Interspecific Contest

What makes the interspecific contest different from intraspecific contest are differences in competing abilities between species, which can be larger than between individuals of a unmarried-species population.

From: Encyclopedia of Ecology , 2008

Construction of Fish Communities in Lakes and Its Abiotic and Biotic Determinants

Thomas Mehner , Sandra Brucet , in Reference Module in Earth Systems and Environmental Sciences, 2021

Competition and resource segregation

Interspecific contest has non been a major focus of studies investigating the structure of lake fish communities. Nosotros are not aware of a study showing that a certain fish species has go excluded from a local lake community every bit the consequence of competition with another local fish species, but invasion by non-native species may induce stronger negative effects. In turn, a dominance of positive co-occurrence patterns of species or habitat-specific groups of species among lakes suggests that facilitation and niche similarity structure richness and composition of lake fish communities substantially ( Fig. 3) (MacDougall et al., 2018). Mutually negative effects on abundance or biomass of competing fish species are more likely (Eloranta et al., 2016), simply systematically negative affluence or biomass correlations between species with similar niches among lakes have not been institute (MacDougall et al., 2018). It has to exist noted that studies of interspecific competition take to consider the frequent ontogenetic niche shifts in freshwater fishes (Werner and Gilliam, 1984), resulting in a succession of dissimilar biotic interactions while fish grow. The effect of competition amongst species is usually best demonstrated past niche segregation between co-occurring fish species (Genner et al., 1999; McMeans et al., 2020). This process is also considered every bit a road towards ecological speciation in lake fish communities (Knudsen et al., 2006). Nonetheless, these young speciation events can be reversed and sympatric species can disappear if anthropogenic influences on lakes destroy the ecological gradients along which niche segregation has been expressed (Vonlanthen et al., 2012).

Fig. 3

Fig. three. Direct and indirect drivers of species richness in fish. Structural-equation-model-derived multivariate relationships amongst integrated abiotic and biotic regulatory factors (blocks   =   degree days, circles   =   lake morphometry, triangles   =   water quality, and biotic factors [ruddy lines]) for the richness and composition of iv major fish functional groups in 721 lakes along an eleven degrees latitudinal gradient in Ontario, Canada (SEM integrative model: n  =   648, MLEST   =   iv.91, Caste of freedom   =   thirteen, P  =   .977). Solid lines indicate negative relationships; dashed lines indicate positive relationships. Arrows indicate the direction of the human relationship. Bold lines indicate stronger relationships, arbitrarily assigned every bit standardized path coefficient values >   0.40. Black lines bespeak abiotic influences on biotic factors and red lines bespeak influences between biotic factors. Functional groups are predator, coastal, pelagic, and minor-prey species.

Reproduced with permission from MacDougall AS, Harvey Eastward, McCune JL, Nilsson KA, Bennett J, Firn J, Bartley T, Grace JB, Kelly J, Tunney TD, McMeans B, Matsuzaki Sis, Kadoya T, Esch Eastward, Cazelles K, Lester N and McCann KS (2018) Context-dependent interactions and the regulation of species richness in freshwater fish. Nature Communications 9: 973.

Stronger effects of competition have been found with respect to size structure of lake fish populations and communities. Interspecific competition modifies the size distribution of lake fish populations, with shifts towards steeper slopes of size spectra where the density of interspecific competitors is high. Even so, the strongest result from competition on size spectra is owing to intraspecific competition (Arranz et al., 2016). The size-at-age declines with increasing density of individuals from the same species, resulting in delayed reproduction and recruitment. Under extreme cases, negative density dependence can induce cyclic population dynamics with alternating strong and weak year classes, due east.one thousand., in vendace (Coregonus albula) (Hamrin and Persson, 1986) or roach (Rutilus rutilus) (Cryer et al., 1986). These results support that biomass and size structure of local fish communities are strongly affected by the available energy provided by autotrophic (phytoplankton, algal biofilms, macrophytes) and heterotrophic (bacteria and fungi) product and the efficiency by which this free energy is transferred to the higher trophic levels (Bartrons et al., 2020).

Read full affiliate

URL:

https://www.sciencedirect.com/science/article/pii/B9780128191668000049

Competition and Competition Models

O. Gilad , in Encyclopedia of Ecology, 2008

Interspecific Competition

Any contest between populations affects the fettle of both. The resources invested (energy, time, and matter) in the contest or abstention of information technology, reduces availability of these resources and adversely affects the reproduction success of the populations. When one population fully or partially depletes a limited resource, the availability of such resource is reduced for its competitor. Due to the cost each population pays in reduction of fitness as a issue of competition, it is advantageous for each party to avoid competition which has resulted in natural pick favoring niche separation, specialization, and diversification.

Interspecific competition is the competition betwixt individuals of dissimilar species. Under this contest type nosotros likewise recognize two types of competition merely the interaction is between individuals of different species and not individuals of the aforementioned population every bit is the case in intraspecific competition.

i.

