Next Article in Journal
A Probabilistic Short-Term Water Demand Forecasting Model Based on the Markov Chain
Next Article in Special Issue
Food Web Responses to Artificial Mixing in a Small Boreal Lake
Previous Article in Journal
Water Policy of Korea for Supplying Safe Groundwater in Rural Areas
Previous Article in Special Issue
Transparency, Geomorphology and Mixing Regime Explain Variability in Trends in Lake Temperature and Stratification across Northeastern North America (1975–2014)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fish but Not Macroinvertebrates Promote Trophic Cascading Effects in High Density Submersed Plant Experimental Lake Food Webs in Two Contrasting Climate Regions

by
Carlos Iglesias
1,*,
Erik Jeppesen
2,3,4,
Néstor Mazzeo
1,5,
Juan Pablo Pacheco
1,
Franco Teixeira-de Mello
1,
Frank Landkildehus
2,
Claudia Fosalba
1,5,
Juan M. Clemente
1 and
Mariana Meerhoff
1,2
1
Departamento de Ecología y Gestión Ambiental, Centro Universitario Regional del Este, Universidad de la República, Maldonado 20000, Uruguay
2
Department of Bioscience, Aarhus University, DK-8600 Aarhus, Denmark
3
Department of Bioscience, Arctic Research Centre, 8000 Aarhus C, Denmark
4
Sino-Danish Centre for Education and Research, Beijing 100190, China
5
Facultad de Ciencias, Universidad de la República, Montevideo 11400, Uruguay
*
Author to whom correspondence should be addressed.
Water 2017, 9(7), 514; https://doi.org/10.3390/w9070514
Submission received: 28 May 2017 / Revised: 27 June 2017 / Accepted: 7 July 2017 / Published: 12 July 2017
(This article belongs to the Special Issue Ecological Responses of Lakes to Climate Change)

Abstract

:
Predators play a key role in the functioning of shallow lakes. Differences between the response of temperate and subtropical systems to fish predation have been proposed, but experimental evidence is scarce. To elucidate cascading effects produced by predators in contrasting climatic zones, we conducted a mesocosm experiment in three pairs of lakes in Uruguay and Denmark. We used two typical planktivorous-omnivorous fish species (Jenynsia multidentata + Cnesterodon decemmaculatus and Gasterosteus aculeatus + Perca fluviatilis) and one littoral omnivorous-predatory macroinvertebrate (Palaemonetes argentinus and Gammarus lacustris), alone and combined, in numbers resembling natural densities. Fish predation on zooplankton increased phytoplankton biomass in both climate zones, whereas the effects of predatory macroinvertebrates on zooplankton and phytoplankton were not significant in either climate zone. Macroinvertebrates (that freely colonized the sampling devices) were diminished by fish in both climate areas; however, periphyton biomass did not vary among treatments. Our experiments demonstrated that fish affected the structure of both planktonic and littoral herbivorous communities in both climate regions, with a visible positive cascading effect on phytoplankton biomass, but no effects on periphyton. Altogether, fish impacts appeared to be a strong driver of turbid water conditions in shallow lakes regardless of climatic zone by indirectly contributing to increasing phytoplankton biomass.

1. Introduction

Aquatic communities and food webs are typically structured by either available resources or fish predation [1], classically known as, respectively, bottom-up and top-down control [2]. However, the strength of these drivers in food webs may differ among lake habitats, such as the pelagic and littoral zone. Habitats and local food webs can be connected by diel or seasonal movements of fish, macroinvertebrates and zooplankton [3,4,5], particularly in shallow lakes.
In temperate lakes, the occurrence of planktivorous fish is known to induce major shifts in size distribution [6,7,8] and sometimes also in the behavior of zooplankton [9,10,11,12,13], indirectly influencing phytoplankton. By also consuming plant-attached macroinvertebrate grazers (e.g., snails), fish may indirectly enhance periphyton growth as evidenced by experiments [14,15,16] and field data [17]; however, not all investigations concur on this point [18].
Fish predation is apparently even stronger in warm lakes than in comparable temperate lakes [19,20,21]; the fish communities in warm lakes typically have higher densities, greater taxonomic and trophic diversity (with predominance of omnivores) and smaller body size [22,23]. Experimental studies have elucidated the effect of planktivorous fish predation on zooplankton communities in subtropical and Mediterranean shallow lakes [24,25,26,27]. Other studies including field surveys have focused on the effects of fish predation on zooplankton structure [27,28] and behavioral shifts [25,29,30,31] or the impact caused by predacious littoral macroinvertebrates [32]. A common feature is that predation on zooplankton in subtropical lakes is high [20,33] and that the zooplankton, therefore, is dominated by small cladocerans, nauplii and rotifers [29,34,35,36,37,38,39], resulting in a lower grazing pressure on phytoplankton in subtropical lakes than in similar lakes located in colder climates [40]. Moreover, Meerhoff and collaborators [20] found a substantially lower abundance of plant-attached macroinvertebrates and lower biomass of periphyton on artificial plants in subtropical than in similar temperate shallow lakes, arguably because the abundant fish in the former feed on and stir up the periphyton to a larger extent there than in temperate lakes. Hence, fish partly become primary consumers, implying that food webs are more truncated in the subtropics [20,41].
By affecting, among other processes, predator-prey interactions, submerged macrophytes may importantly influence the functioning of the whole lake [42,43]. The typically positive effect of submerged macrophytes on water clarity in temperate shallow lakes [44,45] is substantially reduced in warm lakes [46,47], not least because macrophytes do not act as proper refuges for large zooplankton [29,30]; instead, they host small fish [23,48,49], as well as potential predatory macroinvertebrates like the shrimp Palaemonetes argentinus [32], sometimes in extremely high densities [25].
In the present study, we experimentally addressed the role played by fish and predatory macroinvertebrates for the trophic structure in in-lake closed vegetated mesocosms run under different climate conditions. We focused on the trophic interactions among: (i) predators-zooplankton-phytoplankton, hypothesizing that fish and predatory macroinvertebrates indirectly promote phytoplankton growth through trophic cascades in both climatic zones, but more so in the subtropics; (ii) predators-plant-associated macroinvertebrates-periphyton, testing the hypothesis that trophic cascades promoted by fish and predatory macroinvertebrates would strongly promote periphyton growth in temperate systems. By contrast, in the subtropics, fish would not only consume plant-associated macroinvertebrates, but also periphyton, potentially masking the indirect increase in periphyton biomass caused by trophic cascades.

2. Methods

2.1. Study Area

We selected three shallow lakes (Zmax. <3 m in all cases) in Uruguay (30–35° S, Lakes Diario, Blanca and Nutrias) and in Denmark (55–57° N, Lakes Kogleaks, Stigsholm and Bølling), covering a gradient in submerged macrophyte cover (range: 0–70% plant volume inhabited (PVI), sensu [44]), salinity, pH, turbidity, total nutrients and phytoplankton chlorophyll-a (Chl-a) (Table 1). In each lake, we conducted a mesocosm experiment during summer (January–April in Uruguay and August–October in Denmark).

2.2. Experimental Design

We used 16 transparent cylindrical PVC enclosures (diameter 1.2 m) in each of the six lakes (Table 2). The water level within the enclosures varied between 0.8 and 1.1 m during the course of the experiment, approximating a total water volume of 1000 L. The mesocosms were kept open to the atmosphere and fixed by a metal ring penetrating ca. 0.30 m into the sediment, thus ensuring isolation from the outside water. A hard plastic ring and several rubber bands mounted to poles secured the top ca. 0.40 m above the water surface. We placed the mesocosms along both sides of a specially-constructed bridge that allowed us to take samples with minimum disturbance (Figure 1).
Prior to the establishment of the mesocosms, we carefully removed all natural vegetation from the area. During the setup, fish were prevented from entering the mesocosms by placing a small-sized mesh net (500 μm) on the bottom of each mesocosm that was removed before initiating the experiment, allowing a direct contact between sediments and water.
To each mesocosm, we introduced an artificial plant bed mimicking submerged plants (120 stems per module, PVI = 75%). The plants were made of “hairy” 1.0 m-long plastic pieces (originally green Christmas’ tree garlands) with an architecture similar to that of macrophytes such as Myriophyllum sp. or Cabomba sp. (as in [29]). The plants were introduced to the mesocosms two weeks before establishing the treatments to allow periphyton and invertebrate colonization.
Assigned treatments were F (only fish); F + INV (fish + predatory macroinvertebrates); INV (only predatory macroinvertebrates) and CON (no predators added); (details of the design are in Table 2). Since we focused on the potential cascading effects of an assemblage of predators, in each country, we used two typically abundant omnivorous/predatory fish species (“Rio de la Plata one-sided live-bearer” Jenynsia multidentata Jenyns (1942) and “ten spotted live-bearer” Cnesterodon decemmaculatus Jenyns (1942) in the subtropical lakes, and “three-spined stickleback” Gasterosteus aculeatus L. and “European perch” Perca fluviatilis L. in the temperate lakes) and one omnivorous/predatory macroinvertebrate species (Palaemonetes argentinus Nobili (1901) and Gammarus lacustris Sars (1863), in subtropical and temperate systems, respectively). Both the species and the abundances to be added resembled natural densities as described in the available literature [23,50,51] (Table 2), representing a rough fish biomass of 110 g of fresh weight in each treatment in both countries and a ca. 5-times higher predatory macroinvertebrate biomass per mesocosm in the subtropical than in the temperate systems (56 and 12 g, according to natural average densities, respectively). We did not select individuals for sex and used individuals smaller than 7.0 cm (standard length) to ensure predominance of zooplanktivory-invertivory feeding habits. In the case of the “live-bearers”, we avoided introducing pregnant females.

