Weithoff Funcional Classification of PhytoplanktonBy

download Weithoff Funcional Classification of PhytoplanktonBy

of 7

Transcript of Weithoff Funcional Classification of PhytoplanktonBy

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    1/7

    OPINION

    The concepts of plant functional types and functionaldiversity in lake phytoplankton a new understanding

    of phytoplankton ecology?

    GU NTR A M W EI THO F F

    Department of Ecology and Ecosystem Modeling, Institute for Biochemistry and Biology, University of Potsdam, Potsdam,

    Germany

    SU M M A R Y

    1. This is a discussion of the applicability to the phytoplankton of the concepts of plant

    functional types (PFTs) and functional diversity (FD), which originated in terrestrial

    plant ecology.

    2. Functional traits driving the performance of phytoplankton species reflect importantprocesses such as growth, sedimentation, grazing losses and nutrient acquisition.

    3. This paper presents an objective, mathematical way of assigning PFTs and measuring

    FD. Ecologists can use this new approach to investigate general hypotheses [e.g. the

    intermediate disturbance hypothesis (IDH), the insurance hypothesis and synchronicity

    phenomena] as, for example, in its original formulation the IDH makes its predictions

    based on FD rather than species diversity.

    Keywords: functional diversity, functional traits, life strategies, ordination techniques, phytoplankton

    Introduction

    The classification of species aids our understanding of

    the complexity of nature. Such classification is based

    mainly on the morphology and has only recently been

    assisted by genetics. At the same time, species have

    also been classified according to their functional/

    structural properties but, without using distinctly

    measurable characters, such classification has had

    limited application. The classification of the ecological

    strategies of plants proposed by Grime (1977) has

    been influential. He assigned species to three groups

    based on their evolutionary responses to disturbance

    and stress. This CRS concept states that in situationswith a low intensity of disturbance and a low intensity

    of stress, such as nutrient limitation, competitive

    species dominate, the C-strategists. At low distur-

    bance intensity and in a high-stress environmentstress-tolerant species (S-strategists) are expected to

    outcompete others. So-called ruderals (R-strategists)

    dominate in low-stress and high disturbance environ-

    ments. High-stress and high-disturbance environ-

    ments are too hostile for any species to persist. This

    concept has been transferred and adapted to the

    phytoplankton by Reynolds (1988), further modified

    (Reynolds, 1997) and has provided a framework to

    explain phytoplankton succession with respect to

    water column mixing (Reynolds, 1993). It has been

    applied both to natural phytoplankton communities

    (e.g. Huszar & Caraco, 1998; Melo & Huszar, 2000;Weithoff, Lorke & Walz, 2001) and to experimental

    studies (Weithoff, Walz & Gaedke, 2000).

    Besides the CRS concept, Reynolds (1980) intro-

    duced a functional classification by assigning 14

    phytoplankton associations to sets of environmental

    conditions, such as lake size, mixing regime, nutri-

    ents, light and carbon availability, etc. Over the past

    Correspondence: Guntram Weithoff, Department of Ecology and

    Ecosystem Modeling, Institute for Biochemistry and Biology,

    University of Potsdam, Maulbeerallee 2, D-14469 Potsdam,

    Germany. E-mail: [email protected]

    Freshwater Biology (2003) 48, 16691675

    2003 Blackwell Publishing Ltd 1669

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    2/7

    two decades this approach has been refined and an

    upgraded classification was presented recently by

    Reynolds et al. (2002). In its present form, 31 associa-

    tions are described but the new groupings were

    accommodated on intuitive grounds (Reynolds et al.,

    2002). However, the different algae forming a single

    group have similar morphological features which arepowerful predictors of optimum dynamic perform-

    ance (Reynolds & Irish, 1997), although algae with

    different ecological strategies might be well adapted to

    similar environmental conditions. Nevertheless, the

    overall relationships between the algal associations

    and their habitats are quite well founded.

    Recently, terrestrial plant ecologists have revived

    the idea of a functional classification in order to

    predict possible changes in the vegetation as a result

    of global climate change (Lavorel et al., 1997; Smith,

    Shugart & Woodward, 1997). The term plant func-

    tional types (PFTs) was coined (Smith et al., 1993)

    and defined as sets of species showing similar

    responses to the environment and similar effects on

    ecosystem functioning (Gitay & Noble, 1997). In

    terrestrial environments PFTs differ among habitats

    because habitats, ranging from deserts to marshes or

    forests, are inhabited by very different kinds of

    plants, whereas the pelagic phytoplankton commu-

    nities from different water bodies are relatively

    similar. Relevant functional traits of species were

    proposed to be included in the set of functional

    characteristics (Weiher et al., 1999). A functional traitin this context is a feature or property of an organism

    which is measurable and influences one or more

    essential functional processes such as growth, repro-

    duction, nutrient acquisition, etc.

