Effects of Metal Toxicants on the Energy Budgets of Marine Organisms: A Modeling Approach

Principal Investigator:
Roger M. NisbetUniversity of California, Santa Barbara


Research Summary:

Toxic compounds may reduce the fecundity, development rate, and/or survival probability of individual organisms. A large body of literature models the accumulation of toxic compounds in terms of exchange between organism and environment. These models range in sophistication from simple one-compartment models to more detailed descriptions where a growth model plays a role in determining toxicant uptake (Eby et al. 1997; Jackson 1997; Landrum et al. 1992). However, most bioaccumulation models have no feedback term through which toxicants affect organismal performance. One important mechanism for this feedback is that toxicants may cause a reduction in feeding and an increase in respiration. Such effects have been described in terms of scope for growth, defined as the excess of energy assimilation rate over respiration rate (Donkin et al. 1989; Widdows and Donkin 1991). Scope for growth is a particularly appealing concept for ecological applications, as it sidesteps the issue of chemical cause, and focuses directly on physiological effect. However, predicting the consequences of changes in scope for growth requires coupling a growth model with a dynamic model for toxicant exchange and physiological effects. We study these consequences with dynamic energy budget (DEB) models of the physiology of individuals.

DEB models use differential equations to describe the rates at which individual organisms assimilate and utilize energy from food for maintenance, growth, reproduction and development. These rates depend on the state of the organism (age, size, sex , nutritional status, etc.) and the state of its environment, (food density, temperature, toxicant levels, etc.). Solutions of the model equations represent the life history of individual organisms in a potentially variable environment.

One important use of DEB models is to relate observed patterns of growth, development, reproduction and mortality in a particular organism to empirical information on feeding rates and maintenance requirements, the goal of such studies being to get a close match between data and model descriptions for a particular species (McCauley et al. 1990). A second use, which we use in our research, takes a single, parameter- sparse, mechanistic model, and describes a broad spectrum of biological phenomena and life forms. Species differ only in their parameter values. A well-known example of this approach is the von Bertalanffy theory of growth, which with only two parameters fits the growth of many organisms very well (Kooijman 1993). The attractiveness of such parameter sparse, 'general' models is that they are relatively simple, yet they are based on physiological mechanisms. This combination warrants that these models are both testable and practicable for use at higher levels of biological organization.

We use two families of DEB models, that differ in their assumptions on priorities for energy allocation (see Gurney and Nisbet 1998, chapter 4). The first group, "net assimilation" models is represented by a thoroughly studied model developed by Kooijman (1993). The second group, "net production" models are represented by a model of Ross and Nisbet (1990) and by a more recent model of Lika and Nisbet (1998). There are a few situations where the differences between the models are important (see Nisbet et al. 1996; Gurney et al. 1996), but commonly, both models give equally good fits to limited data. Kooijman's (1993) model is currently our default choice; use of the other models allows us to explore the sensitivity of our conclusions to particular model assumptions. Kooijman (1993) discusses the physiological and physico-chemical rationale for the model assumptions and equations, which are listed in Tables 1 and 2.


Table 1.
Model assumptions.


Table 2.
Model equations for isometric organisms with dynamic food and toxicant levels.

reserve dynamics:
growth:
reproduction while growing:
reproduction without growth:
toxicant exchange:
toxic effect functions:
Cdtoxicant concentration in ambientkxarate of toxicant uptake from food
Cxtoxicant concentration in foodmmaintenance rate
Decoefficient for partitioning toxicants
between aquous fraction and reserves
m0maintenance rate in absence of toxicants
Dycoefficient for partitioning toxicants
between aquous fraction and dry structure
[Q]effective toxicant concentration
Escaled energy reserves density[Qnec]no-effect concentration
Erscaled density of energy reserved for reproductionttime
Ffood (scaled density)Vstructural biovolume
Grgrowth investmentVpstructural biovolume of maturing juvenile
KiToxicity scaling factorKcoefficient for partitioning energy between
somatic and reproductive tissues
Kadrate of toxicant uptake from ambientUenergy conductance rate
Kdarate of toxicant removalU0energy conductance rate in
absence of toxicants

Ecotoxicological applications of DEB models require additional assumptions that describe the uptake, release and metabolism of toxic compounds, and the effects of toxicants on the organism's physiology. These are discussed in detail by Muller and Nisbet (1998) who argue that the DEB model must be supplemented by both a toxicokinetic model and a toxic effect model. The toxicokinetic model describes changes in the body burden of toxicant in relation to the state variables describing size and storage, characteristics that determine the rates of toxicant exchange and the potential for toxicant accumulation (Kooijman and Vanharen 1990; Lassiter and Hallam 1990; Van Haren et al. 1994). The toxic effect model specifies how DEB model parameters change in response to body burden of toxicant.


Figure 1.
Growth of larvae of Crassostrea gigas at nominal mercury concentrations of 0 (), 5 (), 10 (), 20 (), and 40 () nM. Model fits represent fast exchange of mercury between larva and ambient (solid line) or slow exchange (broken line); see text for details. Parameter with rapid mercury exchange are: ultimate height 350 µm; initial height 63.3 (± 3.9) µm; von Bertalanffy growth rate 55.7 (± 1.7) 10-3 day-1; toxicity scaling 72.2 (± 6.5) nM; no-effect concentration in ambient 7.4 (± 1.7) nM; and development delay 1.8 (± 0.3) day (data from Beiras and His 1995).