2.3. Sampling and Sample Analysis

We took water samples for physical and chemical analyses and measured in situ parameters (see Table 1) immediately prior to the allocation of treatments (T0) and seven weeks (49 days) later (TF) in all lakes, except for one where some mesocosms had been damaged during a storm (Lake Bølling, Denmark). Since we conducted an extra sampling campaign in Lake Bølling four weeks after T0, those results were considered adequate to represent the final conditions in the lake. We collected depth-integrated water samples for analyses of phytoplankton, zooplankton and nutrients, using a pump that integrated different depths and zones inside each mesocosm.
For quantitative analysis of zooplankton, we filtered 6–8 L water through a 50-μm mesh net and fixed the filtrate with acid Lugol. We performed counts according to [52] and classified the identified cladocerans as pelagic/free-swimming and benthic/plant-associated (following [29]); other zooplankters were categorized as rotifers, calanoid and cyclopoid copepods (both juvenile and adults) and nauplii.
We measured 20 individuals of all of the crustaceans in each sample from the initial and final sampling campaigns and determined the biomass of total zooplankton (dry weight (DW)) using the length/weight relationships of [53]. As an indirect measure of the zooplankton grazing impact on phytoplankton, we estimated the zooplankton to phytoplankton biomass ratio (zoo:phyt; sensu [54]). Phytoplankton Chl-a was measured spectrophotometrically following cold ethanol extraction [55]. For determination of the zoo:phyt ratio, we converted the Chl-a concentration to phytoplankton DW using a Chl-a:C ratio of 30 and a DW:C ratio of 2.2 [54].
We removed one “plant” from each mesocosm (avoiding the outer part for potential wall effects) and shook it inside a 250-μm mesh-sized net to collect the plant-associated macroinvertebrates. Another plant was used to determine the biomass of attached periphyton in the laboratory (as Chl-a [55]). The animals were preserved with alcohol (70%) for taxonomic identification and later classified according to their feeding habits in accordance with the literature [56].
Total macroinvertebrate biomass and biomasses of feeding groups were calculated using length/weight relationships from the literature when available (e.g., [56]) or were directly weighed and expressed in mg. Every individual present was measured to the nearest mm. Based on the literature, the feeding groups were defined as gatherers (some Odonata and some Chironomidae), shredders (Oligochaeta, some Chironomidae and Ephemeroptera), scrapers (Gastropoda), filterers (some Chironomidae and some Trichoptera), piercers (some Trichoptera and Coleoptera) and predators (some Odonata and some Chironomidae). The first four groups were considered potential periphyton consumers, and their biomasses (InvBiom) were used for further calculations of biomass ratios.
The invertebrate to periphyton biomass ratio (InvBiom:periphyt) was also calculated by dividing the sum of gatherers, shredders and scrapers invertebrate biomasses (mg/g plant) by periphyton biomass (μg/g plant) measured as Chl-a (Chl-a concentration converted to periphyton DW, using a Chl-a:C ratio of 30 and a DW:C ratio of 2.2 [54]). We recaptured the added fish and predatory macroinvertebrates (P. argentinus and G. lacustris) after the final sampling (TF) by electro-fishing and by vigorously sweeping a hand net inside the mesocosms.

2.4. Data Analysis

We performed one-way ANOVA analyses to test for initial (after 14 days’ incubation without any treatment) and final between-treatment differences (i.e., F, F + INV, INV and CON) in each lake, in the biomass and abundance of zooplankton, biomass of phytoplankton and periphyton and plant-associated macroinvertebrate abundance. When significant differences emerged, we used Tukey’s HSD post hoc test. Prior to the analyses, we tested for homoscedasticity and the normal distribution of residuals using Cochran’s C-test and visual inspection of fitted values, respectively. When needed, we transformed the data by square-root or 4th root to homogenize variances. For response variables for which analysis assumptions could not be made (i.e., invertebrate biomass and the InvBiom:periphyt ratio), we used the Kruskal–Wallis H-test.

3. Results

3.1. Pelagic Trophic Interactions

After the 49-day experimental period, we observed strong fish effects on zooplankton abundance (Figure 2) and phytoplankton biomass (Figure 3A,B) in both climatic zones. Fish, both alone and together with predatory macroinvertebrates (i.e., similar responses in the F and F + INV treatments), generated strong negative effects on zooplankton, whereas predatory macroinvertebrates alone (INV treatment) generally had no significant effects on any of the response variables (Figure 2).
In all of the Uruguayan lakes, we found lower cladoceran abundances in the presence of fish (one-way ANOVA tests, F3,12 = 4.7, p < 0.05; F3,12 = 4.4, p < 0.05 and F3,12 = 9.8, p < 0.01 for Lakes Blanca, Diario and Nutrias, respectively, Figure 2A,C,E) than in their absence. Daphnia spp. were absent from the fish mesocosms (F and F + INV), but occurred in some of the fishless treatments (INV and CON); however, their abundance fell at the end of the experiment from 43.0 ± 20.0 to 5.1 ± 1.1 and from 5.2 ± 1.3 to 2.1 ± 0.2 ind. L−1 + SE in Lakes Diario and Nutrias, respectively. The abundances of calanoid copepods were also significantly lower in the F and F + INV treatments in Lake Blanca (F3,12 = 8.4, p < 0.01) and Lake Diario (F3,12 = 11.9, p < 0.01). Copepod nauplii were negatively affected by fish only in Lake Blanca (F3,12 = 5.5, p < 0.05), whereas cyclopoid copepods never seemed to be affected by fish or predatory macroinvertebrates in the subtropical lakes.
In contrast, phytoplankton biomass (measured as Chl-a) was higher in the treatments with fish (F3,12 = 3.3, p < 0.05; F3,12 = 10.34, p < 0.01 and F3,12 = 4.3, p < 0.01 in Lakes Blanca, Diario and Nutrias, respectively, Figure 3C). Only in Lake Nutrias did predatory macroinvertebrates alone (INV) affect phytoplankton biomass, with similar effects as in the treatments with fish (F and F + INV; Figure 3C).
Similarly, in two Danish lakes, Lake Kogleaks and Lake Bølling, we found statistically significant effects of fish (both alone, F, and together with Gammarus lacustris, F + INV) on cladoceran abundances (F3,12 = 6.6, p < 0.01 and F3,12 = 8.2, p < 0.01 in Lake Kogleaks and Lake Bølling, respectively). In Lake Stigsholm, in contrast, cladoceran abundances decreased in all treatments. Contrary to the patterns observed in subtropical lakes, here, Daphnia followed the same patterns as total cladocerans. In the three Danish lakes, calanoid copepods occurred in abundances so low that no statistical analysis could be performed. Cyclopoid copepods and nauplii decreased significantly only in the treatments including fish and only in Lake Stigsholm (F and F + INV, F3,12 = 15.7, p < 0.001 and F3,12 = 4.3, p < 0.05) (Figure 2). The predatory macroinvertebrates alone did not affect the zooplankton here either.
Phytoplankton biomass was also higher in the treatments including fish in those lakes where zooplankton was affected (i.e., in Lake Kogleaks and Lake Bølling, F3,12 = 6.1, p < 0.001; F3,12 = 9.3, p < 0.01, respectively). In Lake Stigsholm, however, phytoplankton biomass did not differ among treatments (Figure 3D), despite presumed differences in grazing pressure (Figure 3B). In all lakes in both climatic zones, the potential grazing pressure of zooplankton (estimated as the zoo:phyt biomass ratio) decreased significantly in treatments where fish were present (F and F + INV), but not in the treatment with only predatory macroinvertebrates (Figure 3B). Despite similar trends among the treatments, the estimated grazing pressure varied greatly among lakes, being extremely low in all of the treatments in Lake Nutrias (Uruguay, Figure 3A) and relatively high in all of the treatments in Lake Kogleaks (Denmark, Figure 3B).