Exploitation (consumption) competition. An indirect interaction between species (populations) over a limited resource is where ane population deprives the other of that resources. Indirect interactions may involve three or more different populations of species and will fall under one of three scenarios:

(a)

Classical exploitative contest due to resource depression. Two consumers sharing a mutual casualty may compete indirectly by one competitor consuming the resource before the other has access to it thereby depriving its competitor of resource availability.

(b)

Credible competition. Two prey species may appear to compete because, if either increases, a shared predator too increases, which operates to the detriment of the other prey population.

(c)

Competitive mutualism. Two populations are weak competitors but are both strong competitors with a third population. By each of the weak competitors inhibiting the third (strong) competitor, they both do good. This scenario tin can also ascend in a iv-species system (facilitation), where ii species each eat one of two competing prey species. In this scenario if ane predator species increases, information technology results in decrease in its casualty species which benefits the competing prey species which can now increment. As a consequence, predator 2 increases since its prey is now more arable only in time it decreases the prey availability benefiting casualty one and then on.

2.

Interference competition: allelopathy or interspecific territoriality. This involves straight interaction betwixt species (populations) over a express resource by reducing access of i population to that resource. This tin can exist accomplished past establishing a territory, obtaining a say-so status within a hierarchical species, or releasing toxins into the footing preventing other institute population from establishing themselves. If the competition is for space, 'preemptive competition' may arise, determined mainly past which species arrived at the resources first.

A prepare of relationships between two species competing for a express resource have been proposed by Lotka and Volterra and has been used equally a footing for analysis in ecological studies.

Interspecific Competition and the Lotka–Volterra Model

In 1925, Alfred J. Lotka and Vito Voltera described competition in a set of simplified equations. These equations are a modification of the Verhulst–Pearl logistic equation (see Growth Models) and are based on the aforementioned assumptions.

Lotka–Volterra equations consider two competing species Ten ane and X 2 with an intrinsic growth rate r 1 and r 2 and a conveying capacity of Grand one and K 2 in the absenteeism of contest. 'Carrying capacity' is the maximum number of a population that a habitat can back up. When calculating the simultaneous growth of each species (population) if both occupy the same habitat the relationship betwixt the competing species is as follows:

[3] d X 1 d t = r 1 X 1 K 1 X 1 α 12 X 2 K 1

[4] d 10 2 d t = r 2 X 2 Thousand 2 X 2 α 21 X ane Yard 2

α12 represents the inhibitory effect species Ten 2 has on the population growth of species X 1 and α21 is the inhibotory effect species Ten one has on the population growth of X 2 (competition coefficients).

The inhibitory result of each population growth affects both the population itself (intraspecific competition) and the competing species population (interspecific competition); where both population growth r 1 and r 2 decreases as Ten 1 increases and vice versa. In other words, each individual in X 1 population inhibits X 1 by i/G 1 and inhibits X two past α21/One thousand 2.

The Lotka–Volterra equations take a few assumptions under consideration:

1.

Intrinsic rate growth (r), competition coefficient (α), and carrying chapters (K) are all constants.

two.

Every individual within each population is identical.

3.

The populations are not allowed to diversify.

4.

The environment (habitat) is homogenous.

By evaluating the Lotka–Volterra equations, it is predicted that both competing species will coexist and that the competition between individuals of the same species (intraspecific) is weaker than the contest between the 2 species (interspecific). To sympathise the unlike scenarios predicted by the equations we must first define a cypher isocline represented by a directly line in population growth graphs. The naught isocline for each species states that at any given indicate along that line the species does not increment or decrease. This nada isocline is calculated by setting the growth rate of that species equal to zero (dTen/dt  =   0), and solve for X. Any indicate above the zero isocline indicates that the population of that species is decreasing; whatever point beneath the isocline that the population is increasing.

Where no interspecific contest exists, α12 or 10 two are equal to zero in eqn [3] and α21 or X one are equal to cypher in eqn [4] and both populations will grow sigmoidally every bit described past Verhulst–Peral logistic equation and both populations will accomplish their carrying capacity earlier stabilizing.

Figure 3 illustrates two extremes. The top line examines population X i, where on one extreme population ii is absent and population X ane is at its carrying capacity, X ane  = K 1. In the other extreme, species 10 ane is absent-minded and population X ii is at its carrying chapters where X 2  = K 112. The lesser line examines population X 2. On one extreme X 2 is at its carrying capacity (10 2  = Grand 2) since population X one is absent and X 2 is zero when population X 1 is at its conveying chapters (Ten 1  = Thousand 221). Betwixt the two lines whatsoever possible combination of the two species is possible. Same interpretation tin can exist made for the isocline of species X ii ( Figure 3 ).

Figure 3. Interspecific competition, species X i wins.