    The concept of functional diversity (FD) touches

    another aspect of the functional characterisation of

    species and communities. FD reflects the functional

    multiplicity within a community rather than the

    multiplicity of species. A simple measure of FD is

    the number of co-occurring functional types (Marti-

    nez, 1996), analogous to species richness at the species

    level. A more refined quantification of FD includes a

    distance measure. In such a case the distance between

    species based on their functional traits is calculated,

    i.e. FD is high when species with widely differing

    functional traits are present in the same community.

    Although in many diversity studies a FD is implied,

    species diversity or species richness is used as a

    diversity measure, which may lead to misinterpreta-

    tions. Thus, a new and mathematically quantifiable

    measure for FD is needed. In the following section I

    describe the basis upon which functional traits should

    generally be selected and which traits are useful for

    investigating PFT and FD in phytoplankton. After that

    I discuss a number of aspects in general ecology that

    could be investigated by using the proposed approach.

    Selection of functional traits

    The selection of traits is of crucial importance because

    all subsequent classifications or calculations of eco-

    logical distance (see below) depend on them (Gitay &

    Noble, 1997). The abundance and dynamics of any

    population are driven by ambient standing stock,

    growth and loss processes. For phytoplankton, net

    growth is the sum of intrinsic growth, sedimentation,

    grazing losses and some other less important loss

    factors. Therefore, the traits selected for phytoplank-

    ton should reflect these three main processes, which

    are a surrogate for population performance. In

    general, all traits should be easily measurable (Keddy,

    1992; McIntyre et al., 1999; Walker, Kinzig & Lan-

    gridge, 1999). Thus, for this purpose the capability of a

    species to acquire a particular nutrient through

    N-fixation or phagotrophy, for example, or the

    demand for another like silica, is the appropriate

    information, especially as such information is avail-

    able even for less well-known species. Resource-

    dependent growth rates are scarce in the literatureand more useful for dynamic modelling approaches,

    e.g. in the PROTECH model (Reynolds et al., 2001).

    In the following, I propose six traits as valuable for

    characterising functional aspects of phytoplankton.

    Most of them refer to processes which are also

    reflected in the dynamic PROTECH model (Elliott

    et al., 1999).

    Size

    According to allometric theory, size is a determinant

    of specific physiological activities such as growth, and

    the size range of phytoplankton covers more than five

    orders of magnitude, ranging from autotrophic pico-

    plankton (50 000 lm3). Size in combination with shape also

    affects edibility. For Cladocera, the maximum-sized

    particle that can be ingested depends directly on the

    body size of the animal (Burns, 1968). Additionally,

    1670 G. Weithoff

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    3/7

    size and shape determine the surface to volume ratio

    that in turn influences nutrient uptake.

    Nitrogen fixation

    The potential for nitrogen fixation (e.g. in Nostocales,

    Cyanoprokaryota) gives a competitive advantageunder nitrogen-limiting conditions.

    Demand for silica

    Aside from diatoms, which need silica for their frus-

    tules, Chrysophyceae and Synurophyceae form stato-

    spores (e.g. Ochromonas), bristles and scales (e.g. Synura

    or Mallomonas) also made of silica (Lee, 1999). In addi-

    tion, silica increases the specific weight, which leads to

    higher sedimentation rates, especially in diatoms.

    Phagotrophy

    The ability to ingest bacteria serves as an additional

    source of nutrients and energy for phagotrophs.

    Particularly under nutrient-poor conditions the up-

    take of bacteria may contribute significantly to the

    phosphorus budget of the cell.