3.2. Littoral Trophic Interactions

The interaction patterns observed between predators, plant-associated macroinvertebrates and periphyton were different than those described above between predators, zooplankton and phytoplankton. Here, predators affected the plant-associated macroinvertebrate, but there was not a clear effect on periphyton. Fish (in both the F and F + INV treatments), but sometimes also predatory macroinvertebrates alone (the INV treatments) reduced the abundances of plant-associated macroinvertebrates in both climate zones (Figure 4A,B).
There were only a few differences in the presence and composition of the trophic groups of plant-associated macroinvertebrates between the regions; piercers were not registered in the temperate lakes. In terms of biomass, in both regions, gatherers dominated the community, followed by shredders (Table 3), and they exhibited responses to fish and invertebrate predation throughout the experiment.
The most frequent taxa occurring in the subtropical lakes were Oligochaeta, Chironomidae and Ostracoda, but also Hirudinea and Gasteropoda appeared. Lake Nutrias had a very low abundance of plant-associated macroinvertebrates during the entire study period, preventing meaningful statistical analyses. The density of plant-associated macroinvertebrates was, however, significantly higher in the absence of fish predators in the other two subtropical lakes (F3,12 = 14.0, p < 0.001 and F3,12 = 6.4, p < 0.01 in Lakes Blanca and Diario, respectively, Figure 4A). Fish effects were stronger than those of shrimp (as evidenced in the post hoc tests), which were only noticeable in Lake Blanca (Figure 4A).
In terms of abundance, Oligochaeta and Chironomidae dominated the plant-associated macroinvertebrate communities of the temperate lakes, but also Ostracoda were important in Lakes Kogleaks and Bølling. In Lake Stigsholm, also Trichoptera and Isopoda (e.g., Asellus sp.) occurred. In the three temperate lakes, the negative effects of fish on the density of plant-associated macroinvertebrates were significant (F3,12 = 4.3, p < 0.05; F3,8 = 4.3, p < 0.05 and F3,12 = 5.6, p < 0.05 in Lakes Kogleaks, Bølling and Stigsholm, respectively, Figure 4B). Moreover, in two of the three temperate systems, the effects of the predatory macroinvertebrates were similar to those of fish (as evidenced in the post hoc tests, Figure 4B).
Periphyton biomass increased at the end of the experiment. However, the treatments effects were not significant (Figure 4E,F). Moreover, the potential grazing pressure of plant-associated macroinvertebrates (estimated as the InvBiom:periphyt ratio) decreased, particularly so in Lakes Blanca (Figure 4G) and Bølling (Figure 4H), in the treatments where fish were present (F and F + INV), but not in the treatment with only predatory macroinvertebrates. Despite the similar trends experienced by treatments, the estimated potential grazing pressure varied greatly among the lakes, being extremely low in all the treatments in Lake Nutrias (Uruguay, Figure 4G) and very high in all the treatments in Lake Kogleaks (Denmark, Figure 4H).

4. Discussion

Our experiment highlights the role of predators (mainly fish) as a strong structuring factor in the food webs in shallow lakes, promoting phytoplankton development in both temperate and subtropical climates. During the course of the experiment, we found a significant decrease in phytoplankton biomass and an increase in zooplankton abundance and biomass only when fish were absent. By contrast, in the fish treatments, a cascading effect was revealed by a major reduction of the zooplankton grazing potential and a significant increase in phytoplankton biomass in both climatic zones. Our findings confirm the occurrence of a fish predation-driven structure of zooplankton [2], as suggested also in previous works in subtropical, e.g., [19,26,32,56,57,58,59,60], temperate, e.g., [61,62,63,64], and even sub-polar systems [65,66].
Moreover, our highly structured environment (high artificial PVI of ca. 75%, potentially offering refuge, as shown in [67]) and the availability of alternative food items (such as macroinvertebrates and periphyton attached to the artificial plants) did not counteract the negative effects of fish predation on zooplankton, as could have been expected a priori, but rather concurs with the hypothesis that alternative food sources may enhance fish predation on zooplankton [5]. We found that large-bodied zooplankton like Daphnia spp. were depleted by fish predation, suggesting that the fish effects were strong even at this high artificial plant PVI, regardless of climatic zone.
Positive cascading effects on water transparency at the whole-lake level following changes in the fish community are often observed under mesotrophic or eutrophic conditions, particularly in temperate climates [61,68,69]. Conversely, long-term effective control of phytoplankton biomass by zooplankton is rarely seen in similar lakes in the subtropics in whole-lake studies with the presence of natural fish communities [39,70,71] and has mainly been reported when fish disappear (e.g., with summer fish kills [33,60]). However, the three-level food web tested displayed the typical trophic cascade effects produced by planktivorous fish, resulting in enhanced phytoplankton biomass. This is a typical pattern in temperate eutrophic lake systems [43,72] and also represents the most commonly-observed response in subtropical lakes already in mesotrophic conditions [46]. The main reasons proposed for the observed zooplankton biomass and body size distribution in the subtropics are the typically high densities and preferential use of macrophytes as habitats by fish [19]. Moreover, omnivory seems to be the most common feeding habit of fish there [23,73,74]. Physiological constraints of cladocerans with increased temperatures have also been suggested as causes for the observed patterns in warm systems [75]. The evidence provided by our experimental study confirms these previous empirical suggestions, but in contrast to expectations, there was no clear indication of the proposed differences between temperate and subtropical food web responses [20].
Unexpectedly, in our experiment, neither the shrimp P. argentinus in the subtropics, nor the amphipod G. lacustris in the temperate lakes (unless together with fish) triggered a change in zooplankton abundance and biomass and on phytoplankton biomass, contradicting our expectations based on previous field studies (shrimp [76]; amphipod [51]). In a field and laboratory experiment, González-Sagrario and collaborators identified P. argentinus as a strong predator on zooplankton [32], but their experiment involved a seven-times higher predator density, a much lower plant PVI in their field mesocosms (ca. 10% vs. our 75%) and a much lower experimental volume in the laboratory than our experiments (0.6 L vs. 1000 L), and any or all of these three factors have likely influenced the hunting and feeding efficiency of the shrimp. Besides, in temperate systems, the lack of effects promoted by G. lacustris can also be explained by differences between experimental designs, as we used bottomless mesocosms, allowing the presence of alternative food resources. As an opportunistic feeder, G. lacustris tend to use the most easily available food resources [76,77,78], which was probably not the case for zooplankton in our design where other food items could become available.
Our second hypothesis focused on the behavior of the littoral pathway [4,79]. We did find a clear effect of fish predation on plant-associated macroinvertebrates in all lakes in both countries, as previously suggested for both temperate [15,80] and subtropical shallow lakes [20]. In two out of the three temperate lakes, macroinvertebrate predators (G. lacustris) negatively affected the abundance of plant-associated macroinvertebrates, while in the subtropics, the plant-associated macroinvertebrates were not affected by the added shrimps. These differential effects of macroinvertebrate predation revealed in our experiment constitute the strongest contrast between the food webs in the two regions, and they contradict the idea that shrimp may exert a stronger predation pressure than Gammarus (or other invertebrate predators like Chaoborus larvae [25,81,82]) as suggested from earlier field evidence [20,32].
Contrary to our expectations, however, we found no cascading effects of any of the predators on the biomass of periphyton, suggesting the absence of indirect cascading effects, as well as of direct (i.e., feeding) effects of the used fish or macroinvertebrate predators under our experimental conditions. In the temperate region, the lack of direct effects of fish on periphyton is relatively common in lake studies [83,84,85,86] and in mesocosm experiments [17,85]. In contrast, in the subtropics, we expected a direct negative effect of fish on periphyton, as has previously been found [17,25,86] due to the omnivorous feeding behavior and habitat preferences of most subtropical fish [23]. The absence of (negative) effects may reflect an oversimplification of our predator assemblages, particularly with a lower richness of fish under our experimental than under natural conditions, leading to greater availability of alternative, more nutritious food sources, such as zooplankton and macroinvertebrates. Moreover, the low mean densities of plant-associated macroinvertebrates in our study might be an effect of the short colonization phase and low accessibility to the mesocosms, notably of snails, which may have prevented indirect positive effects of fish on periphyton, as previously reported for temperate systems [17,85,87,88]. Besides, the extremely high PVI used (ca. 75%) would have provided a large habitat surface for periphyton development, potentially high enough to counteract and/or mask any fish feeding effect. Our results on the effects of macroinvertebrates alone and interactive effects with fish must therefore be interpreted with caution. Moreover, if intraguild predation is important [89], more complex experiments resembling natural taxonomic and functional diversity besides densities are needed before firm conclusions can be drawn on the littoral food webs under different climate zones.
Altogether, the experimental evidence presented here concurs with the idea that at whole-system level, planktivorous-omnivorous fish [20,86], but not predatory macroinvertebrates [20,32], might constitute a switch forward mechanism to establish and maintain turbid water conditions in shallow lakes regardless of climatic zone by significantly contributing to phytoplankton biomass development.