Placing the two lines on the same axis enables u.s. to predict to what extent each competing species affects the other. When the two isoclines do non cross, ane of the species will prevent the other species from occupying the aforementioned habitat. In Figure 3 species X 1 volition eventually crusade species X 2 to decline to zero (X i isocline is constitute in a higher place X 2 isocline and as 10 1 approaches Thou 1, Ten 2 approaches zero); Figure four shows the opposite situation where equally species X 2 approaches K 2, species X i approaches zero (X 2 isocline is plant above 10 1 isocline).

Figure four. Interspecific contest, species X 2 wins.

The Lotka–Volterra equations predict that coexistence of the two populations tin only accept place when the two isoclines cantankerous one some other. Four possible scenarios may event from interspecific competition of two species and the outcome depends on how the ii isoclines are in relation to one some other.

Scenario one ( Effigy 5 ). The ii isoclines do not cross and the isocline of population 10 1 is in a higher place that of X 2. Any point located below X ii isocline represents coexistence of both population of species and indication of both species increasing. Any point located above X 1 isocline represents both species decreasing. Points located between the 2 isoclines represent the X 1 below its isocline (increasing) and X 2 above its isocline (decreasing). The result of this scenario is population Ten 1 driving X two into extinction and increasing until 10 i  = Grand 1. This scenario is termed 'competitive exclusion of population X 2 by population 10 1'.

Figure 5. Competitive exclusion of population Ten ii past population X 1.

Scenario 2 ( Figure 6 ). This scenario is the exact reverse of scenario i where the two isoclines do not cross and the isocline of population Ten ii is above that of X 1. Any betoken located below X 1 isocline represents coexistence of both populations of species and indicates both species increasing. Any bespeak located above X 2 isocline represents both species decreasing. Points located between the two isoclines represent X 2 below its isocline (increasing) and X 1 higher up its isocline (decreasing). The result of this scenario is population X ii driving X 1 into extinction and increasing until X 2  = K 2. This scenario is termed 'competitive exclusion of population X 1 past population X two'.

Figure six. Competitive exclusion of population X ane by population 10 ii.

Scenario 3 ( Figure 7 ). Coexistence is possible but when both the populations are experiencing zero population growth. Under this scenario, the coexistence is unstable and will result in one of the species eventfully excluding the other. Nether this scenario X one will exclude X 2 when M one  > K 221 (conveying capacity of population X one is higher than that of X 2 divided past the contest coefficient) which will result in population 1 reaching its carrying capacity (10 one  = K 1) and population Ten 2 driven to extinction. In the effect that X 2 will reach its carrying chapters, since K 2  > K one12, the upshot would find species Ten one driven to extinction and 10 2  = G 2. If both species are at cypher growth (where isoclines cross), any environmental change may shift the species from this point leading ane of the population of species to extinction.

Effigy 7. Coexistence of two species when both species experience zip population growth.

Scenario 4 ( Effigy 8 ). The two isoclines cross each other and each species' carrying capacity is lower than the other's K divided past the competition coefficient. Every bit in scenarios ane and 2, a population increases higher up the isoclines and decreases below the isoclines, but any point plant betwixt the two isoclines represents coexistence of both populations of species leading to a stable situation where the isoclines cross. This situation infers that at the point of isoclines crossing, each individual inside the population is affected more than by individuals of its own population (intraspecific competition) as opposed to being limited by individuals of another population of species (interspecific competition).

Figure eight. Stable coexistence.

Competition Exclusion

When examining the possible scenarios outlined through the Lotka–Volterra equations, it is articulate that in scenarios 1 and 2 one species eventually drives the other to extinction. In 1934, a statement past Georgyi Frantsevitch Gause generated an of import debate in the ecological community regarding species competition. 'Gause's hypothesis (or Gause'southward Law)' states that due to competition, two similar species utilizing similar resources will scarcely ever be within the same niche and each will exist forced to utilize split food sources and modes of life that volition create a relative advantage over the competitor. Gause identified scenario 4 of the Lotka–Volterra scenarios equally the case of different niches which explain the possibility of stable coexistence of two competitive species. In 1960, Garrett Hardin defined the 'competitive exclusion principle' which stated that "consummate competitors cannot coexist", and further defined the Lotka–Volterra scenarios and conclusions.

Tilman's Model of Resource Competition

The Lotka–Volterra equations examine the effect of population size on interspecific contest and species coexistence but do non explore the mechanisms past which the effects of competition occur. David Tilman introduced a model that explored competition betwixt 2 species over limited resources. Resources were defined by Tilman as any factor consumed past the organism and its increase in the environment contributes to an increase in growth rate of the population. Tilman's kickoff key chemical element was examining the effect of ii essential resources to a population. A directly correlation exists betwixt the abundance of a resource and the population size. As the abundance of both resources increases, so does the population growth rate. When both resources decrease in availability, and then does the population size. A zero-growth isocline exists at the boundary betwixt population growth and turn down. A second key element in Tilman's model is resource consumption rate. When two populations of separate species consume a common resource, charge per unit of consumption differs between the species. If 2 species consume the aforementioned 2 resources with different consumption rates and the two zero-growth isoclines are superimposed, several outcomes are observed. In the commencement 2 scenarios ane of the species consumes both resources faster than the other species, driving the second species to extinction ( Figures 9a and 9b ). In the third scenario consumption rate is college for one resource by 1 species and higher for the other resource by the second species. As a result zero-growth isoclines cross, reaching equilibrium. This scenario may result in a stable or unstable equilibrium that depends on the specifics of the species' consumption rates ( Figure 10 ). Tilman specified that in order to predict the direction the equilibrium volition lean, we must know the charge per unit of supply too equally the rate of consumption of each of the resources.