The generality of the DEB-based approach relieves us from the impossible requirement of modeling every combination of toxicant and organism separately. While different toxicants may induce different biochemical responses in an organism, these responses translate into an effect on a small number of energy fluxes. Our formalism allows many possible assumptions on the precise form of this toxic effect; the simplest is that toxicants affect energy transduction in a hyperbolic way (resembling the way non-competitive inhibitors reduce the rate of simple Michaelis-Menten enzyme kinetics). As a result, toxicants directly affect parameters that specify the rates of feeding, assimilation and maintenance, and thereby indirectly reduce growth and reproduction rates (see Table 2 for equations).

Our default model successfully describes how various lipophilic (PAH's) and metallic (copper, cadmium, mercury) compounds, either added singly or in the form of a mixture, change the rates of feeding, respiration and growth in several animal species. Figure 1 shows an example, larval growth of the oyster Crassostrea gigas in the presence of mercury (data from Beiras and His 1995). Because the authors did not measure tissue levels of mercury (unfortunately, experimental studies tend to focus on either bioaccumulation or toxic effects, but rarely both), we fitted the data with our toxic effect model assuming the two extreme possibilities for mercury exchange: very rapid and very slow exchange. Both extremes yield acceptable fits, which shows that the toxic effect model is relatively insensitive to details of toxicant accumulation.

Figure 2. Peak values of reserves in full-grown Mytilus edulis as a function of fluctuations in food availability. The dampening effect of reserves on its dynamics strongly depends on the period at which food availability changes. When this rate is relatively high, the peak value will be low, implying that the reserve density will be cycling with a small amplitude. When the food scaled food density is fluctuating around 0.5, a typical mussel will always survive if the amplitude of the fluctuations in food is less than 0.17; beyond that, survival is a function of the periodicity.

In most natural environments organisms need to cope with temporary shortages of food. Toxicants aggravate the condition of starving organisms, and may accelerate death due to starvation. This mortality factor is a part of our model through mechanisms of 'sublethal' toxic effects, which include a reduction in feeding and an increase in maintenance requirements. In order to quantify this mortality factor, we need to model starvation strategies and specify the condition for death due to starvation in absence of toxicants.

The model we are using assumes that an organism dies when it cannot fulfill its maintenance demands, which happens when energy is released from reserves at an insufficiently rapid rate. The organism is neither supposed to regulate this mobilization rate, nor does it modify its maintenance requirements during periods of starvation. This constrains the survival potential of organisms in variable food environments, and this constraint may be unduly strong. We have conducted a few simulation studies assuming a periodically fluctuating food environment, and found that survival is a function of the period and amplitude. The results indicate that the margins for these parameters are unrealistically small.

Currently, we are studying the behavior of the DEB model with a periodically variable food environment in a systematical way. We use a few target species, among which is the marine mussel Mytilus edulis, that have been properly parameterized, and will study other species through body size scaling relationships, which are part of the model. For a singularly periodically variable food environment, we have been able to derive the environmental conditions granting long-term survival from food shortage, and we have also derived analytical solutions of the long-term dynamics of the state equations. Figure 2 exemplifies some of the model's behavior in a periodically variable food environment. It shows how ultimate peak values of energy reserves depend on the period and the amplitude of food fluctuations. It also shows the boundary conditions for long-term survival from starvation.

In the future, we plan to study starvation behavior in an environment in which food fluctuates stochastically. We expect that model organisms in a stochastically variable environment will be more vulnerable than organisms in a periodic environment. Due to its complexity, this part will be analyzed mostly with numerical studies.

Subsequently, depending on the outcomes of research, we will formulate and test alternative assumptions about energy expenditures during starvation conditions. An organism has in principle two options in order to increase its survival potential during starvation: it can regulate the mobilization of energy from reserves such that maintenance demands can be fulfilled; and it can reduce the requirements for maintenance by closing part of its physiological functions. Both options will potentially improve the realism of the model. The first option prevents a relatively fat animal from dying just because it has insufficient access to its reserves, a consequence of the model we have observed in our simulations. The second option acknowledges that many organisms have a resting stage in which part of the metabolic machinery has been shut down.

In addition to model refinements, we are investigating how toxicological parameters for different compounds can be related. Quantitative Structure Activity Relationship (QSAR) studies establish that factors influencing both toxicant accumulation and toxicant induced mortality depend on physical and chemical properties of a compound. Thus, if one compound shows a stronger effect than another, this may be because it accumulates easier, it is better retained in the animal, it is more toxic, or any combination of the above. In terms of our model, parameters for toxicant exchange and/or toxicity may be a function of physico-chemical properties of a toxicant. Currently we are exploring the uptake process. The system we are analyzing consists of mussels that take up PCBs via food and directly from the ambient. Variables are species of PCB and the fraction taken up with food. The results show a clear dependency between both variables and the octanol-water partitioning coefficient, but we have not yet modeled this dependency successfully. When we will have completed this analysis, we will study the dependency between the toxicity parameters and octanol-water partitioning coefficient, for which (again) the mussel Mytilus edulis will serve as model organism.

REFERENCES