Acknowledgments

We are grateful to Anne Mette Poulsen for manuscript editing and to Tinna Christensen for refinement of the figures. We deeply thank Eti Levi, Kirsten Landkildehus Thomsen, Torben L. Lauridsen, Tommy Silberg, Klaire Houeix, Lúcia Lobão, Marcelo Guerrieri, Nihan Yazgan, Ping Zhong, Guillermo Goyenola, Soledad García, Nicolás Vidal, Natalia Barberán, Alejandra Kröger, Iván González-Bergonzoni, Alejandro D’Anatro, Malvina Masdeu, Mariana Vianna, Santiago Barberán and Daniella Agratti for valuable field assistance. The project was supported by the Ministry of Science, Technology and Innovation of Denmark. The projects EU-MARS (Managing Aquatic ecosystems and water Resources under multiple Stress) funded under the 7th EU Framework Programme, Theme 6 (Environment including Climate Change), Contract No. 603378 (http://www.mars-project.eu). Centre for Lake Restoration (“CLEAR”, a Villum Kann Rasmussen Centre of Excellence project) and Centre for Regional Change in the Earth System (CRES) supported Erik Jeppesen. Sistema Nacional de Investigadores (Agencia Nacional de Investigación e Innovación) supported Mariana Meerhoff, Carlos Iglesias and Néstor Mazzeo and Franco Teixeira de-Mello. Programa de Desarrollo de las Ciencias Básicas) supported Carlos Iglesias, Mariana Meerhoff and Néstor Masseo. Maestría en Ciencias Ambientales supported Néstor Mazzeo. This work was part a PhD scholarship at Aarhus University from the Danish Research Agency to Carlos Iglesias.