Figure ix. Ii species utilize the same resource (resource competition). Each panel represents a scenario where one species consumes both resource faster than the other species causing the other species' extinction.

Figure 10. Ii species utilize the same resources. Each species consumes one resource faster than the other; this may result in a stable or unstable equilibrium.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780080454054006662

Protozoa

William D. Taylor , Robert W. Sanders , in Environmental and Classification of North American Freshwater Invertebrates (Third Edition), 2010

2. Interspecific Competition among Protozoa

Interspecific contest in laboratory cultures of ciliates has been studied by Gause [108] and numerous authors since. Gause provided articulate evidence that environmental factors influenced competitive outcomes.

The Lotka–Volterra equations predict that the winner of exploitative contest for resources in stable environments should exist the species with the greater K value, or conveying capacity, that is, the more efficient user of the resource. Nonetheless, K is usually measured as numbers, not biomass, so smaller species volition tend to have a college K. In bacterivorous ciliates, K does not predict the outcome of contest[181]. Only, in accordance with part of the r/M selection hypothesis, r, or the intrinsic charge per unit of increase, was negatively related to competitive power. (The r/K option hypothesis holds that r, which measures fitness in density-independent or stochastic environments, is negatively correlated amongst species with K, which measures fitness in stable environments.) Thus, the theorem seems to hold for ciliates, except that K, being highly correlated to body size, does not measure out competitive ability in stable environments. As noted in a higher place, r is also correlated with trunk size.

To factor out this correlation, the intrinsic rate of increase can be scaled as r/r due east, where r east is the expected r for a ciliate based on a regression of r on body volume[286]. Therefore, r/r east is a measure of how fast-growing, or r-selected, a ciliate is for its body size. This measure out was found to be negatively correlated with frequency of occurrence in the field and positively correlated with macronuclear ploidy[295]. This enquiry suggests that protozoan communities contain fast-growing but imperceptible opportunists every bit well as more conservative and specialized species that have more modest growth rates but stable populations. A similar moving-picture show emerged from a comparison of the population dynamics of a relatively fast and a relatively slow-growing species in the laboratory[182].

Field studies of exploitative contest in protozoa are few, as are field measurements of growth rates for comparison to laboratory estimates nether conditions of backlog food. Accordingly, nosotros know little most the relative importance of density dependence or independence. Work on planktonic protists suggests that field populations may feel periods of rapid growth and periods of lilliputian or no growth[47,293].

Maguire[191,192] found strong evidence for the exclusion of colonizing species in the genus Colpoda by Paramecium and associated species. Similarly, i member of the Paramecium aurelia species complex dominated sympatric sibling species in contest experiments and predominated in field collections[128,130]. The field population they studied appeared to be food-express. In contrast, sessile ciliates on an artificial substrate in a stream had high growth and mortality rates, implying petty competition[289]. The extensive studies on colonization by protozoa[35] provide indirect prove for species interactions, including contest.

Protozoa and metazoa sometimes compete for resources. Potential metazoan competitors include some rotifers and microcrustaceans, although it is hard in some cases to distinguish between the furnishings of competition and predation[312]. Likewise, the presence of predatory protozoa may touch the outcome of competition between other protozoan species[168].

Read full chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/commodity/pii/B9780123748553000030

PROTOZOA

William D. Taylor , Robert Due west. Sanders , in Environmental and Classification of North American Freshwater Invertebrates (Second Edition), 2001

2. Interspecific Competition among Protozoa

Interspecific contest in laboratory cultures of ciliates has been studied past Gause and numerous authors since. Gause (1934) provided articulate evidence that environmental factors could influence the outcome of competition.

The Lotka–Volterra equations predict that the winner of exploitative competition for resource in stable environments should be the species with the greater K or carrying capacity, that is, the more than efficient user of the resources. All the same, Yard is usually measured as numbers, non biomass, and then smaller species will tend to accept a college G. In his studies of competitive interactions of bacterivorous ciliates, Luckinbill (1979) confirmed that K did not predict the effect of competition. But, in accordance with part of the r/K selection hypothesis, the intrinsic rate of increase, r, was negatively related to competitive ability. [The r/K selection hypothesis holds that r, which measures fitness in density-independent or stochastic environments, is negatively correlated among species with K, which measures fitness in stable environments.] Thus the theorem seems to hold for ciliates, except that K, being highly correlated to body size, does not measure competitive power in stable environments. Equally noted above, r is also correlated with trunk size.