    Motility

    Mobile organisms can migrate into favourable patches

    and counteract sedimentation. This may be of partic-ular advantage in an environment exhibiting steep

    gradients, such as the chemocline in stratified lakes

    (Gervais, 1997). Being non-motile is not a disadvan-

    tage per se as, in shallow waters, sedimentation may

    allow nutrient-depleted diatoms to take up minerals

    from the sediment surface and begin photosynthesis

    again soon after resuspension (Sicko-Goad, Stoermer,

    & Fahnenstiel, 1986). This meroplanktic behaviour can

    be seen as an efficient life strategy in shallow waters

    (Carrick, Aldridge & Schelske, 1993). In addition,

    motility effects nutrient deficiency as the movement of

    cells minimises the hydrate envelope and, thus, the

    diffusive boundary layer for nutrients around the cells

    (Pasciak & Gavis, 1974).

    Shape

    The shape of a cell or colony is important with respect

    to their susceptibility to zooplankton grazing. A

    suitable measure for ingestibility for filter-feeding

    zooplankton is the longest linear dimension (LLD) of

    the food item. As mentioned above, the shape together

    with size also influences the surface to volume ratio,

    which has been used in the PROTECH model to

    predict the maximum specific replication rate (Elliott

    et al., 1999; Reynolds, Irish & Elliott, 2001). In thepresent study, shape and size are treated separately.

    All the above proposed traits are relatively easy to

    determine or data are available in the literature.

    Extensive and time-consuming preparations for the

    taxonomic determination of diatoms or dinoflagellates

    are not necessary and the proposed procedure is

    robust against new taxonomic findings that have

    accelerated recently supported by molecular methods.

    The set of traits selected may differ from one inves-

    tigator to another and some variation is reasonable

    depending on the question under consideration in a

    particular study.

    Data processing and the assignment of functional

    groups and functional diversity

    In this section I discuss different mathematical proce-

    dures for assigning plants to PFTs and recommend a

    promising procedure for calculating FD based on the

    above selected traits. The values for the selected

    functional traits can be categorised as either binary or

    other codes. For some traits, such as silica demand,

    the capability of nitrogen fixation or bacterivory,binary data (0, 1) are recommended. For others a

    categorisation into size classes (size, LLD) is sugges-

    ted. Motility can be classified as: 0, non-motile; 0.5,

    buoyancy regulation (through gas vacuoles); and 1,

    flagellated species which can move in three-dimen-

    sional space. From these data a species-trait matrix is

    created. As all traits are standardised in a range of 0

    1, all traits are regarded to be of equal importance.

    This is not a prerequisite for this approach, because an

    a priori weighting is reasonable when some traits are

    regarded as ecologically more important than others.

    Common methods for determining functional groups

    are multivariate statistics, such as clustering tech-

    niques and principal component analysis (PCA),

    correspondence analysis (CA), canonical correspon-

    dence analysis (CCA) or principal coordination ana-

    lysis (PCoA) (e.g. Jaksic & Medel, 1990; Leishman &

    Westoby, 1992; Pillar, 1999; Usseglio-Polatera et al.,

    2000; Kruk et al., 2002). CCA is a valuable option

    Functional diversity in phytoplankton 1671

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    4/7

    when additional environmental data are available. In

    this case species are ordinated and environmental

    factors are displayed as vectors within the ordination

    (e.g. for phytoplankton: Fangstrom & Willen, 1987;

    Kruk et al., 2002). Potential pitfalls in CCA with

    respect to a functional classification may occur when

    functionally different species co-occur under similarenvironmental conditions, e.g. small cryptophytes

    and large cyanobacteria colonies in eutrophic lakes

    (Sommer et al., 1986).

    For cluster analysis, the species-trait matrix is

    transformed into a distance matrix. On this basis the

    species are clustered and the outcome of the cluster

    analysis is a dendrogram, which can be used to assign

    species into groups. Cutting the dendrogram into

    sub-dendrograms (branches) leads to a number of

    functional groups (Petchey & Gaston, 2002). This

    procedure can be used arbitrarily depending on the

    resolution in question, i.e. whether many or few

    groups should be created. The sum of all branch

    lengths is taken as the FD (Petchey & Gaston, 2002). It

    may be normalised for species number (s) as

    FDnorm FD/s, which reflects the mean branch

    length and therefore allows comparisons between

    samples with different s.

