
Switch to order of decreasing importance. Those lower in the ranking- their order seems
955
arbitrary or would vary with context
956
In a changing, disturbed world it will be increasingly difficult to separate stabilising mechanisms from
957
equalising ones. The temporal turnover of species in communities depends on some species
958
disappearing, others invading, so that many species in a community may be present by courtesy of one
J.B. Wilson & Agnew, chapter 4, Species coexistence, page 34 of 34
959
of the equalising mechanisms and will ultimately be doomed. The multiplicity of possibilities for
960
coexistence should allow the coexistence of a very large number of plant species. The question then
961
becomes: “Why are there so few species in most habitats?”.
962
A plant, however, is sedentary and extends over a spatial volume, necessarily exposed to wide
963
range of environmental conditions. It therefore cannot be confined to a precisely-defined niche. It is the
964
interplay between the potential for high plant diversity in restricted niches and the necessity for plants to
965
tolerate a wide range of environments that encourages us to look for patterns in plant communities. If
966
adaptations to available niches were most of the reason for every species’ occurrence, our enquiry in
967
this book would be less interesting.
968
Footnotes
969
1
970
‘sparse’ would be a better term than ‘rare’, but ‘increase when rare’ is ensconced in the literature, and
so we use it here.
971
TABLES, ILLUSTRATIONS and PLATES
972
Table 4.1: Some Examples of monospecific stands. We exclude monospecificity in a single stratum or
973
974
guild of vegetation, such as a tree species or understorey species.
Table 4.2: Which species has the higher interference ability? The starting biomass for both species was
975
1.00
976
Table 4.3: Competitive hierarchy from Mouquet et al. (2004), strong competitors at the top
977
Table 4.4: Competitive hierarchy of four species in four treatments in Silvertown et al. (1994).
978
Fig. 4.1: The MacArthur and Levins (1967) concept of niche separation along a gradient.
979
Fig. 4.2: Pairs of two species showing relative non-linearity.
980
Fig. 4.3: The effect of interference intensity and environmental favourability on RGR.
981
Fig. 4.4: A circular interference network between three species.
982
Fig. 4.5: Competitive relations in seven species from the University of Otago Botany Lawn. From
983
Roxburgh and Wilson (2000a).
984
Fig. 4.6. The competitive hierarchy from invasion rates in data of Silvertown et al. (1992).
985
Fig. 4.7. Possible causes of intransitivity between three species: A, B and C.
986
Plate 4.1: A monospecific community: Cladium mariscus stand.
987
Plate 4.2: Tyria jacobaeae (cinnabar moth) on Senecio jacobaea (ragwort)
1
all this is in the high-density treatment