To factor out this correlation, Taylor (1978a,b) expressed intrinsic rate of increment as r/r e, where r e is the expected r for a ciliate based on a regression of r on trunk volume. Therefore, r/r e is a measure of how fast-growing or r-selected a ciliate is for its body size. This measure was constitute to exist negatively correlated with frequency of occurrence in the field, and positively correlated with macronuclear ploidy (Taylor and Shuter, 1981). This enquiry suggests that protozoan communities comprise fast-growing but imperceptible opportunists, also as more conservative, specialist species that have more modest growth rates, just stable populations. A similar picture emerged from a comparison of the population dynamics of a relatively fast and a relatively boring-growing species in the laboratory (Luckinbill and Fenton, 1978).

Field studies of exploitative competition in protozoa are few, as are field measurements of growth rates for comparison to laboratory estimates under atmospheric condition of excess food. Accordingly, we know little about the relative importance of density-dependence or independence. Work on planktonic protists suggests that field populations may experience periods of rapid growth and periods of niggling or no growth (Taylor and Johannsson, 1991; Carrick et al., 1992).

Maguire (1963a,b) found potent evidence for the exclusion of colonizing species in the genus Colpoda by Paramecium and associated species. Hairston and Kellerman (1965) and Hairston (1967) found that one member of the Paramecium aurelia species circuitous dominated sympatric sibling species in contest experiments and predominated in field collections. The field population they studied appeared to be nutrient-limited. In contrast, Taylor (1983b) constitute that sessile ciliates on an artificial substrate in a stream had high growth and mortality rates, implying fiddling contest. The all-encompassing studies on colonization by Cairns and co-workers (reviewed in Cairns and Yongue, 1977) provide indirect show for species interactions, including contest.

There is also competition betwixt protozoa and metazoa consuming similar resource. Potential metazoan competitors include some rotifers and crustaceans, although it is difficult in some cases to distinguish between the effects of competition and predation (Wickham and Gilbert, 1993). Likewise, the presence of predatory protozoa may affect the upshot of competition betwixt other protozoan species (Lawler, 1993).

Read total affiliate

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780126906479500041

Authorization☆

Helmut Hillebrand , in Encyclopedia of Ecology (2d Edition), 2019

Competition

Interspecific contest generally leads to a reduction of the contribution of poor competitors to the community, whereas superior competitors are able to gain potency. Thus, competition is a process strongly increasing dominance. The degree of competitive authorization depends on 2 factors: (i) on the asymmetry of the competition and (ii) the fourth dimension for superior species to develop their authorization (meet further below). The former aspect depends on the distribution of traits in an assemblage of competing species. If traits are similar, and thus competitive advantage of a certain trait low, dominance will be low. On the other hand, strong dominance tin can be expected if traits are skewed and competitive advantage of a sure trait is large ( Fig. 3). Competition of terrestrial plants for light is a lucid example for such an disproportionate competition for a unidirectional resources, as competitive success mainly depends on one single trait (establish height), and species growing higher are able to strongly dominate communities. Similarly, filamentous species highly boss many undisturbed assemblages of benthic algae (periphyton) as they are able to compete successfully for calorie-free and water cavalcade nutrients.

Fig. 3

Fig. iii. Conceptual depiction of processes affecting dominance. Grey ovals comprise biotic processes, white ovals abiotic constraints. Straight effects are represented by bold arrows, indirect effects past thin arrows. '+' and '−' characterize processes or traits increasing or decreasing dominance, respectively.

The skewed distribution of successful traits is as well the reason why anthropogenic input of fertilizers ofttimes leads to raise potency (Fig. iiA). Fertilization of terrestrial or aquatic environments (eutrophication) generally enhances the availability of i or few resources, only non of others. Thereby, the assortment of traits leading to competitive success narrows. A typical example is represented by the add-on of phosphorus (P) to lakes, which often leads to the authorisation of cyanobacteria species which demand a lot of P but are able to gear up atmospheric nitrogen (N).

More generally speaking, dominance increases when the number of potentially limiting nutrients decreases (Fig. three). It has been shown that the evenness of lake phytoplankton increases with increasing number of limiting nutrients. Similarly, the evenness of plant assemblages peaks at intermediate ratios of available N:P, just is low at low or loftier Due north:P ratios, respectively. Thus, a strong imbalance of Due north and P, which leads to limitation by either only N or simply P, leads to greater dominance of single species. Similarly, gradients in productivity implicitly are also gradients in the number of limiting resources. Very unproductive terrestrial ecosystems often are systems with stiff water shortage and dominated by species almost successful in accessing water. Very productive terrestrial ecosystems tend to exist express merely past light and over again single species with certain traits will boss. Therefore, dominance is least pronounced in mid-productive ecosystems.