    Ordination techniques such as PCA, CA or PCoA

    result in biplots based on species traits. Species with

    similar ecological traits are located closely together

    whereas widely differing species are distant from

    each other. These techniques aim for a reduction offactors and thus inherit a weighting of traits. In PCA,

    abundant binary data may lead to misinterpretations

    but categorised variables are suitable and, therefore,

    PCA is not a suitable technique for the proposed

    traits. For the generation of biplots in PCA (e.g. PC1

    versus PC2 or PC3 versus PC1) the euclidean distance

    between species is maintained. In contrast to that, the

    resulting biplots from CA display the v2 distance

    between species. Different distance measures, of

    course, influence the outcome of the functional clas-

    sification. In PCoA, the species-trait matrix is trans-

    formed into a distance matrix prior to ordination. This

    means the suitable distance measure for a particular

    investigation can be chosen. This, and robustness even

    when using binary data, makes PCoA a suitable

    technique for the approach suggested. The first three

    axes generated by all three techniques ideally explain

    a large amount of the variation in the data set. Thus a

    three-dimensional ordination can be seen as an

    ecological trait space spanned by the species accord-

    ing to their traits. As in dendrograms, functional

    groups can be created by dissecting the trait space into

    different sub-spaces containing the species forming a

    group. These groups can then be compared with the

    functional classification proposed by Reynolds et al.

    (2002). Additionally, each species has distinct coordi-nates within this ecological trait space, which enables

    us to calculate the variation around the community

    mean as a measure of FD. In other words, the centre of

    gravity of the species present forms the community

    mean and the size of the space around it is a measure

    of the FD. This centre of gravity represents the

    location of a system based on the function of its

    inhabitants and community shifts can be observed

    according to the directional movements of this centre.

    An exemplar comparison of species diversity calcula-

    ted according to the Shannon index, and FD for the

    phytoplankton of Lake Constance in southern Ger-

    many (Gaedke, 1998) is shown in Fig. 1. The species-

    trait matrix was generated as described above and

    was transformed into a distance matrix using the v2

    distance. From the distance matrix a PCoA was run

    and the species coordinates were extracted. For FD the

    sum of the biomass weighted squared distances of

    each species present from the community mean was

    calculated for each of the first three axes (weighted

    according to the eigenvalue of each axis). Despite

    some relation, this figure clearly shows that a high

    species diversity does not necessarily coincide with ahigh FD and at intermediate species diversity FD is

    0.0 0.2 0.40

    1

    2

    3

    4

    5

    Species

    diversity

    FD

    Fig. 1 Comparison of species diversity and FD from 845 phy-

    toplankton samples from Lake Constance over a 21-year period.

    1672 G. Weithoff

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    5/7

    highly variable. This proposed measure is new and

    based on distinct functional traits determining the

    functional distance between species. Therefore, it

    overcomes existing problems with traditional diver-

    sity indices (Hurlbert, 1971) and enables new insights

    into phytoplankton and general ecology.

    Applications of functional classifications and FD

    The intermediate disturbance hypothesis

    and ecosystem function

    Over the past few decades ecologists have searched

    for the main factors driving diversity. Among others,

    the frequency and intensity of disturbances (the

    intermediate disturbance hypothesis (IDH); Connell,

    1978), the productivity of a system (e.g. Rosenzweig,

    1995) and predation (including herbivory) have been

    shown to influence diversity directly. The IDH states

    that diversity peaks at an intermediate frequency or

    intensity of disturbance as a result of the co-existence

    of pioneers, stress-tolerants and ruderals. Thus, the

    IDH relies on the concept of FD, but empirical studies

    testing this hypothesis have concentrated almost

    entirely on species diversity (but see Weithoff et al.,

    2001). This same inconsistency appears when species

    richness (species number per unit area/volume) is

    considered. Calculation of FD overcomes this incon-

    sistency and ecologists are encouraged to perform

    studies on the IDH in its original sense using FD.In recent years the role of diversity has been

    investigated with respect to ecosystem function (e.g.

    Schulze & Mooney, 1993; Daz & Cabido, 2001; Kinzig,

    Pacala & Tilman, 2001), but no defined measure of FD

    was used. I suggest the use of FD, in the way proposed

    in this paper, to compare different studies and provide

    a more objective way of characterising FD. In aquatic

    sciences, FD has often been neglected and the conse-

    quences for ecosystem processes of FD in the phyto-

    plankton should be investigated in the future.

    The insurance hypothesis and functional redundancy

    Important questions in community ecology are Why

    are there so many species? (Hutchinson, 1961) and

    what role do the rare species play in a given

    community? (Gaston, 1994, 1996). In many habitats,

    a few species dominate the biomass of a community

    whilst many others share a very limited amount. The

    question arises, therefore, as to whether the rare

    species are specialised and have found their ecological

    niche or whether they are in a transient state of being

    outcompeted or having recently colonised the habitat.