Author Contributions

Carlos Iglesias designed and run the experiment, processed zooplankton samples, analyzed data and wrote the manuscript draft. Erik Jeppesen and Mariana Meerhoff contributed to the experimental design, statistical analysis and revisions of the Manuscript. Claudia Fosalba, Franco Teixeira-de Mello, Juan Pablo Pacheco and Nestor Mazzeo contributed to the experimental run and revisions. Juan Maria Clemente processed macroinvertebrates and participated during sampling campaigns and revisions. All authors commented on the MS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Carpenter, S.R.; Kitchell, J.F.; Hodgson, J.R.; Cochran, P.A.; Elser, J.J.; Elser, M.M.; Lodge, D.M.; Kretchmer, D.; He, X.; von Ende, C.N. Regulation of lake primary productivity by food web structure. Ecology 1987, 68, 1863–1876. [Google Scholar] [CrossRef]
  2. McQueen, D.J.; Post, J.R.; Mills, E.L. Trophic relationships in freshwater pelagic ecosystems. Can. J. Fish. Aquat. Sci. 1986, 43, 1571–1581. [Google Scholar] [CrossRef]
  3. Burks, R.L.; Lodge, D.M.; Jeppesen, E.; Lauridsen, T.L. Diel horizontal migration of zooplankton: Costs and benefits of inhabiting littoral zones. Freshw. Biol. 2002, 47, 343–365. [Google Scholar] [CrossRef]
  4. Schindler, D.E.; Scheuerell, M.D. Habitat coupling in lake ecosystems. Oikos 2002, 98, 177–189. [Google Scholar] [CrossRef]
  5. Vander Zanden, M.J.; Vadeboncoeur, Y. Fishes as integrators of benthic and pelagic food webs in lakes. Ecology 2002, 83, 2152–2161. [Google Scholar] [CrossRef]
  6. Brooks, J.L.; Dodson, S.I. Predation, body size and composition of plankton. Science 1965, 150, 28–35. [Google Scholar] [CrossRef] [PubMed]
  7. Brucet, S.; Boix, D.; Nathansen, L.W.; Quintana, X.D.; Jensen, E.; Balayla, D.; Meerhoff, M.; Jeppesen, E. Effects of temperature, salinity and fish in structuring the macroinvertebrate community in shallow lakes: Implications for effects of climate change. PLoS ONE 2012, 7, e30877. [Google Scholar] [CrossRef] [PubMed]
  8. Hrbácek, J.; Dvorakova, M.; Korinek, V.; Procházkóva, L. Demonstration of the effect of the fish stock on the species composition of zooplankton and the intensity of metabolism of the whole plankton association. Verh. Int. Ver. Limnol. 1961, 14, 192–195. [Google Scholar]
  9. Burks, R.L.; Lodge, D.M. Cued in: Advances and opportunities in freshwater chemical ecology. J. Chem. Ecol. 2002, 28, 1901–1917. [Google Scholar] [CrossRef] [PubMed]
  10. Kairesalo, T.; Tátrai, I.; Luokkanen, E. Impacts of waterweed (elodea canadiensis michx) on fish-plankton interactions in the lake littoral. Verh. Int. Ver. Limnol. 1998, 26, 1846–1851. [Google Scholar]
  11. Lauridsen, T.L.; Lodge, D.M. Avoidance of daphnia magna by fish and macrophytes: Chemical cues and predator-mediated use of macrophyte habitat. Limnol. Oceanogr. 1996, 41, 794–798. [Google Scholar] [CrossRef]
  12. Romare, P.; Hansson, L.A. A behavioural cascade: Top-predator induced behavioural shifts in planktivorous fish and zooplankton. Limnol. Oceanogr. 2003, 48, 1956–1964. [Google Scholar] [CrossRef]
  13. Timms, R.M.; Moss, B. Prevention of growth of potentially dense phytoplankton populations by zooplankton grazing, in the presence of zooplanktivorous fish, in a shallow wetland ecosystem. Limnol. Oceanogr. 1984, 29, 472–486. [Google Scholar] [CrossRef]
  14. Brönmark, C.; Weisner, S.E.B. Indirect effects of fish community structure on submerged vegetation in shallow, eutrophic lakes: An alternativa mechanism. Hydrobiologia 1992, 243, 293–301. [Google Scholar] [CrossRef]
  15. Jones, J.I.; Waldron, S. Combined stable isotope and gut contents analysis of food webs in plant dominated, shallow lakes. Freshw. Biol. 2003, 48, 1396–1407. [Google Scholar] [CrossRef]
  16. Liboriussen, L.; Jeppesen, E. Temporal dynamics in epipelic, pelagic and epiphytic algal production in a clear and a turbid shallow lake. Freshw. Biol. 2003, 48, 418–431. [Google Scholar] [CrossRef]
  17. Jones, J.I.; Sayer, C.D. Does the fish-invertebrate-periphyton cascade precipitate plant loss in shallow lakes? Ecology 2003, 84, 2155–2167. [Google Scholar] [CrossRef]
  18. Bertolo, A.; Lacroix, G.; Lescher-Moutoué, F.; Cardinal-Legrand, C. Plankton dynamics in planktivore and piscivore dominated mesocosm. Arch. Hydrobiol. 2000, 147, 327–349. [Google Scholar] [CrossRef]
  19. Havens, K.E. Size structure and energetics in a plankton food web. Oikos 1998, 81, 346–358. [Google Scholar] [CrossRef]
  20. Meerhoff, M.; Clemente, J.M.; Teixeira de Mello, F.; Iglesias, C.; Pedersen, A.R.; Jeppesen, E. Can warm climate-related structure of littoral predator assemblies weaken the clear water state in shallow lakes? Glob. Chang. Biol. 2007, 13, 1888–1897. [Google Scholar] [CrossRef]
  21. Havens, K.E.; Pinto-Coelho, R.M.; Beklioğlu, M.; Christoffersen, K.S.; Jeppesen, E.; Lauridsen, T.L.; Mazumder, A.; Méthot, G.; Alloul, B.P.; Tavşanoğlu, U.N.; et al. Temperature effects on body size of freshwater crustacean zooplankton from greenland to the tropics. Hydrobiologia 2015, 743, 27–35. [Google Scholar] [CrossRef]
  22. Lazzaro, X. Do the trophic cascade hypothesis and classical biomanipulation approaches apply to tropical lakes and reservoirs? Verh. Int. Ver. Limnol. 1997, 26, 719–730. [Google Scholar]
  23. Teixeira-de Mello, F.; Meerhoff, M.; Pekcan-Hekim, Z.; Jeppesen, E. Substantial differences in littoral fish community structure and dynamics in subtropical and temperate shallow lakes. Freshw. Biol. 2009, 54, 1202–1215. [Google Scholar] [CrossRef]
  24. Boveri, M.; Quirós, R. Cascading trophic effects in pampean shallow lakes: Results of a mesocosm experiment using two coexisting fish species with different feeding strategies. Hydrobiologia 2007, 584, 215–222. [Google Scholar] [CrossRef]
  25. Iglesias, C.; Goyenola, G.; Mazzeo, N.; Meerhoff, M.; Rodó, E.; Jeppesen, E. Horizontal dynamics of zooplankton in subtropical lake blanca (uruguay) hosting multiple zooplankton predators and aquatic plant refuges. Hydrobiologia 2007, 584, 179–189. [Google Scholar] [CrossRef]
  26. Tavşanoğlu, Ü.N.; Cakiroğlu, A.I.; Erdogan, Ş.; Meerhoff, M.; Jeppesen, E.; Beklioğlu, M. Sediments, not plants, offer the preferred refuge for daphnia against fish predation in mediterranean shallow lakes: An experimental demonstration. Freshw. Biol. 2012, 57, 795–802. [Google Scholar]
  27. Sinistro, R. Top-down and bottom-up regulation of planktonic communities in a warm temperate wetland. J. Plankton Res. 2009, 32, 209–220. [Google Scholar] [CrossRef]
  28. Castro, B.B.; Marques, S.M.; Gonçalves, F. Habitat selection and diel distribution of the crustacean zooplankton from a shallow mediterranean lake during the turbid and clear water phases. Freshw. Biol. 2007, 52, 421–433. [Google Scholar] [CrossRef]
  29. Meerhoff, M.; Iglesias, C.; Teixeira-de Mello, F.; Clemente, J.M.; Jensen, E.; Lauridsen, T.L.; Jeppesen, E. Effects of habitat complexity on community structure and predator avoidance behaviour of littoral zooplankton in temperate versus subtropical shallow lakes. Freshw. Biol. 2007, 52, 1009–1021. [Google Scholar] [CrossRef]
  30. Meerhoff, M.; Fosalba, C.; Bruzzone, C.; Mazzeo, N.; Noordoven, W.; Jeppesen, E. An experimental study of habitat choice by Daphnia: Plants signal danger more than refuge in subtropical lakes. Freshw. Biol. 2006, 51, 1320–1330. [Google Scholar] [CrossRef]
  31. Tavşanoğlu, Ü.N.; Brucet, S.; Levi, E.E.; Bucak, T.; Bezirci, G.; Özen, A.; Johansson, L.S.; Jeppesen, E.; Beklioğlu, M. Size-based diel migration of zooplankton in mediterranean shallow lakes assessed from in situ experiments with artificial plants. Hydrobiologia 2015, 753, 47–59. [Google Scholar] [CrossRef]
  32. González-Sagrario, M.; Balseiro, E.; Ituarte, R.; Spivak, E. Macrophytes as refuge or risky area for zooplankton: A balance set by littoral predacious macroinvertebrates. Freshw. Biol. 2009, 54, 1042–1053. [Google Scholar] [CrossRef]
  33. Iglesias, C.; Mazzeo, N.; Meerhoff, M.; Lacerot, G.; Clemente, J.; Scasso, F.; Kruk, C.; Goyenola, G.; García-Alonso, J.; Amsinck, S.; et al. High predation is of key importance for dominance of small-bodied zooplankton in warm shallow lakes: Evidence from lakes, fish exclosures and surface sediments. Hydrobiologia 2011, 667, 133–147. [Google Scholar] [CrossRef]
  34. Branco, C.W.C.; Rocha, M.I.A.; Pinto, G.F.S.; Gomara, G.A.; Filippo, R.D. Limnological features of funil reservoir (R.J., Brazil) and indicator properties of rotifers and cladocerans of the zooplankton community. Lakes Reser. Res. Manag. 2002, 7, 87–92. [Google Scholar] [CrossRef]
  35. Crisman, T.L.; Beaver, J.R. Applicability of planktonic biomanipulation for managing eutrophication in the subtropics. Hydrobiologia 1990, 200/201, 177–185. [Google Scholar] [CrossRef]
  36. Dumont, H.J. On the diversity of the cladocera in the tropics. Hydrobiologia 1994, 272, 27–38. [Google Scholar] [CrossRef]
  37. Fernando, C.H. Zooplankton, fish and fisheries in tropical freshwaters. Hydrobiologia 1994, 272, 105–123. [Google Scholar] [CrossRef]
  38. García, P.R.; Nandini, S.; Sarma, S.S.S.; Valderrama, E.R.; Cuesta, I.; Hurtado, M.D. Seasonal variations of zooplankton abundance in the freshwater reservoir valle bravo (Mexico). Hydrobiolgia 2002, 467, 99–108. [Google Scholar] [CrossRef]
  39. Kruk, C.; Rodríguez-Gallego, L.; Meerhoff, M.; Quintans, F.; Lacerot, G.; Mazzeo, N.; Scasso, F.; Paggi, J.C.; Peeters, E.T.H.M.; Scheffer, M. Determinants of biodiversity in subtropical shallow lakes (Atlantic coast, Uruguay). Freshw. Biol. 2009, 54, 2628–2641. [Google Scholar] [CrossRef]
  40. Meerhoff, M.; Teixeira-de Mello, F.; Kruk, C.; Alonso, C.; Gonzalez-Bergonzoni, I.; Pacheco, J.P.; Arim, M.; Beklioglu, M.; Brucet, S.; Goyenola, G.; et al. Environmental warming in shallow lakes: A review of effects on community structure as evidenced from space-for-time substitution approach. Adv. Ecol. Res. 2012, 46. [Google Scholar] [CrossRef]
  41. Iglesias, C.; Meerhoff, M.; Johansson, L.S.; González-Bergonzoni, I.; Mazzeo, N.; Pacheco, J.P.; Teixeira-de Mello, F.; Goyenola, G.; Lauridsen, T.L.; Søndergaard, M.; et al. Stable isotope analysis confirms substantial differences between subtropical and temperate shallow lake food webs. Hydrobiologia 2017, 784, 111–123. [Google Scholar] [CrossRef]
  42. Carpenter, S.R.; Lodge, D.M. Effects of submersed macrophytes on ecosystem processes. Aquat. Bot. 1986, 26, 341–370. [Google Scholar] [CrossRef]
  43. Jeppesen, E.; Jensen, J.P.; Søndergaard, M.; Lauridsen, T.; Pedersen, L.J.; Jensen, L. Top-down control in freshwater lakes: The role of nutrient state, submerged macrophytes and water depth. Hydrobiologia 1997, 342, 151–164. [Google Scholar] [CrossRef]