Environmental harshness has often been supposed to reduce competition, which however has been falsified in recent years. Instead harsh conditions oftentimes require sure adaptations and the array of traits leading to survival and competitive success are too highly narrowed for organisms in extraordinarily harsh atmospheric condition. Thus dominance of single species is the full general case in these environments.

The second aspect, time of development, describes whether the competitive reward of a species can fully develop (Fig. 3). The gain of competitive authorization takes fourth dimension, while the extent of time strongly depends on the generation time of the organisms and the already-mentioned disproportion in competitive advantage. A highly superior fast-growing species will proceeds dominance more rapidly than a slow-growing species which has only a marginal competitive advantage compared to the cooccurring species. Competitive potency in competing microalgae tin can arise within days or weeks, whereas it may last years in long-lived copse or in ecological similar moss species in a bog.

Because the proceeds of competitive dominance takes time, spatial too every bit temporal heterogeneity may stop or even reverse this process. Spatial heterogeneity tin be provided past patchiness in resource supply or past biotic or abiotic architecture affecting competitive success. Patchy resource supply creates a mosaic of resource ratios and competitively advantageous strategies. Architecture ofttimes opens up refuges or represents the basis for the success of other life-forms (such as epiphytes). Temporal heterogeneity is often investigated in the form of fluctuations in resource availability, which also reduces competitive dominance or reverses competitive ranks of the interacting species.

Disturbances, divers as an irregular mortality-inducing effect, has been proposed equally a univariate explanation of dominance and variety patterns, well-nigh prominently by the intermediate disturbance hypothesis (IDH), which predicts highest species richness at disturbance regimes of intermediate frequency and intensity. Similarly, it predicts that low-disturbance ecosystems are dominated by competitively superior species, whereas high disturbance creates dominance by species able to cope with the extreme mortality in these systems. Dominance is thus expected to be low in mid-disturbance regimes, where dominance opens niche opportunities for species past creating patchiness of environmental weather condition allowing species to differentially limited life-history tradeoffs. Strong evidence has accumulated that how disturbance (and likewise consumption every bit a biotic mortality agent) affects community structure is coupled to the productivity of the system, which affects the rate of biomass accrual and dispersal and thus the speed of competitive displacement in a community (come across below).

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489111583

Indirect Effects in Ecology

V. Krivtsov , in Encyclopedia of Ecology, 2008

Most Ordinarily Studied Indirect Effects

Among a plethora of possible indirect furnishings, there are v that accept been studied most ordinarily. Their essence is depicted in Figure 1 and is briefly explained beneath.

Figure 1. Diagrams of the most commonly studied indirect affects. Direct furnishings are shown using solid lines, while indirect furnishings (just the effects relevant to the accompanying discussion are illustrated) using dotted lines. Interaction modification is illustrated using a dashed line. Numbers in the compartments are used solely for labeling to distinguish between unlike compartments, and do not chronicle to any kind of hierarchy. Also, the box sizes do not bear whatsoever relevance to the sizes or significance of the compartments drawn, and the relative size of the arrows relates neither to the effect's strength no to the preferential directionality. See further explanations in the text. (a) Interspecific contest; (b) credible competition; (c) trophic cascade; (d) indirect mutualism involving exploitative competition; (e) indirect mutualism involving interference competition; (f) interaction modification. Modified from Wootton JT (1994) The nature and consequences of indirect furnishings in ecological communities. Annual Review of Ecology and Systematics 25: 443–466.

Interspecific competition

Interspecific contest (likewise called exploitative competition) takes place whenever two (or several) species compete for the same resources. In Figure 1a , an increase in Component ane volition pb to the increased consumption of the shared resource (Component 2), and consequently to the decrease in a competitor (Component 3). Examples of this include, for example, two predators sharing the same prey, or 2 microbial species whose growth is limited by the availability of the same nutrient.

Apparent competition

Apparent competition occurs when ii species have a common predator. In Effigy 1b an abundant population of species 1 sustains a high-density population of predator ii, who, in turn, may limit the population of some other prey species 3. From practical point of view, it is worth noting here that this situation sometimes happens equally an unwanted result in biocontrol, when a biocontrol agent (species 2), specifically introduced to command a target (species one), may increase the gamble of a nontarget's (species 3) extinction.

Trophic cascades

Trophic cascades involve propagation of the upshot along a vertical trophic chain consisting of three or more than components connected past grazing or predation. In Figure 1c , an increase/subtract in Component 4 will pb to the decrease/increase in Component 3, increment/decrease in Component 2, and subtract/increase in Component 1. These effects are especially well studied in aquatic nutrient bondage (meet examples beneath), but take likewise been studied in terrestrial systems.

Information technology is worth pointing out, nonetheless, that the structure of real ecosystems hardly ever fits tidily into the concepts of unproblematic trophic levels (eastward.g., omnivory is widespread in nature), and trophic cascades, therefore, are often complicated by the interlinks within and among trophic levels (east.chiliad., in terrestrial ecosystems insectivorous birds prey on predatory, herbivorous, and parasitoid insects, and the resulting effect of birds on the primary producers and their damage by herbivory may, therefore, depend on the specific species and the conditions involved). In item, proper consideration of detritus contributions to the energy flows may prove the 'trophic cascade' simplification unsuitable, as the detritus compartment often has directly links to a number of trophic levels.

Indirect mutualism and commensalism

Indirect mutualism and commensalism involve a consumer–resource interaction coupled with either exploitative ( Figure 1d ) or interference ( Effigy 1e ) competition. For instance, starfish and snails reduce the abundance of mussels, a dominant space occupier, and increase the abundance of inferior sessile species. The presence of grazers on oyster farms in Australia increases oyster recruitment past removing algae, who otherwise preempt the available spaces. In Figure 1d , an increase in species one should lead to a decrease in species 2 and an increment in species 3. The latter positive effect would propagate upwards the right branch of the diagram, increasing the abundances of species 4 and 5. This state of affairs arises when, for example, planktivorous fish preferentially feeding on large zooplankton indirectly increase the abundance of small zooplankton. Cases involving interference contest are well known from, for example, the intertidal environment, where birds increment the affluence of acorn barnacles by consuming limpets that otherwise dislodge the young barnacles off the rock.

Interaction modification

Interaction modification occurs when the human relationship between a species pair is modified by a third species ( Figure 1f ). Examples include positive effects of macroalgae on zooplankton through interference with the hunting potential of fish and irresolute of a chemical's bioavailability due to the activeness of a species, when the chemic in question is of import for the functioning of another species (e.g., acids produced by one microbial population may increment bioavailability of compounds that are bound or unaccessible for some other microbial population).

It is worth pointing out that 'interaction modification' is often, and quite rightly, considered as a principally different type of indirect effect. Past coupling interaction modifications with other types of relationships (e.one thousand., trophic), one may get in at possibilities of numerous (including very circuitous) relationships. Ane of the more elementary of such combinations may be exemplified ( Figure ii ) with an indirect effect of grazers and certain agricultural practices on the population density of foxes (Vulpes vulpes) and the rodent Marmota bobac in Eastern Europe (V. Takarsky, personal communication): lower grazing rates pb to a denser and taller grass cover, enabling more than successful hunting of predators. Conversely, college grazing rates pb to a lower grass encompass, thus enhancing the detection of predators by the rodents. As a result, increase in grazing may accept an indirect positive effect on the Marmota bobac population, and an indirect negative effect on the population of foxes.

Figure 2. Diagram illustrating a positive indirect effect of grazing on Marmota bobac population resulting from a combination of consumer–resources relationships with an interaction-modification human relationship. See further explanations in the text.

It should also exist noted that some of the known examples of ammensalism and commensalism do actually fit in the description either of a simple interaction modification or interaction modification coupled with a number of tropic relationships. For instance, the bioavailability example described above has been quoted past Atlas and Bartha as an example of commensalism. If, yet, the chemical in question is non nutritional, only harmful for the second species, so the relationship fits the criteria for ammensalism. In a like vein, protocooperative and mutualistic relationships are easily envisaged from certain combinations of interaction modifications and tropic relationships.

It is worth pointing out that although the indirect relationships listed above are mainly studied in relation to pairs of biological species, they are applicable to a wider range of organization components. It should also be noted that many more types of indirect effects are easily envisaged from various possible combinations betwixt interacting compartments, and quite a few have indeed been observed in nature. For example, Menge distinguished 83 subtypes of indirect effects. However, an effort to exemplify every possible type of indirect effects would exist outside the scope of this article. The readers could easily construct, for example, many further types of indirect furnishings combining the most commonly studied ones depicted in Figure ane . In a real world, ecosystem components simultaneously take part in a multitude of interactions, and it is therefore appropriate to name information technology an interaction web. In fact, the number of possible kinds of indirect effects is probable to be limited only by the number of system components considered.

Read total affiliate

URL:

https://www.sciencedirect.com/science/commodity/pii/B9780080454054006935

Sea Urchins: Biology and Ecology

Timothy R. McClanahan , Nyawira A. Muthiga , in Developments in Aquaculture and Fisheries Scientific discipline, 2020

3.3 Interactions with fish

No interspecific contest between fish and Echinometra has been reported, other than the studies of Williams (1979, 1980, 1981), which institute low levels of aggression toward E. viridis. Consequently, interference competition between Echinometra and fish is probably non very common simply may occur with territorial damselfish in some reef ecosystems dominated by thickets of Acropora (Mapstone et al., 2007). Mapstone et al. (2007) report that the damselfish Stegastes nigricans and S. lividus may reduce E. mathaei numbers from their territories, and this regulates the erosion rates and affluence of Acropora thickets. At the intermediate E. mathaei population levels maintained past Stegastes, Due east. mathaei appears to promote the persistence of the Acropora thickets only not at the high or low ends of abundance. Resource competition may be mutual when Echinometra reaches loftier population levels, equally indicated by experiments that institute slightly increased herbivore fish abundance (damselfish and parrotfish) and grazing rates in areas where East. mathaei populations were reduced (McClanahan et al., 1994, 1996). The increased erosion rates by loftier numbers of Echinometra may likewise influence coral encompass and topographic complexity and the associated fish fauna.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/commodity/pii/B9780128195703000287