    Occupying a distinct niche makes a species less

    susceptible to competitive exclusion. The insurance

    hypothesis states that rare species form a functionalbackup for dominant species, and that they can

    respond quickly to disturbance thus increasing the

    resilience of ecosystem processes. There is theoretical

    and experimental evidence from plankton and other

    communities supporting this hypothesis (McGrady-

    Steed, Harris & Morin 1997; Naeem & Li, 1997; Yachi

    & Loreau, 1999; Fonseca & Ganade, 2001). Studies on

    the insurance hypothesis should be based on quanti-

    fied relative functional redundancy among species, as

    is possible if the proposed procedure is applied.

    Ecosystem shifts

    Ecosystems continuously undergo changes caused by

    season, climate change or multiple anthropogenic

    impacts, such as eutrophication, chemical pollution or

    disturbances in the cycling of nutrients or water. Such

    changes may cause gradual or rapid shifts from one

    state to another (Scheffer et al., 2001). In each case, the

    community exhibits a directional change which can be

    traced by ordinating the community or its centre of

    gravity at any given time in an ecological space based

    on the abundance of the functional entities present. Afunctional analysis of phytoplankton long-term data

    may reveal such directional shifts, and inter-lake

    comparisons are facilitated by the functional ataxo-

    nomical approach. Such shifts in communities may

    then be related to climate change, eutrophication or

    other factors on a broader geographical scale. This

    will increase our understanding of perturbed systems

    and provide a guide to better management.

    Conclusions

    The aim here is to encourage phytoplankton ecologists

    to adopt functional concepts and to apply them in

    phytoplankton research. The selection of functional

    traits is crucial and requires broad discussion among

    researchers working in different types of lakes. The

    traits proposed in this contribution are also open to

    debate and different traits or different modalities may

    in fact prove more useful. One of the trait selection

    Functional diversity in phytoplankton 1673

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    6/7

    criteria was practicability. Many phytoplankton data

    sets already exist and the proposed traits could offer

    new insights into phytoplankton and general ecology,

    without laboriously collecting new data. Compared

    with higher plants, generation times in algae are very

    short and, even within a season, true succession takes

    place (Sommer et al., 1986). This should allow plank-ton ecologists to investigate hypotheses of general

    ecological interest and lead to a better perception of

    limnological studies within the general ecological

    community (Reynolds, 1998). The application of func-

    tional classifications and FD is by no means an

    argument for neglecting a careful species determin-

    ation and/or autecological research, although it is seen

    as a very valuable addition to traditional taxonomic

    approaches and to the list of phytoplankton associa-

    tions proposed by Reynolds et al. (2002).

    Acknowledgments

    This manuscript benefited greatly from the comments

    by F. Gervais, G. Fussmann, U. Gaedke, V. Bissinger

    and E.M. Bell. A. Hildrew helped with clarity. Th.

    Kumke gave valuable statistical advice. The data in

    Fig. 1 were acquired within the Integrated Research

    Project (SFB) 248 The cycling of matter in Lake

    Constance and successor projects.

    References

    Burns C.W. (1968) The relationship between body size of

    filter-feeding Cladocera and the maximum size of

    particles ingested. Limnology and Oceanography, 33,

    675678.

    Carrick H.J., Aldridge F.J. & Schelske C.L. (1993) Wind

    influences phytoplankton biomass and composition in

    a shallow, productive lake. Limnology and Oceanogra-

    phy, 38, 11791192.

    Connell J.H. (1978) Diversity in tropical rain forests and

    coral reefs. Science, 199, 13021310.

    Daz S. & Cabido M. (2001) Vive la difference: plant

    functional diversity matters to ecosystem processes.Trends in Ecology and Evolution, 16, 646655.

    Elliott J.A., Reynolds C.S., Irish A.E. & Tett P. (1999)

    Exploring the potential of the PROTECH model to

    investigate phytoplankton community theory. Hydro-

    biologia, 414, 3743.

    Fangstrom I. & Willen E. (1987) Clustering and canonical

    correspondence analysis of phytoplankton and envir-

    onmental variables in Swedish lakes. Vegetatio, 71, 8795.

    Fonseca C.R. & Ganade G. (2001) Species functional

    redundancy, random extinctions and the stability of

    ecosystems. Journal of Ecology, 89, 118125.