  44. Canfield, D.E., Jr.; Shireman, J.V.; Colle, D.E.; Haller, W.T.; Watkins, C.E.I.; Maceina, M.J. Prediction of chlorophyll a concentrations in florida lakes: Importance of aquatic macrophytes. Can. J. Fish. Aquat. Sci. 1984, 41, 497–501. [Google Scholar] [CrossRef]
  45. Scheffer, M.; Hosper, S.H.; Meijer, M.-L.; Moss, B.; Jeppesen, E. Alternative equilibria in shallow lakes. Trends Ecol. Evol. 1993, 8, 275–279. [Google Scholar] [CrossRef]
  46. Jeppesen, E.; Meerhoff, M.; Jakobsen, B.A.; Hansen, R.S.; Søndengaard, M.; Jensen, J.P.; Lauridsen, T.; Mazzeo, N.; Branco, C.C. Restoration of shallow lakes by nutrient control and biomanipulation—The successful strategy varies with lake size and climate. Hydrobiologia 2007, 581, 269–285. [Google Scholar] [CrossRef]
  47. Kosten, S.; Lacerot, G.; Jeppesen, E.; da Motta Marques, D.; van Nes, E.; Mazzeo, N.; Scheffer, M. Effects of submerged vegetation on water clarity across climates. Ecosystems 2009, 12, 1117–1129. [Google Scholar] [CrossRef]
  48. Conrow, R.; Zale, A.V.; Gregory, R.W. Distributions and abundances of early stages of fishes in a florida lake dominated by aquatic macrophytes. Trans. Am. Fish. Soc. 1990, 119, 521–528. [Google Scholar] [CrossRef]
  49. Meerhoff, M.; Mazzeo, N.; Moss, B.; Rodríguez-Gallego, L. The structuring role of free-floating versus submerged plants in a subtropical shallow lake. Aquat. Ecol. 2003, 37, 377–391. [Google Scholar] [CrossRef]
  50. Liboriussen, L.; Jeppesen, E.; Bramm, M.E.; Lassen, M.F. Periphyton-macroinvertebrate interactions in light and fish manipulated enclosures in a clear and a turbid shallow lake. Aquat. Ecol. 2005, 39, 23–39. [Google Scholar] [CrossRef]
  51. Wilhelm, F.M.; Schindler, D.W. Effects of gammarus lacustris (crustacea: Amphipoda) on plankton community structure in an alpine lake. Can. J. Fish Aquat. Sci. 1999, 56, 1401–1408. [Google Scholar] [CrossRef]
  52. Paggi, J.C.; de Paggi, S.J. Primeros estudios sobre el zooplancton de las aguas lóticas del paraná medio. Physis 1974, 33, 94–114. [Google Scholar]
  53. Dumont, H.J.; Van de Velde, I.; Dumont, S. The dry weight estimate of biomass in a selection of cladocera, copepoda and rotifera from the plankton, periphyton and benthos of continental waters. Oecologia 1975, 19, 75–97. [Google Scholar] [CrossRef] [PubMed]
  54. Jeppesen, E.; Søndergaard, M.; Kanstrup, E.; Petersen, B.; Henriksen, R.B.; Hammershøj, M.; Mortensen, E.; Jensen, J.P.; Have, A. Does the impact of nutrients on the biological structure and function of brackish and freshwater lakes differ? Hydrobiologia 1994, 275–276, 15–30. [Google Scholar] [CrossRef]
  55. Jespersen, A.M.; Christoffersen, K. Measurements of chlorophyll-a from phytoplankton using ethanol as extraction solvent. Arch. Hydrobiol. 1987, 109, 445–454. [Google Scholar]
  56. Merrit, W.; Cummings, K. An Introduction to the Aquatic Insects of North America; Kendal/Hunt: Dubuque, Iowa, 1984. [Google Scholar]
  57. Fontanarrosa, M.; Chaparro, G.; de Tezanos Pinto, P.; Rodriguez, P.; O’Farrell, I. Zooplankton response to shading effects of free-floating plants in shallow warm temperate lakes: A field mesocosm experiment. Hydrobiologia 2010, 646, 231–242. [Google Scholar] [CrossRef]
  58. Iglesias, C.; Mazzeo, N.; Goyenola, G.; Fosalba, C.; Teixeira-de Mello, F.; Garcia, S.; Jeppesen, E. Field and experimental evidence of the effect of jenynsia multidentata, a small omnivorous-planktivorous fish, on the size distribution of zooplankton in subtropical lakes. Freshw. Biol. 2008, 53, 1797–1807. [Google Scholar] [CrossRef]
  59. Mazzeo, N.; Iglesias, C.; Teixeira-de Mello, F.; Borthagaray, A.; Fosalba, C.; Ballabio, R.; Larrea, D.; Vilches, J.; García, S.; Pacheco, J.; et al. Trophic cascade effects of Hoplias malabaricus (Characiformes, Erythrinidae) in subtropical lakes food webs: A mesocosm approach. Hydrobiologia 2010, 644, 325–335. [Google Scholar] [CrossRef]
  60. Nagdali, S.S.; Gupta, P.K. Impact of mass mortality of a mosquito fish, gambussia affinis, on the ecology of a freshwater eutrophic lake (Lake Naini Tal, India). Hydrobiologia 2002, 468, 45–52. [Google Scholar] [CrossRef]
  61. Carpenter, S.R.; Kitchell, J.F. The Trophic Cascade in Lakes; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar]
  62. Gyllström, M.; Hansson, L.A.; Jeppesen, E.; García-Criado, F.; Gross, E.; Irvine, K.; Kairesalo, T.; Kornijow, R.; Miracle, M.R.; Nykänen, M.; et al. The role of climate in shaping zooplankton communities of shallow lakes. Limnol. Oceanogr. 2005, 50, 2008–2021. [Google Scholar] [CrossRef]
  63. Moss, B.; Stephen, D.; Balayla, D.M.; Bécares, E.; Collings, S.E.; Fernández-Aláez, C.; Fernández-Aláez, M.; Ferriol, C.; García, P.; Gomá, J.; et al. Continental-scale patterns of nutrient and fish effects on shallow lakes: Synthesis of a pan-european mesocosm experiment. Freshw. Biol. 2004, 49, 1633–1649. [Google Scholar] [CrossRef]
  64. Vakkilainen, K.; Kairesalo, T.; Hietala, J.; Balayla, D.M.; Becares, E.; Van de Bund, W.J.; Van Donk, E.; Fernández-Aláez, M.; Gyllström, M.; Hansson, L.A. Response of zooplankton to nutrient enrichment and fish in shallow lakes: A pan-european mesocosm experiment. Freshw. Biol. 2004, 49, 1619–1632. [Google Scholar] [CrossRef]
  65. Hayden, B.; Myllykangas, J.P.; Rolls, R.J.; Kahilainen, K.K. Climate and productivity shape fish and invertebrate community structure in subarctic lakes. Freshw. Biol. 2017, 62, 990–1003. [Google Scholar] [CrossRef]
  66. Skoglund, S.; Knudsen, R.; Amundsen, P. Selective predation on zooplankton by pelagic arctic charr, salvelinus alpinus, in six subarctic lakes. J. Ichthyol. 2013, 53, 849–855. [Google Scholar] [CrossRef]
  67. Schriver, P.; Bøgestrand, J.; Jeppesen, E.; Søndergaard, M. Impact of submerged macrophytes on fish-zooplankton-phytoplankton interactions: Large-scale enclosure experiments in a shallow eutrophic lake. Freshw. Biol. 1995, 33, 255–270. [Google Scholar]
  68. Ekvall, M.K.; Urrutia-Cordero, P.; Hansson, L.-A. Linking cascading effects of fish predation and zooplankton grazing to reduced cyanobacterial biomass and toxin levels following biomanipulation. PLoS ONE 2014, 9, e112956. [Google Scholar] [CrossRef] [PubMed]
  69. Jeppesen, E.; Lauridsen, T.L.; Kairesalo, T.; Perrow, M. Impact of submerged macrophytes on fish-zooplankton interactions in lakes. In The Structuring Role of Submerged Macrophytes in Lakes; Jeppesen, E., Søndergaard, M., Søndergaard, M., Christoffersen, K., Eds.; Springer: New York, NY, USA, 1997; Volume 131, pp. 91–114. [Google Scholar]
  70. Gelós, M.; Teixeira-de Mello, F.; Goyenola, G.; Iglesias, C.; Fosalba, C.; García-Rodríguez, F.; Pacheco, J.; García, S.; Meerhoff, M. Seasonal and diel changes in fish activity and potential cascading effects in subtropical shallow lakes with different water transparency. Hydrobiologia 2010, 646, 173–185. [Google Scholar] [CrossRef]
  71. Pacheco, J.; Iglesias, C.; Meerhoff, M.; Fosalba, C.; Goyenola, G.; Teixeira-de Mello, F.; García, S.; Gelós, M.; García-Rodríguez, F. Phytoplankton community structure in five subtropical shallow lakes with different trophic status (Uruguay): A morphology-based approach. Hydrobiologia 2010, 646, 187–197. [Google Scholar] [CrossRef]
  72. Persson, L.; Anderson, G.; Hamrin, S.F.; Johansson, L. Predation regulation and primary production a long the productivity gradient of temperate lake ecosystem. In Complex Interactions in Lake Communities; Carpenter, S.E., Ed.; Springer: New York, NY, USA, 1988; pp. 45–65. [Google Scholar]
  73. Winemiller, K.O. Spatial and temporal variation in tropical fish trophic networks. Ecol. Monogr. 1990, 60, 331–367. [Google Scholar] [CrossRef]
  74. González-Bergonzoni, I.; Meerhoff, M.; Davidson, T.; Teixeira-de Mello, F.; Baattrup-Pedersen, A.; Jeppesen, E. Meta-analysis shows a consistent and strong latitudinal pattern in fish omnivory across ecosystems. Ecosystems 2012, 15, 492–503. [Google Scholar] [CrossRef]
  75. Weetman, D.; Atkinson, D. Evaluation of alternative hypotheses to explain temperature-induced life history shifts in daphnia. J. Plankton Res. 2004, 26, 107–116. [Google Scholar] [CrossRef]
  76. Collins, P.A.; Paggi, J.C. Feeding ecology of Macrobrachium borellii (Nobili) (Decapoda: Palaemonidae) in the flood valley of the river paraná, argentina. Hydrobiologia 1998, 362, 21–30. [Google Scholar] [CrossRef]
  77. Felten, V.; Tixier, G.; Guerold, F.; De Billy, V.D.C.; Dangles, O. Quantification of diet variability in a stream amphipod: Implications for ecosystem functioning. Fundam. Appl. Limnol. 2008, 170, 303–313. [Google Scholar] [CrossRef]
  78. Hellmann, C.; Worischka, S.; Mehler, E.; Becker, J.; Gergs, R.; Winkelmann, C. The trophic function of Dikerogammarus villosus (Sowinsky, 1894) in invaded rivers: A case study in the Elbe and Rhine. Aquat. Invasions 2015, 10, 385–397. [Google Scholar] [CrossRef]
  79. Vadeboncoeur, Y.; Vander Zanden, M.J.; Lodge, D.M. Putting the lake back together: Reintegrating benthic pathways into lake food web models. Bioscience 2002, 52, 44–54. [Google Scholar] [CrossRef]
  80. Diehl, S.; Kornijów, R. Influence of submerged macrophytes on trophic interactions among fish and macroinvertebrates. In The Structuring Role of Submerged Macrophytes in Lakes; Jeppesen, E., Søndergaard, M., Søndergaard, M., Christoffersen, K., Eds.; Springer: New York, NY, USA, 1997; Volume 131, pp. 24–46. [Google Scholar]
  81. Arcifa, M.S. Feeding habits of chaoboridae larvae in a tropical brazilian reservoir. Rev. Bras. Biol. 2000, 60, 591–597. [Google Scholar] [CrossRef]
  82. Castilho-Noll, M.S.M.; Arcifa, M.S. Mesocosm experiment on the impact of invertebrate predation on zooplankton of a tropical lake. Aquat. Ecol. 2007, 41, 587–598. [Google Scholar] [CrossRef]
  83. Bécares, E.; Gomá, J.; Fernández-Aláez, M.; Fernández-Aláez, C.; Romo, S.; Miracle, M.R.; Ståhl-Delbanco, A.; Hansson, L.-A.; Gyllström, M.; Van de Bund, W.J. Effects of nutrients and fish on periphyton and plant biomass across a european latitudinal gradient. Aquat. Ecol. 2008, 42, 561–574. [Google Scholar] [CrossRef]
  84. Hansson, L. Factors regulating periphytic algal biomass. Limnol. Oceanogr. 1992, 37, 322–328. [Google Scholar] [CrossRef]
  85. Brönmark, C.; Vermaat, J.E. Complex fish-snail-epiphyton interactions and their effects on submerged freshwater macrophytes. In The Structuring Role of Submerged Macrophytes in Lakes; Jeppesen, E., Søndergaard, M., Søndergaard, M., Christoffersen, K., Eds.; Springer: New York, NY, USA, 1997; pp. 47–68. [Google Scholar]