A Primer on Ecological Relationships among Freshwater Invertebrates

James H. Thorp , D. Christopher Rogers , in Field Guide to Freshwater Invertebrates of N America, 2011

C Physical Command of Communities

While interspecific competition and predation are present in all aquatic ecosystems to some extent, physical factors also play a role in determining the distribution and abundance of many invertebrates. The predominant view among ecologists is that physical factors play a greater just not exclusive role in controlling diversity within lotic than lentic systems. However, this consensus tends to ignore ephemeral pools, where the length of the hydroperiod is critical to ecosystem structure and function. Inside permanent systems, however, streams are probably more than subject to physical disturbances than lakes. Small streams and those in more arid regions experience more variable and less-predictable pulses of water (chosen flood pulses if the stage peak exceeds bankfull) following pelting events than do larger lotic systems. As menstruum pulses increase in severity, more of the substrate is moved, thereby disturbing the resident invertebrates and their food resource.

While flow pulses may reduce overall densities in a stream by disturbing substrates and washing species downstream or on to land, this physical disturbance also favors species able to reproduce chop-chop or seek shelter from the loftier currents. In this way, the physical disturbance may preclude a superior competitor from driving a subordinate to extinction and thereby allow greater species diversity in the community. Flood pulses likewise benefit species that need to reproduce on the flooded floodplains. Fish that are floodplain specialists especially benefit from regular, predictable, and extended flood pulses, but this may as well favor many invertebrates, though this has rarely been studied. Stream ecologists now believe that a natural flow regime of fluctuating flows is vital to the health of all lotic ecosystems.

Ecologists often focus on the furnishings of frequent flow pulses or annual inundation pulses, but the periodicity and degree of floods and droughts over periods of 1–100 years (menstruum history) and >100 years (flow regime) tin be very of import to aquatic and floodplain species of animals and plants.

Another physical disturbance is ice. Far northern wetlands frequently freeze to the lesser, but resident invertebrates typically have life history adaptations assuasive their survival. Streams less commonly freeze to the bottom, but whenever ice touches the lesser of a stream or air current-swept littoral zone of a lake, information technology can scour the bottom and eliminate all surface species. Stream communities will apace recolonize afterwards the thaw, just shallow lake bottoms may permanently show a difference in community composition within and beneath ice-scoured areas.

All benthic invertebrates are strongly affected by the size, texture, permeability, stability, limerick, and location of the substrate. Substrates provide sites for resting, food acquisition, reproduction, and evolution as well as refuges from predators and inhospitable concrete conditions.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780123814265000041

Food Specialization☆

Richard Svanbäck , Daniel I. Bolnick , in Encyclopedia of Ecology (Second Edition), 2019

Ecological opportunity

Both predation and interspecific competition can restrict populations from using specific resources or habitats. By ecological opportunity we mean other factors, that hinders niche divergence, such as patch size, and microhabitat/resource diversity. Equally larger patch sizes can concord more diverse prey populations habitat fragmentation may take large furnishings on among private diet variation. For example, it has been shown that habitat fragmentation in eustarine tidal wetlands leads to lower resource diversity consequently decreasing among individual diet variation in the gray snapper ( Lutjanus griseus), a predatory fish. Similarly, information technology has also been shown that among individual diet variation in the fruit bat, Rousettus aegyptiacus, was higher in bound when the number of establish species bearing fruits was also higher.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489109248

Many Footling Hammers: Ecological Direction of Crop-Weed Interactions

Matt Liebman , Eric R. Gallandt , in Environmental in Agronomics, 1997

1 Temperature Conditions

The outcome of interspecific competition between plants can be markedly afflicted by temperature conditions ( Pearcy et al, 1981; Flint and Patterson, 1983; Holt, 1988; Radosevich and Roush, 1990; Wall, 1993). Although very little research effort has been directed toward exploiting differential responses to temperature in weed management strategies, results of an experiment conducted by Weaver et al. (1988) indicate that more attention to this subject could be worthwhile. After quantifying differences betwixt tomato and four weed species in their germination and emergence responses to temperature, Weaver et al. (1988) predicted the sowing temperatures at which tomato would emerge before the weeds and consequently experience less contest from them. In addition to choosing planting times based on soil temperature equally a means of improving weed control, soil temperature might exist manipulated intentionally for weed management purposes. Significant differences tin occur, for example, between tillage systems in early season soil temperature characteristics (Johnson and Lowery, 1985; Cox et al, 1990; Dwyer et al, 1995) that bear upon crop, and probably, weed emergence.

Read full affiliate

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780123782601500105