    Gaedke U. (1998) Functional and taxonomical properties

    of the phytoplankton community: interannual variab-

    ility and response to re-oligotrophication. Archiv fur

    Hydrobiologie Special Issues: Advances in Hydrobiology,

    53, 119141.

    Gaston K.J. (1994) Rarity. Chapman & Hall, London.

    Gaston K.J. (1996) Biodiversity A Biology of Numbers and

    Differences. Blackwell, Oxford.

    Gervais F. (1997) Diel vertical migration of Cryptomonas

    and Chromatium in the deep chlorophyll maximum of a

    eutrophic lake. Journal of Plankton Research, 19, 533550.

    Gitay H. & Noble I.R. (1997) What are functional types and

    how should we seek them? In: Plant Functional Types

    Their Relevance to Ecosystem Properties and Global Change

    (Eds T.M. Smith, H.H. Shugart & F.I. Woodward), pp.

    319. Cambridge University Press, Cambridge.

    Grime J.P. (1977) Evidence for the existence of threeprimary strategies in plants and its relevance to

    ecological and evolutionary theory. American Natural-

    ist, 111, 11691194.

    Hurlbert S.H. (1971) The nonconcept of species diversity:

    a critique and alternative parameters. Ecology, 52, 577

    586.

    Huszar V.L.M. & Caraco N. (1998) The relationship

    between phytoplankton composition and physical

    chemical variables: a comparison of taxonomic and

    morphologicalfunctional approaches in six temperate

    lakes. Freshwater Biology, 40, 118.

    Hutchinson G.E. (1961) The paradox of the plankton.American Naturalist, 95, 137147.

    Jaksic F.M. & Medel R.G. (1990) Objective recognition of

    guilds: testing for statistically significant species

    clusters. Oecologia, 82, 8792.

    Keddy P.A. (1992) A pragmatic approach to functional

    ecology. Functional Ecology, 6, 621626.

    Kinzig A.P., Pacala S.W. & Tilman D. (2001) The Func-

    tional Consequences of Biodiversity. Princeton University

    Press, Princeton.

    Kruk C., Mazzeo N., Lacerot G. & Reynolds C.S. (2002)

    Classification schemes for phytoplankton: a local

    validation of a functional approach to the analysis of

    species temporal replacement. Journal of PlanktonResearch, 24, 901912.

    Lavorel S., McIntyre S., Landsberg J. & Forbes T.D.A.

    (1997) Plant functional classifications: from general

    groups to specific groups based on response to

    disturbance. Trends in Ecology andEvolution, 12, 474478.

    Lee R.E. (1999) Phycology. Cambridge University Press,

    Cambridge.

    1674 G. Weithoff

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675

  • 7/29/2019 Weithoff Funcional Classification of PhytoplanktonBy

    7/7

    Leishman M.R. & Westoby M. (1992) Classifying plants

    into groups on the basis of associations of individual

    traits evidence from Australian semi-arid wood-

    lands. Journal of Ecology, 80, 417424.

    Martinez N.D. (1996) Defining and measuring functional

    aspects of biodiversity. In: Biodiversity: A Biology of

    Numbers and Differences (Ed. K.J. Gaston), pp. 114148.

    Blackwell, Oxford.

    McGrady-Steed J., Harris P.M. & Morin P.J. (1997)

    Biodiversity regulates ecosystem predictability. Nature,

    390, 162165.

    McIntyre S., Daz M., Lavorel S. & Cramer W. (1999)

    Plant functional types and disturbance dynamics

    introduction. Journal of Vegetation Sciences, 10, 604

    608.

    Melo S. & Huszar V.L.M. (2000) Phytoplankton in an

    Amazonian flood-plain lake (Lago Batata, Brasil): diel

    variation and species strategies. Journal of Plankton

    Research, 22, 6376.

    Naeem S. & Li S. (1997) Biodiversity enhances ecosystemreliability. Nature, 390, 507509.

    Pasciak W.J. & Gavis J. (1974) Transport limitation of

    nutrient uptake in phytoplankton. Limnology & Ocea-

    nography, 19, 881888.

    Petchey O.L. & Gaston K.J. (2002) Functional diversity

    (FD), species richness and community composition.

    Ecology Letters, 5, 402411.

    Pillar V.D. (1999) On the identification of optimal plant

    functional types. Journal of Vegetation Sciences, 10, 631

    640.

    Reynolds C.S. (1980) Phytoplankton assemblages and

    their periodicity in stratifying lake systems. HolarcticEcology, 3, 141159.