  86. Jeppesen, E.; Meerhoff, M.; Holmgren, K.; González-Bergonzoni, I.; Teixeira-de Mello, F.; Declerck, S.; De Meester, L.; Søndergaard, M.; Lauridsen, T.; Bjerring, R.; et al. Impacts of climate warming on lake fish community structure and potential effects on ecosystem function. Hydrobiologia 2010, 646, 73–90. [Google Scholar] [CrossRef]
  87. Cao, Y.; Li, W.; Jeppesen, E. The response of two submerged macrophytes and periphyton to elevated temperatures in the presence and absence of snails: A microcosm approach. Hydrobiologia 2014, 738, 49–59. [Google Scholar] [CrossRef]
  88. Mahdy, A.; Hilt, S.; Filiz, N.; Beklioğlu, M.; Hejzlar, J.; Özkundakci, D.; Papastergiadou, E.; Scharfenberger, U.; Šorf, M.; Stefanidis, K.; et al. Effects of water temperature on summer periphyton biomass in shallow lakes: A pan-european mesocosm experiment. Aquat. Sci. 2015, 77, 499–510. [Google Scholar] [CrossRef]
  89. Hart, D. Intraguild predation, invertebrate predators, and trophic cascades in lake food webs. J. Theor. Biol. 2002, 218, 111–128. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Eight plastic enclosures were fixed to each side of a bridge, and an artificial plant bed was introduced to all enclosures two weeks before sampling to allow periphyton and macroinvertebrates to colonize the plastic structures.
Figure 1. Eight plastic enclosures were fixed to each side of a bridge, and an artificial plant bed was introduced to all enclosures two weeks before sampling to allow periphyton and macroinvertebrates to colonize the plastic structures.
Water 09 00514 g001
Figure 2. Effects of experimental treatments on the abundance of zooplankton in subtropical (pannels A, C and E) and temperate (panels B, D and F)) lakes. Dotted and horizontal lines represent the initial conditions (just before treatment assignment and 15 days after fish exclusion), while columns show the final conditions (49 days after fish and macroinvertebrate addition). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axes. Letters indicate groups based on Tukey’s post hoc tests when ANOVA tests were significant.
Figure 2. Effects of experimental treatments on the abundance of zooplankton in subtropical (pannels A, C and E) and temperate (panels B, D and F)) lakes. Dotted and horizontal lines represent the initial conditions (just before treatment assignment and 15 days after fish exclusion), while columns show the final conditions (49 days after fish and macroinvertebrate addition). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axes. Letters indicate groups based on Tukey’s post hoc tests when ANOVA tests were significant.
Water 09 00514 g002
Figure 3. Effects of treatments on potential zooplankton grazing pressure (A,B) and phytoplankton biomass (C,D) in subtropical (left) and temperate (right) lakes. Dotted and horizontal lines represent initial values (just before treatment assignment and 15 days after fish exclusion), while columns depict final values (49 days after fish and invertebrate addition). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axis. Letters indicate groups according to Tukey’s post hoc tests when ANOVA tests were significant.
Figure 3. Effects of treatments on potential zooplankton grazing pressure (A,B) and phytoplankton biomass (C,D) in subtropical (left) and temperate (right) lakes. Dotted and horizontal lines represent initial values (just before treatment assignment and 15 days after fish exclusion), while columns depict final values (49 days after fish and invertebrate addition). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axis. Letters indicate groups according to Tukey’s post hoc tests when ANOVA tests were significant.
Water 09 00514 g003
Figure 4. Density (A,B) and biomass (C,D) of plant-associated macroinvertebrates, periphyton biomass (E,F) and potential macroinvertebrates grazing pressure (G,H) in subtropical (left) and temperate (right) lakes. Dotted and horizontal lines represent initial values (15 days after fish exclusion and just before treatment assignment), while columns represent final conditions (49 days after treatment additions). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axes. Letters indicate groups according to Tukey’s post hoc tests when ANOVA tests were significant.
Figure 4. Density (A,B) and biomass (C,D) of plant-associated macroinvertebrates, periphyton biomass (E,F) and potential macroinvertebrates grazing pressure (G,H) in subtropical (left) and temperate (right) lakes. Dotted and horizontal lines represent initial values (15 days after fish exclusion and just before treatment assignment), while columns represent final conditions (49 days after treatment additions). F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Error bars represent ±1 standard error. Please note the different scales on the y-axes. Letters indicate groups according to Tukey’s post hoc tests when ANOVA tests were significant.
Water 09 00514 g004
Table 1. Main limnological parameters of the lakes where the experiments were conducted. Shaded = subtropical Uruguayan lakes (30–35° S) and non-shaded = temperate Danish lakes (55–57° N). Shown are mean temperature, dissolved oxygen concentration (DO2), pH, water turbidity (Nephelometric Turbidity Unit, NTU) and concentration of phytoplankton chlorophyll-a (Chl-a) at the beginning (start) and at the end of the experiments. Distinctive traits of the lakes, including submerged plant percentage of volume inhabited (PVI), total nitrogen (TN) and total phosphorus (TP) concentrations, are also provided.
Table 1. Main limnological parameters of the lakes where the experiments were conducted. Shaded = subtropical Uruguayan lakes (30–35° S) and non-shaded = temperate Danish lakes (55–57° N). Shown are mean temperature, dissolved oxygen concentration (DO2), pH, water turbidity (Nephelometric Turbidity Unit, NTU) and concentration of phytoplankton chlorophyll-a (Chl-a) at the beginning (start) and at the end of the experiments. Distinctive traits of the lakes, including submerged plant percentage of volume inhabited (PVI), total nitrogen (TN) and total phosphorus (TP) concentrations, are also provided.
DiarioBlancaNutriasKogleaksStigsholmBølling
startendstartendStartendstartendstartendstartEnd
Temperature (oC)28.020.522.521.427.020.216.68.315.110.414.713.9
DO2 (mg L−1)9.77.06.710.47.48.45.79.110.59.38.77.6
pH8.67.58.47.616.145.86.97.58.47.57.17.2
Conductivity (mS cm−1)5666173183167582595542210185116116
Turbidity (NTU)6.32014.218.226.339.712.42.13.63.914.717.3
Phyto. Chl-a (µg L−1)15.79.646.556.115.74.78.94.711.48.28.410.7
Distinctive traitsSlightly brackishCyanobacteria bloomsHumicSlightly brackishCyanobacteria bloomsHumic
PVI (%)>75%<25%0%ca. 50%ca. 50%0%
TN (µg L−1)9701391.5670233022751600
TP (µg L−1)89.265.9122.5214.255182.8
Table 2. Fish and macroinvertebrate predator species and numbers added to the mesocosms in the treatments: F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). F and F + INV treatments had individuals of both fish species (fish densities at the beginning, expressed as ind. m−2, are given in parenthesis). For each species, biomass was estimated by multiplying introduced densities by average weight and expressed in grams of fresh weight. The selected densities of predators were based on information in the literature: * [23]; + [50]; ¤ [51].
Table 2. Fish and macroinvertebrate predator species and numbers added to the mesocosms in the treatments: F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). F and F + INV treatments had individuals of both fish species (fish densities at the beginning, expressed as ind. m−2, are given in parenthesis). For each species, biomass was estimated by multiplying introduced densities by average weight and expressed in grams of fresh weight. The selected densities of predators were based on information in the literature: * [23]; + [50]; ¤ [51].
Fish Species 1Fish Species 2Macroinvertebrates
SubtropicalNameCnesterodon decemmaculatusJenynsia multidentataPalaemonetes argentinus
Density50 (42) *40 (33) *120 (100) *
Biomass26.580.856
TemperateNameGasterosteus aculeatusPerca fluviatilisGammarus lacustris
Density12 (10) +6 (5) *240 (200) ¤
Biomass29.281.812
Table 3. Trophic group biomass (mg g−1 plant) of plant-attached invertebrates at the start (To) and at the end of the experiment. (CON (Control); F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Upper panel shows the subtropical Uruguayan lakes (30–35° S) and the lower panel the temperate Danish lakes (55–57° N). Standard errors are in brackets.
Table 3. Trophic group biomass (mg g−1 plant) of plant-attached invertebrates at the start (To) and at the end of the experiment. (CON (Control); F (fish); F + INV (fish + predatory macroinvertebrates) and INV (predatory macroinvertebrates). Upper panel shows the subtropical Uruguayan lakes (30–35° S) and the lower panel the temperate Danish lakes (55–57° N). Standard errors are in brackets.
DiarioBlancaNutrias
ToCONFINVF + INVToCONFINVF + INVToCONFINVF + INV
Predator0,01 (0,0)0,0 (-)0,0 (-)0,04 (0,1)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)
Gatherer14,8 (5,8)90,5 (41,4)1,3 (0,4)2,9 (0,7)3,2 (0,1)17,2 (6,3)19,3 (9,2)10,1 (2,0)21,2 (12,5)2,3 (1,0)0,0 (-)0,0 (-)0,0 (-)0,6 (0,2)0,0 (-)
Shredder1,9 (1)0,0 (-)0,0 (-)0,23 (0,3)0,0 (-)0,8 (0,5)0,0 (-)3,8 (0,6)0,0 (-)0,0 (-)0,1 (0,0)0,0 (-)0,0 (-)0,4 (0,1)0,2 (0,1)
Filterer0,09 (0,09)0,03 (0,01)0,04 (0,02)0,0 (-)0,0 (-)1,9 (1,9)0,06 (0,06)0,0 (-)0,05 (0,03)0,0 (-)0,0 (-)0,01 (0,0)0,0 (-)0,0 (-)0,0 (-)
Scrapers0,14 (0,14)0,0 (-)0,0 (-)37,8 (37,8)40,9 (40,9)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)
Piercer0,02 (0,02)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)
Total Biomass15,3 (5,8)90,5(41,4)1,3 (0,4)40,7 (37,8)44,0 (40,9)19,1 (6,2)19,4 (9,1)10,5 (2,1)21,2 (12,5)22,3 (0,9)0,01 (0,0)0,01 (0,0)0,0 (-)0,7 (0,2)0,02 (0,01)
StigsholmKogleaksBøllingso
ToCONFINVF + INVToCONFINVF + INVToCONFINVF + INV
Predator0,3 (0,2)0,6 (0,6)0,1 (0,02)0,5 (0,2)0,09 (0,09)0,01 (0,0)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,02 (0,01)0,0 (-)0,0 (-)0,0 (-)0,02 (0,02)
Gatherer1,5 (0,9)3,4 (0,3)0,02 (0,0)0,8 (0,3)0,1 (0,08)6,4 (2,3)36,7 (19,4)0,9 (1,7)0,01 (0,01)0,4 (0,2)4,5 (1,4)12,8 (8,2)4,2 (1,4)6,2 (2,5)2,9 (1,6)
Shredder3,7 (2,7)7,6 (6,0)1,4 (0,8)4,9 (3,95)0,01 (0,01)0,02 (0,01)0,05 (0,0)0,01 (0,01)0,0 (-)0,01 (0,01)1,0 (0,9)0,3 (0,3)8,4 (8,4)12,5 (12,5)0,02 (0,01)
Filterer0,0 (-)0,0 (-)0,0 (-)0,0 (-)2,2 (2,2)0,0 (-)0,0 (-)0,0(-)0,0 (-)0,0 (-)0,5 (0,2)0,0 (-)2,1 (2,1)2,2 (2,2)0,8 (0,8)
Scrapers0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)
Piercer0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)0,0 (-)
Total Biomass5,5 (3,7)11,6 (9,3)1,6 (0,8)6,2 (3,3)0,4 (0,1)6,4 (2,3)36,7 (1,9)0,9(1,7)0,01 (0,0)0,4 (0,2)6,1 (1,2)13,1 (8,2)12,8 (9,8)18,9 (1,3)3,0 (1,6)