    Reynolds C.S. (1988) Functional morphology and the

    adaptive strategies of freshwater phytoplankton. In:

    Growth and Reproductive Strategies of Freshwater Phyto-

    plankton (Ed. C.D. Sandgren), pp. 388433. Cambridge

    University Press, Cambridge.

    Reynolds C.S. (1993) Scales of disturbance and their role

    in plankton ecology. Hydrobiologia, 249, 157171.

    Reynolds C.S. (1997) Vegetation Processes in the Pelagic. A

    Model for Ecosystem Theory. ECI, Oldendorf.

    Reynolds C.S. (1998) The state of freshwater ecology.

    Freshwater Biology, 39, 741753.

    Reynolds C.S. & Irish A.E. (1997) Modelling phytoplank-ton dynamics in lakes and reservoirs: the problem of

    in-situ growth rates. Hydrobiologia, 349, 517.

    Reynolds C.S., Irish A.E. & Elliott J.A. (2001) The

    ecological basis for simulating phytoplankton re-

    sponses to environmental change (PROTECH). Ecolo-

    gical Modelling, 140, 271291.

    Reynolds C.S., Huszar V., Kruk C., Naselli-Flores L. &

    Melo S. (2002) Towards a functional classification of

    the freshwater phytoplankton. Journal of Plankton

    Research, 24, 417428.

    Rosenzweig M.L. (1995) Species Diversity in Space and

    Time. Cambridge University Press, Cambridge.

    Scheffer M., Carpenter S.R., Foley J.A., Folke K. & Walker

    B. (2001) Catastrophic shifts in ecosystems. Nature, 413,

    591596.

    Schulze E.D. & Mooney H.A. (1993) Biodiversity and

    Ecosystem Function. Springer-Verlag, New York.

    Sicko-Goad L., Stoermer E.F. & Fahnenstiel G. (1986)

    Rejuvenation of Melosira granulata (Bacillariophyceae)

    resting cells from anoxic sediments of Douglas Lake,

    Michigan. 1. Light microscopy and 14C uptake. Journal

    of Phycology, 22, 2228.

    Smith T.M., Shugart H.H., Woodward F.I. & Burton P.J.

    (1993) Plant functional types. In: Vegetation Dynamics

    and Global Change (Eds A.M. Solomon & H.H. Shugart),

    pp. 272292. Chapman & Hall, New York.

    Smith T.M., Shugart H.H. & Woodward F.I. (1997) Plant

    Functional Types Their Relevance to Ecosystem Propertiesand Global Change. Cambridge University Press, Cam-

    bridge.

    Sommer U., Gliwicz Z.M., Lampert W. & Duncan A.

    (1986) The PEG-model of seasonal succession of

    planktonic events in fresh waters. Archiv fur Hydro-

    biologie, 106, 433471.

    Usseglio-Polatera P., Bournaud P., Richoux P. & Tachet

    H. (2000) Biological and ecological traits of benthic

    freshwater macroinvertebrates: relationships and defi-

    nition of groups with similar traits. Freshwater Biology,

    43, 175205.

    Walker B., Kinzig A. & Langridge J. (1999) Plant attributediversity, resilience, and ecosystem function: the

    nature and significance of dominant and minor

    species. Ecosystems, 2, 95113.

    Weiher E., van der Werf A., Thompson K., Roderick M.,

    Garnier E. & Eriksson O. (1999) Challenging Theo-

    phrastus:a commoncore list of plant traitsfor functional

    ecology. Journal of Vegetation Sciences, 10, 609620.

    Weithoff G., Lorke A. & Walz N. (2000) Effects of water-

    column mixing on bacteria, phytoplankton, and

    rotifers under different levels of herbivory in a shallow

    eutrophic lake. Oecologia, 125, 91100.

    Weithoff G., Walz N. & Gaedke U. (2001) The inter-

    mediate disturbance hypothesis species diversity orfunctional diversity? Journal of Plankton Research, 23,

    11471155.

    Yachi S. & Loreau M. (1999) Biodiversity and ecosystem

    productivity in a fluctuating environment: the insur-

    ance hypothesis. Proceedings of the National Academy of

    Sciences USA, 96, 14631468.

    (Manuscript accepted 23 June 2003)

    Functional diversity in phytoplankton 1675

    2003 Blackwell Publishing Ltd, Freshwater Biology, 48, 16691675