Share and Cite

MDPI and ACS Style

Iglesias, C.; Jeppesen, E.; Mazzeo, N.; Pacheco, J.P.; Mello, F.T.-d.; Landkildehus, F.; Fosalba, C.; Clemente, J.M.; Meerhoff, M. Fish but Not Macroinvertebrates Promote Trophic Cascading Effects in High Density Submersed Plant Experimental Lake Food Webs in Two Contrasting Climate Regions. Water 2017, 9, 514. https://doi.org/10.3390/w9070514

AMA Style

Iglesias C, Jeppesen E, Mazzeo N, Pacheco JP, Mello FT-d, Landkildehus F, Fosalba C, Clemente JM, Meerhoff M. Fish but Not Macroinvertebrates Promote Trophic Cascading Effects in High Density Submersed Plant Experimental Lake Food Webs in Two Contrasting Climate Regions. Water. 2017; 9(7):514. https://doi.org/10.3390/w9070514

Chicago/Turabian Style

Iglesias, Carlos, Erik Jeppesen, Néstor Mazzeo, Juan Pablo Pacheco, Franco Teixeira-de Mello, Frank Landkildehus, Claudia Fosalba, Juan M. Clemente, and Mariana Meerhoff. 2017. "Fish but Not Macroinvertebrates Promote Trophic Cascading Effects in High Density Submersed Plant Experimental Lake Food Webs in Two Contrasting Climate Regions" Water 9, no. 7: 514. https://doi.org/10.3390/w9070514

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop