ArticlePDF Available

Ecoenzymatic Stoichiometry and Ecological Theory

Authors:

Abstract and Figures

The net primary production of the biosphere is consumed largely by microorganisms, whose metabolism creates the trophic base for detrital foodwebs, drives element cycles, and mediates atmospheric composition. Biogeochemical constraints on microbial catabolism, relative to primary production, create reserves of detrital organic carbon in soils and sediments that exceed the carbon content of the atmosphere and biomass. The production of organic matter is an intracellular process that generates thousands of compounds from a small number of precursors drawn from intermediary metabolism. Osmotrophs generate growth substrates from the products of biosynthesis and diagenesis by enzyme-catalyzed reactions that occur largely outside cells. These enzymes, which we define as ecoenzymes, enter the environment by secretion and lysis. Enzyme expression is regulated by environmental signals, but once released from the cell, ecoenzymatic activity is determined by environmental interactions, represented as a kinetic casca...
Content may be subject to copyright.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Ecoenzymatic Stoichiometry
and Ecological Theory
Robert L. Sinsabaugh1and Jennifer J. Follstad Shah2
1Biology Department, University of New Mexico, Albuquerque, New Mexico 87131;
email: rlsinsab@unm.edu
2Watershed Sciences Department, Utah State University, Logan, Utah 84322
Annu. Rev. Ecol. Evol. Syst. 2012. 43:313–43
First published online as a Review in Advance on
September 4, 2012
The Annual Review of Ecology, Evolution, and
Systematics is online at ecolsys.annualreviews.org
This article’s doi:
10.1146/annurev-ecolsys-071112-124414
Copyright c
2012 by Annual Reviews.
All rights reserved
1543-592X/12/1201-0313$20.00
Keywords
ecoenzyme, ecological stoichiometry, biogeochemistry, decomposition,
microbial growth, resource allocation model
Abstract
The net primary production of the biosphere is consumed largely by
microorganisms, whose metabolism creates the trophic base for detrital
foodwebs, drives element cycles, and mediates atmospheric composition.
Biogeochemical constraints on microbial catabolism, relative to primary
production, create reserves of detrital organic carbon in soils and sediments
that exceed the carbon content of the atmosphere and biomass. The produc-
tion of organic matter is an intracellular process that generates thousands
of compounds from a small number of precursors drawn from intermediary
metabolism. Osmotrophs generate growth substrates from the products of
biosynthesis and diagenesis by enzyme-catalyzed reactions that occur largely
outside cells. These enzymes, which we define as ecoenzymes, enter the
environment by secretion and lysis. Enzyme expression is regulated by envi-
ronmental signals, but once released from the cell, ecoenzymatic activity is
determined by environmental interactions, represented as a kinetic cascade,
that lead to multiphasic kinetics and large spatiotemporal variation. At the
ecosystem level, these interactions can be viewed as an energy landscape
that directs the availability and flow of resources. Ecoenzymatic activity
and microbial metabolism are integrated on the basis of resource demand
relative to environmental availability. Macroecological studies show that the
most widely measured ecoenzymatic activities have a similar stoichiometry
for all microbial communities. Ecoenzymatic stoichiometry connects the
elemental stoichiometry of microbial biomass and detrital organic matter
to microbial nutrient assimilation and growth. We present a model that
combines the kinetics of enzyme activity and community growth under
conditions of multiple resource limitation with elements of metabolic and
ecological stoichiometry theory. This biogeochemical equilibrium model
provides a framework for comparative studies of microbial community
metabolism, the principal driver of biogeochemical cycles.
313
Click here for quick links to
Annual Reviews content online,
including:
• Other articles in this volume
• Top cited articles
• Top downloaded articles
• Our comprehensive search
Further
ANNUAL
REVIEWS
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
1. INTRODUCTION
The net primary production (NPP) of the biosphere is consumed largely by microorganisms, whose
metabolism creates the trophic base for detrital foodwebs, drives global carbon (C) and nutrient
cycles, and mediates atmospheric composition. Biogeochemical constraints on these catabolic pro-
cesses, relative to primary production, create reserves of detrital organic C in soils and sediments
that exceed the C content of the atmosphere and biomass by a factor of two or more (Cole et al.
2007, Houghton 2007).
The production of organic molecules is an intracellular process, broadly similar across do-
mains, that generates thousands of compounds from a small number of precursors drawn from
intermediary metabolism, fueled by the consumption of C from the environment. For autotrophs,
the C source is CO2. For osmotrophic prokaryotes and fungi, the C sources are low–molecular
mass compounds, with individual taxa limited to a small number (1–20) of growth substrates.
These growth substrates are generated from the myriad products of biosynthesis and diagenesis
by enzyme-catalyzed reactions that occur largely outside cells (Burns 1978, Chr´
ost 1991, Burns &
Dick 2002, Shukla & Varma 2011, Trasar-Cepeda et al. 2011, Dick 2012). The production of these
catabolic enzymes is directed by environmental signals in relation to cellular resources, but once
released into the environment by secretion or lysis, their activity and turnover are determined by
complex physicochemical and biochemical interactions. The extracellular catabolism of organic
matter, i.e., decomposition, is considered a rate-controlling step in the global C cycle. For this
reason, considerable research has focused on environmental enzyme reactions across molecular
to biosphere scales.
Herein, we briefly review the history of environmental enzyme research with a focus on
concepts and models that link these enzyme activities to biogeochemical processes, microbial
metabolism, and ecological theory. We refer readers to previous reviews and texts for background
on the biochemistry of particular enzymes and the methodology for measuring reaction rates. We
conclude with a synthesis of empirical data and model consensus that describes the equilibrium re-
lationships among environment resource availability, microbial metabolism, and enzymatic indices
of catabolic potential, the three components of the successional loop that drives decomposition
(Sinsabaugh et al. 2002).
2. TERMINOLOGY AND ENZYMES OF INTEREST
Several terms have been used to describe the distribution or origin of enzymes found outside
of cells, including extracellular enzymes, ectoenzymes, exoenzymes, abiontic enzymes, and free
enzymes, with varying definitions. In recent papers, we use the term ecoenzyme to broadly
encompass all enzymes located outside the confines of intact cell membranes regardless of
whether such enzymes enter the environment by secretion or lysis (Sinsabaugh et al. 2009).
This definition provides the closest correspondence between environmental enzyme activity and
organic matter decomposition.
The most studied ecoenzymes catalyze the degradation of the largest environmental sources of
organic C, nitrogen (N), and phosphorus (P) (Figure 1). The largest organic C pool is structural
polysaccharides that form the cell walls and matric glycolates of plants and microorganisms, fol-
lowed by lignins and other secondary polyphenolic molecules, storage polysaccharides, and lipids.
Polysaccharide degradation is primarily hydrolytic; lipid and phenolic degradation is primarily
oxidative. The organic N pool includes polymers of amino acids and aminosaccharides, which are
sources of C and N. The organic P pool includes labile nucleic acids and phospholipids and more
recalcitrant P storage products, principally inositol phosphates.
314 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Microbial
biomass
CO2
Signal induction
Derepression
Repression
CELL
Catabolic repression
ENVIRONMENT
Induced expression
Constitutive expression
Endproduct
inhibition
Enzymes
ENZYMES
Organic
phosphorus
Nucleic acids
Phospholipids
Inositol phosphates
Polysaccharides
Cellulose
Hemicellulose
Pectin
Starch
Chitin
Peptidoglycan
Exopolysaccharides
Organic
nitrogen
Protein
Chitin
Peptidoglycan
Polyphenols
and
hydrocarbons
Lignin
Tannin
Cutin
Suberin
Fatty acids
Humic substances
EEA
processing Assimilation Production
Figure 1
Expanded successional loop that links microbial production, detrital organic matter, and ecoenzyme activity
(EEA). (Upper loop) Principal environmental sources of organic carbon, nitrogen, and phosphorus. (Lower
loop) Expression of extracellular enzymes is controlled by signal pathways linked to environmental cues.
Some enzymes are produced constitutively, usually at low levels. More generally, production is upregulated
by induction-derepression pathways and downregulated by catabolic repression pathways controlled by
environmental signals and cellular resources. Once released, the activity of extracellular enzymes is subject to
environmental controls such as end-product inhibition. Intracellular enzymes released through cell lysis are
subject to the same suite of controls. A broader description of environmental controls is presented in Figure 2.
Within these classes, the ecoenzymatic activities (EEAs) most commonly measured generate
low–molecular mass products that can be directly consumed by microorganisms. These include
α-andβ-1,4-glucosidase, which catalyze the terminal reactions in the hydrolysis of storage and
structural glucans; leucine and alanine aminopeptidase, which hydrolyze the two most abundant
amino acids from the N-terminus of polypeptides; β-1,4-N-acetylglucosaminidase, which
catalyzes the terminal reaction in the hydrolysis of chitin; and phosphatase, which hydrolyzes
phosphate from phosphoesters. Phenol oxidases and peroxidases, which use molecular oxygen
www.annualreviews.org Ecoenzymatic Stoichiometry 315
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
and peroxide, respectively, as electron acceptors, catalyze the oxidative degradation of lignin and
the formation of humus. These choices are partially methodological; it is easier to assay activities
using soluble substrates that yield soluble products than to study the degradation of insoluble
polymers or humic complexes. This bias has value in that reactions that yield assimilable products
are the ones most directly linked to microbial metabolism (Meyer-Reil 1987, Hoppe et al. 1988,
M¨
unster 1991). In the molecular sieve model (Burns 1978), such enzymes are intermediate
agents in reaction pathways that connect the activities of polymer-degrading enzymes distributed
throughout the environmental matrix and cell membrane permeases that transport substrates into
the cell. In many cases, the enzymes that catalyze the terminal reactions in polymer degradation
are localized on cell surfaces and periplasmic spaces.
3. HISTORY OF ECOENZYMATIC RESEARCH
Skujin¸ ˇ
s (1978) summarized the history of soil enzyme research, beginning with the first report of
catalase activity in soil in 1899. Through the 1970s, the major research topics were the source of
soil enzymes (plant or microbe, intracellular or extracellular), the physicochemical interactions
of enzymes with the soil matrix, correlating EEA with soil properties, and using EEA to classify
soils. Phosphatase and urease (amidohydrolase) activities were of particular interest because of
their role in generating P and N for plant growth.
Progress was limited by the lack of sensitive methods for measuring EEA and microbial dynam-
ics. But there was a consensus that phosphatase activity increased in response to N fertilization,
an indication that microbial communities allocate resources in relation to environmental nutrient
availability (Skujin¸ˇ
s 1978). In the final synthesis chapter of Soil Enzymes, Burns (1978) described
the soil system as a molecular sieve of stabilized matric enzymes linked to cell surface enzymes,
periplasmic enzymes, and permeases.
Overbeck (1991) summarized the history of aquatic enzyme research, beginning with a 1906
paper on proteolytic activity in surface water. The importance of ecoenzymes in aquatic systems
was highlighted by ZoBell (1943), who included ecoenzymes in his model of the organization
of attached microbial communities. The text Microbial Enzymes in Aquatic Environments (Chr´
ost
1991) contains several reviews and perspectives on the role of ecoenzymes in the organization
of aquatic microbial communities. In particular, Hoppe (1991) added ecoenzymes as an integral
component of the microbial loop model for the trophic organization of planktonic microbial
communities, and Wetzel (1991) described the functional organization of aquatic ecosystems as a
system of stored, immobilized enzymes.
The questions of interest to aquatic researchers differed from those of soil researchers. Many
studies correlated phosphatase, aminopeptidase and glucosidase activities to the composition
and turnover of dissolved organic and inorganic nutrient pools and rates of microbial produc-
tion (Chr´
ost 1991), establishing the fundamental relationships on which subsequent models are
based.
By the time Enzymes in the Environment was published (Burns & Dick 2002), the perspectives
of soil and aquatic researchers had largely converged through improvements in methodology
that facilitated comparisons across systems and through conceptual advances such as the biofilm
concept (Characklis & Marshall 1990) that established a common paradigm for the organization
of attached microbial communities. During the past decade, EEA studies have merged with the
biomics paradigm, which considers ecological communities as metagenomes and metaproteomes,
while simulation models with increased mechanistic resolution connect EEA to ecological pro-
cesses and theory.
316 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Km
kcat
Ea
Enzyme concentration (V
max
)
Competitive inhibition
Noncompetitive inhibition
Diusion (D)
Nonspecic sorption
Eective sorption
Temperature
Water potential (Ψ)
pH
Figure 2
Environmental control of ecoenzymatic activity: a kinetic cascade. The kinetics of enzyme reactions can be
characterized by the parameters kcat (the rate of substrate conversion), Km(the half-saturation constant), and
Ea(activation energy), which are determined by enzyme structure. A cascade of enzyme-environment
interactions affects the apparent values of these parameters. Enzyme interactions with reaction substrates and
products (competitive inhibition), other reactive species (noncompetitive inhibition), and particle surfaces
generally reduce kcat and increase Kmand Eavalues. Other rate-controlling matric effects include
sorption-desorption reactions and diffusion rates of enzymes and substrates. These effects, in turn, are
influenced by broader environmental variables such as pH, temperature, and water potential. As a result,
kinetic parameters show multiphasic distributions and high spatiotemporal variation.
4. ENZYME EXPRESSION AND ENVIRONMENTAL CONTROL
4.1. Kinetic Cascade
Approximately 1–4% of the production of heterotrophic microbial communities is used to make
enzymes for secretion into the environment; many more are released through lysis (Maire et al.
2012). Modeling studies suggest that production of extracellular enzymes has first priority on cell
metabolism above the level of maintenance respiration (Schimel & Weintraub 2003, Moorhead
& Sinsabaugh 2006, Moorhead et al. 2012, Wang et al. 2012a). Given the energy and material
cost of producing enzymes whose control is lost to the cell, expression is closely regulated at
the level of transcription. Specifics vary among enzymes and organisms, but the general model is
that transcription is ultimately linked to environmental signals (Figure 1). These signals may be
substrates consumed by the cell, indicators of toxicity, or quorum sensing molecules (DeAngelis
et al. 2008). However, once released from the cell, whether by direct secretion or cell lysis, EEA is
determined by a hierarchy of interactions that can be represented as a kinetic cascade (Figure 2).
In the kinetic cascade, ecoenzyme function, measured by the parameters kcat (the rate of
substrate conversion), Km(the half-saturation constant), and Ea(activation energy), is determined
by interactions with reaction substrates and products (competitive inhibition), other reactive
species (noncompetitive inhibition), and matric interactions that control the sorption-desorption
and diffusion of enzymes and substrates (Resat et al. 2012). These molecular processes are also
www.annualreviews.org Ecoenzymatic Stoichiometry 317
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
0
10
20
30
40
50
60
70
80
90
100
S min P min N min CH2O CHO SOC ROC
Apparent Ea (kJ mol–1)
Figure 3
Mean apparent activation energy (Ea±standard deviation) for enzyme catalyzed processes in litter and soils
based on short-term temperature manipulations. Enzyme data for sulfur (S), phosphorus (P), and nitrogen
mineralization (N min), carbohydrate hydrolysis (CH2O), and polyphenol oxidation (CHO) come from
Table 1. Values for mineralization (respiration) of soil organic carbon (SOC) and recalcitrant organic
carbon (ROC) are taken from Gillooly et al. (2001) and Ramirez et al. (2012), respectively.
influenced by broader environmental variables such as pH, temperature, and water potential.
As a result, EEA exhibits multiphasic kinetics (Overbeck 1975, 1991; McLaren 1978; Azam &
Hodson 1981; Vrba et al. 2004) and high spatiotemporal variation (Sinsabaugh et al. 2008).
4.2. Ecoenzymatic Activities and Thermodynamics
In general, environmental interactions decrease the apparent Vmax and increase the apparent Km
and Eaof EEAs. At a macroecological scale, apparent Eareceives the most attention because of its
debatable utility for predicting changes in decomposition processes, microbial respiration, and C
storage as a consequence of warming climate (Davidson & Jannsens 2006, Sierra 2012). In general,
the activation energy of a process increases with its complexity, i.e., the number of reactions
(Sierra 2012). Short-term temperature manipulations of soils show that the apparent Eaof enzymes
mediating sulfur, P, and N mineralization are lower on average (30–34 kJ mol1) than the apparent
Eaof enzymes involved in lignocellulose degradation (44–47 kJ mol1)(Figure 3). An independent
compilation by Wang et al. (2012b) yielded mean Eaestimates of 54 kJ mol1for phenol oxidase
and peroxidase, 43 kJ mol1for β-glucosidase, 34 kJ mol1for β-endoglucanase, and 32 kJ mol1
for cellobiohydrolase. The apparent Eaof C mineralization (65 kJ mol1), measured as respiration,
is greater than the apparent Eaof ecoenzymatic reactions, and values for the mineralization of
recalcitrant C range up to 85 kJ mol1(Table 1,Figure 3).
The totality of reactions that direct resource flow to microorganisms, which include sorption-
desorption and diffusion as well as catalysis, can be represented as an activation energy landscape
(Gfeller et al. 2007). As system complexity increases, mean activation energies increase (Figure 4).
In laboratory trials, the temperature sensitivity of EEA is typically measured by adding substrate at a
saturating concentration, which simplifies the energy landscape by overwhelming potential kinetic
bottlenecks associated with sorption-desorption and diffusion processes. In contrast, respiration
responses are often measured in situ without substrate amendment and the more complex landscape
contributes to greater apparent Eavalues.
In an ecosystem context, changes in the energy landscape translate to changes in the availability
of substrates for EEA and microbial metabolism. A dynamic energy landscape complicates the
318 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Table 1 Apparent activation energies (Ea) for ecoenzymes that catalyze sulfur, P, and N mineralization; carbohydrate
hydrolysis; and polyphenol oxidation
Reference System Enzyme Ea(kJ mol1)
Trasar-Cepeda et al. 2007 Soil Sulfatase 29.6
Elsgaard & Vinther 2004 Soil Sulfatase 42.2
Tabatabai & Bremner 1970 Soil Sulfatase 25.3
Oshrain & Wiebe 1976 Soil Sulfatase 29.0
Beil et al. 1995 Free Sulfatase 46.6
Ramirez-Martinez & McLaren 1966 Clay loam Phosphatase 63.2
Kaziev 1975 Soil Phosphatase 23.0
Menezes-Blackburn et al. 2011 Free Phytase 35.8
Stone et al. 2012 Soils N-acetylglucosamindase 44.1
Bremner & Mulvaney 1978 Soils Urease 52.2
Ambus 1993 Riparian soil Denitrification 64.9
Peterjohn 1991 Desert soil Denitrification 41.0
Frankenberger & Tabatabai 1991b Soils L-glutaminase 32.4
Frankenberger & Tabatabai 1991a Soils L-asparaginase 26.6
Trasar-Cepeda et al. 2007 Grassland soil Casein protease 38.0
Trasar-Cepeda et al. 2007 Grassland soil L-argininase 23.3
Trasar-Cepeda et al. 2007 Grassland soil Urease 29.5
McClaugherty & Linkins 1990 Chitinase 25.4
Trasar-Cepeda et al. 2007 Grassland soil Cellulase 48.6
Stone et al. 2012 Soil α-glucosidase 38.7
Stone et al. 2012 Soil β-glucosidase 41.5
Stone et al. 2012 Soil β-xylosidase 46.8
Stone et al. 2012 Soil Cellobiohydrolase 52.8
Trasar-Cepeda et al. 2007 Grassland soil β-glucosidase 28.6
McClaugherty & Linkins 1990 Forest soil Exocellulase 44.8
McClaugherty & Linkins 1990 Forest soil Endocellulase 50.4
Kahkonen et al. 2001 Pine soil β-glucosidase 56.3
Davidson et al. 2012 Forest soil β-glucosidase 61.8
Kocabas et al. 2008 Free Phenol oxidase 42.3
Zhang et al. 2008 Soils Laccase 44.8
Di Nardo et al. 2004 Oak litter Laccase 55.0
Di Nardo et al. 2004 Oak litter Peroxidase 60.0
McClaugherty & Linkins 1990 Forest soil Laccase 54.4
McClaugherty & Linkins 1990 Forest soil Peroxidase 39.6
Valtcheva et al. 2003 Wood pulp Laccase 22.3
Lo et al. 2001 Free Laccase 12.4
Aktas¸ et al. 2001 Free Laccase 57.0
Acevedo et al. 2010 Clay bound Manganese peroxidase 51.9
Acevedo et al. 2010 Free Manganese peroxidase 34.4
Davidson et al. 2012 Soil Phenol oxidase 32.5
Annuar et al. 2009 Bound Laccase 23.0
www.annualreviews.org Ecoenzymatic Stoichiometry 319
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
System
complexity
Free energy
Reaction coordinate
Figure 4
The effect of enzyme-environment interactions on the availability of resources can be represented as an
activation energy landscape. In this conception, resource availability is related to free energy changes
associated with catalytic, sorption-desorption and diffusion reactions involving enzymes, substrates,
products, and inhibiters. As the matric and enzymatic complexity of the system increases, the energy
landscape becomes more complex and apparent activation energies trend greater. Colors indicate increasing
elevation on the vertical axis.
extrapolation of process rates to future conditions using Arrhenius models because altered resource
flows affect microbial metabolism and enzyme expression. Consequently, the equilibrium between
environmental resources, EEAs, and microbial metabolism is not easily linked to temperature.
5. ECOENZYMATIC ACTIVITY AND ECOLOGICAL STOICHIOMETRY
The metabolic basis of EEA stoichiometry is organismal control over enzyme expression based
on environmental signals (Figure 1). The kinetic cascade (Figure 2) that controls the activity and
turnover of ecoenzymes provides feedback by generating assimilable substrates and other signals of
nutrient availability in relation to growth requirements. Studies of EEA stoichiometry, although
not described by that term, have proceeded largely independently of ecological stoichiometry
theory (Sterner & Elser 2002), which originated from comparisons of the elemental ratios of
biomass composition.
5.1. Phosphorus
The most studied case of stoichiometric control of EEA is the generally inverse relationship
between phosphatase activity and environmental P availability (Reichardt et al. 1967, Berman
1970, Jones 1972, Speir & Ross 1978, Wetzel 1981, Chr ´
ost & Overbeck 1987). Correlations
and relative response magnitudes vary widely across aquatic and terrestrial systems in relation to
measures of organic and inorganic P concentration or availability. But overall, the generalization
is well supported by both small-scale and large-scale studies (e.g., Olander & Vitousek 2000;
Sinsabaugh et al. 2008; Hill et al. 2010a,b, 2012; Marklein & Houlton 2012; Williams et al.
2012).
320 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
5.2. Nitrogen
Relating the activities of proteases, aminopeptidases, and amidohydrolases to measures of N avail-
ability is tenuous because N sources are more varied, and amino acids and amino sugars can be
important sources of C as well as N. Proteins are the largest source of organic N. In soils, Abe &
Watanabe (2004) found that peptide N accounted for 66–90% of the N content of humic acids.
In planktonic systems, amino acids support a large fraction of bacterial production, ranging up
to 100% with a median value of about 40% (Kirchman 2003). As a result, relationships between
proteolytic activities and inorganic or organic N concentrations are variable. High-resolution
planktonic studies (e.g., Hollibaugh & Azam 1983, Someville & Billen 1983, Meyer-Reil 1987,
Cunningham & Wetzel 1989, Billen 1991, Hoppe 1991, M¨
unster 1991) generally show a close
linkage between the amino acid hydrolysis, amino acid uptake, and microbial metabolism with
turnover rates for free amino acids ranging from minutes to a few hours. Amendment studies show
that aminopeptidase activity can be depressed by additions of nitrate, ammonium, and amino acids
and induced by protein addition (Chr´
ost 1991, Foreman et al. 1998).
Jones et al. (2009) found that turnover rates for amino acids across a broad range of soils were
similar to those in aquatic systems, i.e., 1–4 h, and suggested that enzymatic hydrolysis was the
rate-limiting step for microbial metabolism. A study of 84 soils by Hofmockel et al. (2010) found
that potential rates of proteolysis, measured as amino acid production, were similar to rates of
N mineralization (amidohydrolase activity) in all ecosystems except semiarid grasslands, where
proteolysis rates were six times greater than amidohydrolase. In addition, proteolytic activity was
inversely related to extractable ammonium. In a study of leaf litter decomposition, Wanek et al.
(2010) used 15N isotopic dilution of labeled amino acids to show that rates of proteolysis exceeded
N mineralization by eight fold.
After protein, the aminopolysaccharides of chitin and peptidoglycan are the most abundant
sources of N. About 16% of the dry mass of filamentous fungi is chitin (Dahiya et al. 2006).
Zhang & Amelung (1996) reported that hydrolysable muramic acid, glucosamine, mannosamine,
and galactosamine concentration in soils was 6.2–9.5% of soil C. In a study of the Delaware
estuary, Kirchman & White (1999) found that potential chitinase activity usually exceeded 14C-
chitin degradation rates, but both measures were of the same order of magnitude. From these
data and other studies, they estimated that 10% of bacterial production in marine systems is
supported by chitin. Zeglin et al. (2012) found that N-acetylglucosamine and chitin addition
to coniferous forest soils increased β-N-acetylglucosaminidase activity, N mineralization, and
microbial respiration. Olander & Vitousek (2000) found that β-N-acetylglucosaminidase activity
decreased as soil N increased along a Hawaiian chronosequence, but N addition depressed β-N-
acetylglucosaminidase activity only in the youngest N-limited soil.
Experimental N amendment studies have been conducted in nearly every ecosystem type with
no clear trend for either proteolytic or chitinolytic enzyme activities. Reported responses, if sig-
nificant, are generally small. Enowashu et al. (2009) filtered N from rainwater in a spruce forest
and measured the response of 15 N-acquiring enzyme activities. Some activities increased, e.g.,
urease (amidohydrolase), whereas, others decreased, including leucine aminopeptidase and β-
N-acetylglucosaminidase. Zeglin et al. (2007) also found complementary responses in grassland
ecosystems: where the leucine aminopeptidase to β-N-acetylglucosaminidase ratio was high, N
amendment reduced leucine aminopeptidase activity and increased β-N-acetylglucosaminidase;
where the ratio was low, N deposition depressed β-N-acetylglucosaminidase and increased leucine
aminopeptidase. These examples highlight that there is no simple relationship between bulk mea-
sures of N availability and microbial N-acquiring EEAs at the ecosystem scale. A further com-
plication is that N addition to soils depresses microbial biomass and respiration (Treseder 2008,
Ramirez et al. 2012).
www.annualreviews.org Ecoenzymatic Stoichiometry 321
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Because amidohydrolases are the proximate agents of N mineralization, analogous to phos-
phatase and P mineralization, these activities may be more directly linked to N availability and
uptake than proteases or chitinases. But these activities are not commonly measured in ecological
studies, with the exception of urease (Bremner & Mulvaney 1978) whose activity can be induced
by urea fertilization and repressed by ammonium (Mobley et al. 1995, Hasan 2000). Blank (2002)
found that the activities of four amidohydrolases were positively correlated with N mineralization
and plant N uptake in riparian habitats colonized by an invasive crucifer. In a broad survey of soils,
Hofmockel et al. (2010) found a 1:1 ratio of protease to amidohydrolase activities.
Cellular N metabolism is mediated by the activities of glutamate dehydrogenase (GDH) and
glutamine synthetase (GS) (Marzluf 1997). GDH catalyzes a low-affinity NH+
4assimilation re-
action, the interconversion of glutamate and α-ketoglutarate; GS catalyzes a high-affinity NH+
4
assimilation reaction, the interconversion of glutamate and glutamine (Hoch et al. 2006). Jorgensen
et al. (1999) found that dissolved free amino acids were the dominant N sources for estuarine and
oceanic bacterioplankton and that greater cell-specific amino acid assimilation was associated with
greater cell-specific leucine aminopeptidase activity, a higher ratio of GDH:GS activities, and a
lower cell-specific respiration rate. Hoch & Bronk (2007) found that amendments of amino acids
and ammonium to marine bacterioplankton increased growth and repressed GS and assimilatory
nitrate reductase activities. Similarly, Geisseler et al. (2009) found that the GDH:GS activity ratio
for soil microbial communities was inversely related to the C:N availability.
5.3. Carbon
P and N mineralization are predominantly hydrolytic reactions. The ecoenzymatic degradation
of carbon molecules requires a large number of both hydrolytic and oxidative reactions, whose
stoichiometry is linked to the composition and oxidation state of available organic matter.
5.3.1. Polysaccharides. Cellulose production accounts for about half of terrestrial NPP
(Ericksson et al. 1990). Hemicelluloses, polymers of xylose, mannose, galactose, arabinose, and
glucose, compose 20–30% of the mass of plant cell walls (Ericksson et al. 1990). In aquatic sys-
tems, the capsules and glycocalyces of microorganisms and biofilms are composed of complex
polysaccharides whose mass can far exceed that of the microorganisms themselves (Decho 1990,
Leppard 1995, Pereira et al. 2009, Bellinger et al. 2010). Given the range of monomers, linkages,
and crystallinity, polysaccharide degradation requires the interaction of diverse enzymes (Warren
1996). In aquatic ecosystems, β-andα-glucosidase activities are correlated with the abundance of
dissolved polysaccharides and the turnover and uptake of neutral monosaccharides (Hoppe 1983,
Hoppe et al. 1988, M¨
unster 1991, Arnosti 2011). Polysaccharides are estimated to support 20%
of bacterial production in marine systems (Kirchman 2003, Piontek et al. 2011). In terrestrial
ecosystems, the activities of β-glucosidase and other cellulolytic enzymes are correlated with mi-
crobial metabolism and rates of mass loss from plant litter (Sinsabaugh et al. 1992, 1994; Jackson
et al. 1995; Allison & Vitousek 2004; Snajdr et al. 2011).
5.3.2. Lignin and humus. Lignin, the most recalcitrant component of plant litter, accounts for
about one-quarter of terrestrial NPP. The degradation of lignin and humus is mediated by an
array of oxidases, peroxidases, dehydrogenases, and supporting enzymes that vary widely among
taxa (Rabinovich et al. 2004, Baldrian 2006, Sinsabaugh 2010, Theuerl & Buscot 2010, Bugg
et al. 2011, Strong & Claus 2011). Many of these enzymes are produced for purposes other than
C acquisition including morphogenesis, response to oxidative stress, antimicrobial defense, and
322 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
detoxification of reactive phenols. Many activities are nonspecific and may involve the generation
of secondary organic and inorganic redox mediators.
Fungi, particularly Basidiomycota, are the most efficient lignin degraders in terrestrial sys-
tems, but bacteria are the ultimate consumers of most lignin-derived C. Lignin degradation is a
rate-limiting process in litter decomposition and a lignocellulose index (ratio of lignin to lignin
plus cellulose) of 0.7 is considered the limit of decomposition, i.e., the transition between plant
litter and soil organic matter (Berg & McClaugherty 2010). In soils, the mining of humus for
polysaccharide and polypeptide C and N may lead to a positive correlation between oxidative and
hydrolytic enzyme activities, but more commonly these activities are uncorrelated (Sinsabaugh
2010, Sinsabaugh & Follstad Shah 2011, Sinsabaugh et al. 2011). Phenol oxidase and peroxidase
activities in soils and sediments generally increase with the concentration of polyphenols, but
other factors, including pH and oxygen levels and the availability of manganese and iron medi-
ators, are also important. In soils, N saturation reduces phenol oxidase activity, decomposition
rates, microbial respiration, and microbial biomass (Knorr et al. 2005, Treseder 2008, Sinsabaugh
2010, Ramirez et al. 2012). Oxidative activities are thought to play a major role in mediating C
sequestration (Freeman et al. 2001, 2004; Collins et al. 2008; Sinsabaugh & Follstad Shah 2011).
6. MICROBIAL RESOURCE ALLOCATION
6.1. Conceptual Model
Correlating EEAs with the availability of target substrates provides empirical relationships
for decomposition models based on the successional loop and corroborates the utility of EEA
measurements for inferring microbial nutrient needs in relation to supply. At the community and
ecosystem scale, the research focus shifts from the biochemistry of specific enzymes to microbial
metabolism as the driver of decomposition processes. For terrestrial systems, early research
involved correlating EEA with soil respiration. Correlating EEA to rates of litter decomposition
began in the 1980s (Sinsabaugh et al. 1981), but it was problematic to correlate snapshots of
highly dynamic activities with cumulative changes in litter mass or composition. Integrating EEA
over time provided the first statistical models for mass-loss rates as a function of potential EEAs
(Sinsabaugh & Linkins 1993). In aquatic systems, it is much easier to measure the turnover of
dissolved organic matter pools and its relationship to EEA and microbial metabolism (Chr ´
ost
1991). The 3H-thymidine and 3H-leucine assays for bacterial productivity, which came into
widespread use in the 1980s, facilitated empirical analyses of the microbial loop model (Fuhrman
& Azam 1982).
The development of resource allocation models in the 1990s provided a conceptual paradigm
for integrating EEA and microbial metabolism (Sinsabaugh & Moorhead 1994). This effort par-
alleled the development of the biofilm concept, which extended a common model of microbial
community organization across disciplines (Characklis & Marshall 1990). The view of microbial
communities as interacting consortia with 102–104populations and multiple resource require-
ments moved microbial ecology beyond the single limiting-nutrient approach of the Monod (1949)
and Droop (1977) models and moved EEA studies closer to the field of ecological stoichiometry
(Zinn et al. 2004, Cherif & Loreau 2007, Danger et al. 2008).
6.2. Empirical Studies of Resource Allocation
In his history of soil enzyme research, Skujin¸ˇ
s (1978) noted that N fertilization of soils generally
increased phosphatase activity. A meta-analysis of soil N and P fertilization studies by Marklein &
Houlton (2012) confirms this effect for a wide variety of soils. In aquatic systems, several studies
www.annualreviews.org Ecoenzymatic Stoichiometry 323
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
showed that the ratio of glucosidase:aminopeptidase activities shifts in response to nutrient
availability (e.g., Hoppe 1991, Christian & Karl 1995).
Sinsabaugh et al. (1992, 1993) showed that differences in decomposition rate of birch sticks
placed at six sites were directly related to cellulolytic enzyme activities and inversely related to the
activities of enzymes involved in P and N acquisition. These studies provided the empirical basis
for a model that linked EEA to mass-loss rates using ratios of C-, N-, and P-acquiring EEAs as
indicators of microbial resource allocation (Sinsabaugh & Moorhead 1994).
Foreman et al. (1998) evaluated the resource allocation model by adding eight C and N amend-
ments to bacterioplankton communities from a eutrophic river sampled on eight dates over an
annual cycle. The responses of five enzymatic activities varied widely over the year with a mix
of induction-depression, feedback inhibition, and resource allocation effects. On average, ammo-
nium and leucine amendments increased bacterial production and growth efficiency and reduced
enzymatic activity (feedback inhibition). Albumin had similar effects on metabolism, but increased
peptidase and phosphatase activities (resource allocation). Glucose, cellobiose, and starch increased
production, respiration, and glycosidase activities (induction-derepression). Vanillin and tannin
increased respiration, but reduced growth efficiency and enzymatic activities. Across treatments,
EEAs were closely correlated with productivity and weakly correlated with respiration. Response
ratios were greatest for α-andβ-glucosidase followed by phosphatase and leucine aminopepti-
dase. These ratios were inversely related to ambient Vmax and apparent Km. The results supported
the resource allocation model, but illustrated that multiple mechanisms underlie shifts in ratios of
EEAs.
Allison & Vitousek (2005) tested the resource allocation model in a low-nutrient tropical
soil amended with simple and complex (insoluble) C, N, and P substrates, measuring respiration
and the activities of β-glucosidase, glycine aminopeptidase, and acid phosphatase. The activi-
ties of enzymes directed at complex nutrients generally increased in response to the addition of
complementary simple nutrients. However, ecoenzymatic responses were tempered by the ac-
tivity of enzymes stabilized on soil particles, and therefore uncoupled from microbial responses.
Hernandez & Hobbie (2010) conducted a similar experiment adding nine substrates singly or
in varying combinations and quantities to a low-nutrient grassland soil. Ecoenzymatic responses
included complementation consistent with resource allocation as well as substrate induction-
derepression. Respiration rates were directly related to the sum of β-glucosidase, α-glucosidase,
phosphatase, and leucine aminopeptidase activities.
Sinsabaugh & Follstad Shah (2010) compared the apparent activation energy of bacterioplank-
ton production in two rivers sampled over an annual cycle to EEA kinetics. Bacterial consumption
of carbohydrates and proteins varied by season due to changes in substrate concentration and
the EEAs that mediated substrate turnover. Production was closely related to the ecoenzyme-
mediated generation of assimilable substrates from dissolved carbohydrate, protein, and organic
phosphate pools. The analyses demonstrated that EEA kinetics and stoichiometry can be used to
resolve nutrient and temperature constraints on microbial community metabolism.
Collectively, these studies, and others, support resource allocation and multiple resource lim-
itation models for the functional organization of microbial communities. Changes in substrate
availability caused by altered inputs or shifts in the activation energy landscape affect the stoi-
chiometry of EEA and nutrient supply, altering microbial metabolism (Allison et al. 2007).
6.3. Macroecological Patterns of Resource Allocation
Recently, EEA data sets extending to continental and global scales make it possible to compare EEA
patterns to large-scale biogeochemical trends and evaluate models that link EEA stoichiometry to
metabolic and stoichiometric theories of ecology.
324 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Sinsabaugh et al. (2008) described the distribution of β-1,4-glucosidase (BG), leucine
aminopeptidase (LAP), β-1,4-N-acetylglucosaminidase (NAG), acid (alkaline) phosphatase (AP),
phenol oxidase and peroxidase activities in soils in relation to edaphic (soil pH, soil organic mat-
ter) and climatic (mean annual temperature, mean annual precipitation) variables. All activities
correlated with soil pH. Ratios of C:N and C:P acquiring hydrolytic activities reflected latitudinal
trends in P and N availability, with a global mean BG:(NAG+LAP):AP ratio of approximately
1:1:1. Sinsabaugh et al. (2009) extended these analyses, showing that terrestrial soils and fresh-
water sediments have similar stoichiometry with slopes near 1.0 for regressions of ln(BG) versus
ln(AP) and ln(BG) versus ln(NAG+LAP). They proposed that EEAs connect the stoichiometric
and metabolic theories of ecology by reflecting the equilibrium between the elemental composi-
tion of microbial biomass and detrital organic matter and the efficiencies of nutrient assimilation
and growth. Sinsabaugh et al. (2010) evaluated one of these predictions by normalizing EEA to
microbial productivity rates and showing that the regression slopes of ln(BG) versus ln(AP) and
ln(BG) versus ln(NAG+LAP) for plankton and biofilm communities differ in proportion to the
elemental C:P and C:N ratios of biomass.
Sinsabaugh & Follstad Shah (2011) analyzed the relationships between hydrolytic and oxidative
activities in soils and presented a conceptual model that links organic matter recalcitrance and EEA
stoichiometry. As N becomes increasingly concentrated in humus, N acquisition becomes linked
to phenol oxidase activity, and N availability to microbial communities is lower than elemental C:N
ratios suggest. The growth rate hypothesis (GRH) proposes that microbial growth rates are related
to the cellular P quota (Frost et al. 2006, Allen & Gillooly 2009). As microbial growth slows with
increasing organic matter recalcitrance, P demand should decrease relative to N. Normalizing
hydrolytic activities to phenol oxidase activity reduces the slope of the ln(BG) versus ln(AP)
regression and increases the slope of ln(BG) versus ln(NAG+LAP), reflecting the shift toward N
acquisition expected as microbial growth rates decline with increasing organic matter recalcitrance.
Kelley et al. (2011) conducted a meta-analysis of 34 studies that measured soil enzyme responses
to elevated atmospheric CO2treatments. β-N-acetylglucosaminidase activity increased signifi-
cantly with CO2enrichment across studies, suggesting increased N demand. Activities directed
toward lignocellulose degradation and P mineralization increased on average, but the responses
were not statistically significant across all studies. Consequently, shifts in EEA stoichiometry
varied with ecosystem type and duration of treatment.
Williams et al. (2012) sampled 50 streams, measuring planktonic EEA and bacterial abundance
and production as well as taking several measures of land use. Abundance, productivity, and
EEA were positively related to nutrient availability and anthropogenic land use. The ratio of β-
glucosidase to alkaline phosphatase activity approached 1:1 with increasing anthropogenic land
use and total dissolved N. The ratio of leucine aminopeptidase to alkaline phosphatase activity
approached 1:1 with increased dissolved organic C and N. Ecoenzymatic C:N:P ratios moved
closer to 1:1:1 as bacterial turnover increased.
Sinsabaugh et al. (2011) analyzed EEA data collected from 2,200 stream sites nationwide by the
US Environmental Protection Agency. The data included nine hydrolytic activities; phenol oxidase
and peroxidase activities; elemental analyses of sediment C, N, and P content; and dehydrogenase
activity as a measure of microbial respiratory potential. On average, EEAs in stream sediments are
two to five times greater per gram C than those of terrestrial soils. The mean ratios of BG:AP and
BG:(NAG+LAP) were 1.64 ±0.18 [95% CI (confidence interval)] and 1.83 ±0.04, respectively,
compared to 0.62 ±0.04 and 1.43 ±0.22 for soils, reflecting differences in the mean elemental
C:P and C:N ratios of sediment and soil (57 versus 186 and 18.2 versus 14.3, respectively).
Here, we extend these stoichiometric analyses of soils and sediments by pooling data from
studies of limnetic and marine plankton (Table 2). The slopes for the standardized major axis
www.annualreviews.org Ecoenzymatic Stoichiometry 325
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Table 2 α-glucosidase (AG), β-glucosidase (BG), alkaline phosphatase (AP), leucine aminopeptidase (LAP), and
β-N-acetylglucosaminidase (NAG) activities (in nanomoles per hour per liter) of planktonic microbial communities
Reference System NAG BG AP LAP NAG BG/AP BG/LAP
Hendel & Marxsen 1997 Streams 84.11 37.9 54.1 0.108 0.076
Williams et al. 2012 Streams 37 12.9 71.0 138 0.182 0.093
Sinsabaugh et al. 1997 River 51 23.5 32.0 148 759 0.216 0.042
Findlay et al. 1998 River 18 10.0 19.4 179 179 14.1 0.108 0.108
Sinsabaugh & Foreman 2001 River 10 37.7 99.2 227 243 0.437 0.408
Boucher & Debroas 2009 Lake 25 3.64 3.71 3.74 69.6 0.992 0.053
Pamer et al. 2011 Lakes 10 61.0 852 0.072
Vrba et al. 2004 Lakes 10 3.25 8.97 12.9
M¨
unster et al. 1992 Humic lake 32 53.9 134 14.1 0.402 3.823
Mudryk & Sk´
orczewski 2004 Estuarine lake 5 96.4 166 1646 3223 56.8 0.101 0.052
Caruso et al. 2005 Littoral ponds 76.49 177 209 0.037 0.031
Taylor et al. 2003 Estuary 47 10.4 26.1 8.62 0.398
Celussi et al. 2009a Ross Sea 14 1.67 2.20 1.54 0.759
Celussi et al. 2009b Ross Sea 23 0.61 1.54 22.3 0.60 0.396 0.027
Monticelli et al. 2003 Ross Sea 80.26 0.64 79.1 0.406 0.003
Hoppe 1983 Baltic Sea 54.38 3.63 15.3 45.7 4.32 0.237 0.079
Caruso 2010 Mediterranean Sea 10 0.70 34.8 16.0 0.020 0.044
Karner & Rassoulzadegan 1995 Mediterranean Sea 50.31 0.84 24.8 0.034
Grossart & Simon 2002 Red Sea 221.0 26.5 0.792
Baltar et al. 2009 Atlantic Ocean 20.04 0.04 0.75 5.5 0.053 0.007
Christian & Karl 1995 Pacific Ocean 60.06 8.0 0.008
Hoppe & Ullrich 1999 Indian Ocean 31 1.46 2.91 9.93 0.502 0.147
Mean 366 9.90 7.60 31.3 71.0 5.47 0.243 0.107
regressions of ln(BG) versus ln(AP) and ln(BG) versus ln(NAG+LAP) are similar to those for
soils and sediments, showing that the ecoenzymatic scaling of planktonic communities does
not differ substantially from that of attached microbial communities (Figure 5). However, the
normalization constants and mean EEA ratios for plankton are much lower than those of soils
and sediments [BG:AP and BG:(NAG+LAP) ratios are 0.26 and 0.12, respectively], reflecting
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−→
Figure 5
Global regressions for β-glucosidase (BG), acid (alkaline) phosphatase (AP), leucine aminopeptidase (LAP),
and β-N-acetylglucosaminidase (NAG). For sediments and soils, activity units are in nanomoles per hour
per gram of organic matter; for plankton, activity units are in nanomoles per hour per liter. (a) Standardized
major axis (SMA) regressions for ln(BG) versus ln(AP). Freshwater sediments (blue): b=0.946 ±0.030
(95% CI), a=1.02 ±0.29, R2=0.43, n=2,208, geomean BG:AP ratio is 1.636 ±0.179. Soils (orange):
b=1.162 ±0.056, a=–1.715 ±0.481, R2=0.40, n=929, geomean BG:AP ratio is 0.617 ±0.045.
Plankton ( gray): b=0.855 ±0.506, a=–0.836 ±0.219, R2=0.72, n=292, geomean BG:AP ratio is
0.243 ±0.086. The slope of the soil regression is significantly greater than that of the sediment regression
(p<0.05). (b) SMA regressions for ln BG versus ln (NAG+LAP). Freshwater sediments (blue): b=1.098 ±
0.028, a=–0.42 ±0.28, R2=0.62, n=2,208, geomean BG:(NAG+LAP) ratio is 1.832 ±0.036. Soils
(orange): b=1.091 ±0.063, a=–0.59 ±0.503, R2=0.16, n=929, geomean BG:(NAG+LAP) ratio is
1.434 ±0.220. Plankton ( gray): b=1.276 ±0.176, a=−3.300 ±0.808, R2=0.50, n=106, geomean
BG:(NAG+LAP) ratio is 0.123 ±0.139. The slopes of the sediment, soil and plankton regressions are not
significantly different. The BG versus LAP regression for plankton: b=1.086 ±0.104, a=–2.509 ±
0.521, R2=0.23, n=292, geomean BG/LAP ratio is 0.091 ±0.068. Trend lines highlight the range of
values and do not correspond with the SMA regressions. Freshwater sediment data from Sinsabaugh et al.
(2011). Soil data from Sinsabaugh et al. (2009). Plankton data from Table 2
326 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
P limitation on growth and the importance of dissolved proteins and aminopolysaccharides as
sources of both C and N.
Some of the residual variance associated with the global regressions for soils, sediments, and
plankton (Figure 5) can be attributed to differences in nutrient availability, pH, and other param-
eters that differ at the ecoregional scale. For example, within the soil data, which have the weakest
regressions, the alkaline soils of arid lands have high aminopeptidase and phenol oxidase activities
(relative to β-glucosidase) compared to acidic soils; and highly weathered tropical soils have high
phosphatase activities (relative to β-glucosidase) compared to higher latitude soils (Sinsabaugh
et al. 2008). Further data collection may lead to specific stoichiometries for individual soil orders.
–4
–2
0
2
4
6
8
10
12
14
2 0 2 4 6 8 10121416
ln (β-glucosidase activity)
ln (phosphatase activity)
–4
–2
0
2
4
6
8
10
12
14
2 4 6 8 10 12 14 16
ln (β-glucosidase)
ln (β-N-acetylglucosamindase
+ leucine aminopeptidase)
a
b
Freshwater sediments
Freshwater sediments
Soils
Soils
Plankton
Plankton
www.annualreviews.org Ecoenzymatic Stoichiometry 327
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
0
0.10
0.20
0.30
0.40
0.50
0.60
0.70
NAG bXylo aGluc CBH bGal bGlcur aGal aArab
Activity: β-glucosidase ratio
River plankton
Arid soil
All soils
Freshwater sediments
Freshwater biolm
Lake plankton
0
0.20
0.40
0.60
0.80
1.00
1.20
1.40
1.60
1.80
Ala Gly Ser Arg Tyr Pro Asp Glu Asn
Activity: Leu aminopeptidase ratio
a
b
Figure 6
Mean glycosidase and aminopeptidase activities relative to β-glucosidase from representative studies. River
bacterioplankton data from Sinsabaugh & Foreman (2001; n=10); arid soil data from Stursova et al. (2006;
n=24–48); all soils data from Sinsabaugh et al. (2008; n=948–1012); freshwater sediments from Hill
et al. (2012; n=2,200); freshwater biofilm data from L.L.P. Lehto and B.H. Hill (unpublished data; n=
72); lake plankton data from Vrba et al. (2004; n=360). Abbreviations: NAG, β-N-acetylglucosaminidase;
bXylo, β-xylosidase; aGluc, α-glucosidase; CBH, cellobiohydrolase; bGal, β-galactosidase; bGlcur,
β-glucuronidase; aGal, α-galactosidase; aArab, α-arabinosidase; Ala, alanine; Gly, glycine; Ser, serine;
Arg, arginine; Tyr, tyrosine; Pro, proline; Asp, aspartate; Glu, glutamate; Asn, asparagine aminopeptidase.
From the freshwater sediment data, Hill et al. (2012) were able to resolve EEA stoichiometries
for nine ecoregions.
β-glucosidase and leucine aminopeptidase are the most widely measured glycosidase and
aminopeptidase activities because they generally yield the greatest potential rates and hydrolyze
the most abundant substrates. Some studies include other measures. The expectation is that
activities will scale in proportion to the relative abundance of substrates; that appears to be the
case (Figure 6), but there are few studies that combine high-resolution analyses of substrate
composition with corresponding EEAs. We note that leucine and alanine, the two most abundant
328 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
protein amino acids, are hydrolyzed by the same enzymes, but hydrolysis of coumarin-linked
alanine is generally greater than coumarin-linked leucine even though the latter is more
commonly used (and costs more).
7. MODELS
7.1. Ecological Theory
The metabolic and stoichiometric theories of ecological function are founded on the kinetics
and structure of cellular components that mediate production and respiration (Sterner & Elser
2002, Brown et al. 2004, Gillooly et al. 2005). Each theory is based on invariance rules and
allometric relationships (Allen & Gillooly 2009, Doi et al. 2010). The metabolic theory extrapolates
the functions of terminal metabolic units involved in C fixation, respiration, and translation to
organismal and ecological levels of organization based on thermodynamics and the limitations of
material distribution, which differ for prokaryotes, protists, and metazoans (DeLong et al. 2010).
The stoichiometric theory extends the elemental composition of cellular components to ecological
processes and organization using cellular growth models and ratios of nutrient availability. These
theories are interrelated through the threshold element ratio (TER) and the GRH.
The TER is the elemental C:N or C:P ratio of available growth substrates at which control
of cellular or community metabolism switches from energy supply (C) to nutrient supply (N, P)
(Sterner & Elser 2002, Elser et al. 2003, Frost et al. 2006, Doi et al. 2010). Frost et al. (2006)
defined the TER as
TERC:P =(AP/GE) ·BC:P and TERC:N =(AN/GE) ·BC:N,1.
where APand ANare assimilation efficiencies for P and N, GE is microbial growth efficiency
with respect to C, and BC:P and BC:N are the elemental C:P and C:N ratios of microbial biomass.
Doi et al. (2010) presented a TER definition that replaces the AP/GE and AN/GE terms with
GEmax
P/GEmax
Cand GEmax
N/GEmax
C, which are maximum growth efficiencies with respect to C, N,
and P. This definition emphasizes that the TER is the condition for optimal growth. For microbial
communities and invertebrates, the ratio of GEmax
C/GEmax
Nis approximately 0.77; the GEmax
C/GEmax
P
is more variable with a mean of about 0.3 (Herron et al. 2009, Jones et al. 2009, Doi et al. 2010,
Zeglin et al. 2012). Moorhead et al. (2012) developed a simulation model for microbial community
growth on C and C+N substrates in which TER is defined as BC/N/CUE (carbon use efficiency).
Because TER links biomass composition and growth efficiency, values vary by the extent to which
elemental homeostasis is maintained in relation to external resource supply (Gusewell & Gessner
2009, Hladyz et al. 2009, Franklin et al. 2011, Hall et al. 2011).
Elemental homeostasis is itself dependent on growth rate. The GRH predicts that the cellular P
quota increases with growth rate because growth rate is directly related to ribosome density (Elser
et al. 2003, Doi et al. 2010), and ribosomes are the largest P stocks in many cells. Franklin et al.
(2011) show that the GRH does not apply under conditions of N limitation, because cells have
little capacity to alter their C:N stoichiometry by altering the relative abundance of biomolecules.
Sinsabaugh & Follstad Shah (2011) showed that increased N limitation associated with the hu-
mification of organic matter reverses the GRH by slowing growth and resource allocation to
P-acquiring enzymes.
7.2. Substrate Availability, Ecoenzymatic Activity, and Growth
The generation of an assimilable substrate from a single extracellular enzymatic reaction can be
represented by the Michaelis-Menten function (Michaelis & Menten 1913):
V=Vmax ·S/[Km+S],2.
www.annualreviews.org Ecoenzymatic Stoichiometry 329
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
where V is the reaction rate; Vmax is the maximum reaction rate when available enzyme capacity
is saturated; S is environmental substrate concentration; and Kmis the half-saturation constant
for the enzyme, i.e., S at Vmax/2. The Monod model uses the same formulation to represent the
growth of single cell organisms as a function of the environmental concentration of a limiting
substrate (Monod 1949):
μ=μmax ·S/[Ks+S],3.
where μis growth rate, μmax is the maximum growth rate, and KSis the half-saturation constant.
The presumption is that μis controlled by the substrate affinity and active transport capacity of
integral membrane proteins (Button 1993). In simulation models, growth is more complex because
the energy and materials represented by substrate consumption must be allocated to a variety of
maintenance processes, only resource consumption beyond these requirements contributes to
growth (Moorhead et al. 2012, Wang et al. 2012a).
The Monod model is not broadly applicable to microbial communities because single sub-
strate limitation is not the typical condition (Chen & Christensen 1985, Zinn et al. 2004, Cherif
& Loreau 2007, Danger et al. 2008). Within microbial communities the thresholds for sub-
strate limitation of growth vary across populations. At any given state, there are many resources
that are limiting the growth of one or more of the 102–104extant populations. In addition,
(a) populations have some physiological capacity to adapt to changing nutrient availabilities
and (b) the acquisition of various substrates may not be independent. The latter occurs because
(a) the organic matter available for degradation by heterotrophic microorganisms has a limited
range of biochemical and stoichiometric composition and (b) the acquisition of multiple resources
is integrated through optimal resource allocation strategies selected to maximize growth. If envi-
ronmental conditions remain stable for an extended period, community adaptation also includes
succession of populations. As a result, the relationship between microbial community produc-
tion rate and resource availability is less subject to discrete thresholds and sequential limitations
than single populations. It is possible to posit growth models for specific types of resource in-
teractions (Danger et al. 2008). But for microbial communities as a whole, we propose that the
relationship among multiple resources and community production may be generally described
as
μ=μmax ·{(S1·S2·...Sn)/[(KS1 +S1)·(KS2 +S2)·(...KSn +Sn)]}1/n.4.
This community growth model relates production to the geometric mean of multiple resource
availabilities (Figure 7). In this formulation, the acquisition of multiple resources is neither wholly
independent nor fully integrated. In microbial communities, the assimilation of a limiting nutri-
ent into the community is mediated by populations that have the greatest affinity for the sub-
strate (lowest KS). But once acquired by the community, nutrients can be internally recycled.
Within biofilms, nutrient concentrations can be one or more orders of magnitude greater than
the environmental concentration (Hall-Stoodley et al. 2004, Van Horn et al. 2011). As a re-
sult, community growth rates should exceed those predicted by the Monod model for a single
population.
In the metabolic theory of ecology and ecological stoichiometry theory, microbial growth
efficiency, rather than growth rate, is linked to the availability of limiting nutrients, commonly N
and P relative to C, at both the organismal and community levels (Equation 1). The Michaelis-
Menten and Monod models, and the metabolic theory of ecology and ecological stoichiometry
theory models derived from them, can be related to the stoichiometry of the EEAs that mediate
microbial nutrient acquisition from environmental organic matter (Sinsabaugh et al. 2009, 2010,
330 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Substrate 2
Substrate 1
0.8–1.0
0.6–0.8
0.4–0.6
0.2–0.4
0–0.2
µ/µmax
Figure 7
Microbial community growth rate as a fraction of maximal growth rate (μ/μmax; see Equation 4 in text)
based on availability of two substrates.
2011; Sinsabaugh & Follstad Shah 2011):
EEAC/NBC/N/LC/NTERC/N/BC/N=AN/GEC,5.
EEAC/PBC/P/LC/PTERC/P/BC/P=AP/GEC,6.
where EEAC/Nand EEAC/Pare ratios of ecoenzymatic C acquisition activity to ecoenzymatic N
and P acquisition activity, and LC/Nand LC/Pare the C:N and C:P ratios of labile organic matter
in the environment. Using the most widely measured activities, we define
EEAC/N=BG/(LAP +NAG),7.
EEAC/P=BG/AP,8.
where BG =β-1,4-glucosidase activity; LAP =leucine aminopeptidase activity; NAG =β-
1,4-N-acetylglucosaminidase activity; and AP =acid (alkaline) phosphatase activity.
Following Equations 5 and 6, the generation of assimilable N and P substrates from environ-
mental sources via EEA, expressed as a fraction relative to C, can be represented as
SC/N=BC/N/LC/N·1/EEAC/N,9.
SC/P=BC/P/LC/P·1/EEAC/P.10.
Sc/nand SC/Pare scalars for the relative availability of hydrolyzed N and P in relation to micro-
bial community composition. Microbial growth efficiency as a function of the stoichiometry of
nutrients generated by EEA can be represented as
GEC=GEmax
C·SC/N/[KC/N+SC/N]=AN·BC/N/TERC/N,11.
GEC=GEmax
C·SC/P/[KC/P+SC/P]=AP·BC/P/TERC/P,12.
where KC/Nand KC/Pare half-saturation constants for GECbased on the stoichiometry of C:N
and C:P availabilities. GECapproaches GEmax
C, approximately 0.60, when SC/NKC/N,S
C/P
KC/P, and the ratios BC/N/TERC/Nand BC/P/TERC/Pare near 0.6 (assuming ANand AP=1).
www.annualreviews.org Ecoenzymatic Stoichiometry 331
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Equations 11 and 12 propose that community growth efficiency is a saturating function of N and P
availability relative to C, linking substrate generation via EEA to microbial substrate assimilation
and biomass composition.
If community metabolism is controlled by the resource in least supply, following the Liebig
assumption, Equations 11 and 12 will generate independent estimates of GEC. In that case, com-
munity growth efficiency is equal to the lower estimate. At the other extreme, the equations
should yield identical estimates of GECif the generation and assimilation of C, N, and P are
perfectly integrated. For reasons presented above, neither condition is likely to apply. Following
the community growth model (Equation 4), we merge Equations 11 and 12 as
GEC=GEmax
C·{(SC/N·SC/P)/[(KC/N+SC/N)·(KC/P+SC/P)]}0.5
={[AN·BC/N/TERC/N]·[AP·BC/P/TERC/P]}0.5.13.
This biogeochemical equilibrium model expresses growth efficiency at the community level as the
geometric mean of the N and P supply, relative to C. The models for biogeochemical equilibrium
(Equation 13) and community growth (Equation 4) can be merged using GEC=μ/(μ+R),
where R is the community respiration rate. Equation 13 can be simplified to
GEC=GEmax
C·(Nsat ·Psat)0.5=(Ncon ·Pcon )0.5.14.
by defining Nsat and Psat as stoichiometric indices of environmental N and P saturation relative to
C,
Nsat =SC/N/(KC/N+SC/N)andP
sat =SC/P/(KC/P+SC/P),15.
and Ncon and Pcon as indices of microbial N and P consumption relative to C,
Ncon =AN·BC/N/TERC/Nand Pcon =AP·BC/P/TERC/P.16.
7.3. Model Prediction
For most ecosystems, the mean GECfor heterotrophic microbial communities is near GECmax/2
or 0.30 (Manzoni et al. 2012). In aquatic ecosystems, bacterial respiration (BR) generally increases
sublinearly with production (BR =3.42BP0.61; del Giorgio & Cole 1998). Consequently, GEC
also generally increases with production, approaching an asymptote of approximately 0.50, with
a global average of 0.26 (del Giorgio & Cole 1998). For decomposing plant litter in terrestrial
ecosystems, microbial C use efficiency, an indirect measure of growth efficiency, averages about
0.30 based on C:N stoichiometry (Manzoni & Porporato 2009, Manzoni et al. 2010). There are
fewer data for terrestrial soils, and most measurements involve calculating growth yield using
labile substrate (usually glucose) amendments. In some cases, N and P additions are also included.
Most of these estimates are in the range of 0.50–0.60 (Frey et al. 2001, Manzoni et al. 2012).
Herron et al. (2009) reported a microbial growth efficiency of 0.46 for semiarid soils using 13C-
labeled acetic acid vapor. These growth efficiency estimates for labile substrates approach maximal
values observed in culture. Community growth efficiency under ambient substrate and nutrient
conditions is probably lower and comparable to values for other ecosystems.
In theory, at GEC=0.5·GECmax balanced flows of C, N, and P would set
SC/N=KC/N=SC/P=KC/P=0.5 (Equations 11 and 12). The data set for terrestrial soils
(Figure 5) lacks estimates of LC/N,L
C/P,B
C/N,andB
C/Pfor individual samples. Using mean
values for LC/N(14.3), LC/P(186), BC/N(8.6), and BC/P(60) presented by Cleveland & Liptzin
(2007), the estimated mean values of SC/Nand SC/Pare 0.42 and 0.52, respectively. Using these
values and assuming KC/N=KC/P=0.5, the estimated mean growth efficiency for soil micro-
bial communities is 0.29 (Equation 13; Figure 8). The stream sediment data (Figure 5) include
332 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
0
0.10
0.20
0.30
0.40
0.50
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Growth eciency
SC/N × SC/P
Soils
Plankton
Sediments
0
0.10
0.20
0.30
0.40
0.50
0.60
0 1.0 2.0 3.0 4.0
Growth eciency
SC/N × SC/P
Figure 8
Biogeochemical equilibrium model (Equation 13) that predicts microbial community growth efficiency from
the elemental C:N and C:P ratios of biomass (BC/N,B
C/P) and environmental organic matter (LC/N,L
C/P),
and the ratios of ecoenzymatic activities (EEAs) that mediate C, N, and P acquisition (EEAC/N, EEAC/P).
SC/N=BC/N/LC/N·1/ EEAC/Nand SC/P=BC/P/LC/P·1/ EEAC/P. Circles are mean values for the soil,
sediment, and plankton data sets shown in Figure 5.(Inset) Predicted growth efficiencies for freshwater
sediments; this data set includes LC/Nand LC/Pmeasures for each sample (n=2,096; Hill et al. 2012).
LC/Nand LC/Pestimates for each sample, but not values for BC/Nand BC/P. Using mean BC/N
and BC/Pvalues of 8.6 and 60, respectively, for all samples, the mean values of SC/Nand SC/Pfor
stream sediments are 0.26 (95% CI =0.02, n=2192) and 0.64 (95% CI =0.15, n=2192),
respectively. The differences in SC/Nand SC/P, relative to soils is driven by the low C:P ratio of
sediments (57), which approximates BC/P.UsingK
C/N=KC/P=0.5, the mean GECfor stream
sediments is 0.27 (SD =0.10, 95% CI =0.004) (Figure 8), which is similar to the mean GEC
of 0.26 estimated by del Giorgio & Cole (1998).
Compared to soils and sediments, the mean BG/(LAP+NAG) ratio for plankton (0.123) is low
(Table 2;Figure 5) and yields a SC/Nvalue of 3.1 using a mean BC/Nof 6.6 (Sinsabaugh & Follstad
Shah 2010) and a mean LC/Nof 17.3 (Seitzinger et al. 2005), suggesting that BG/(LAP+NAG) is
not a good indicator of EEAC/Nfor bacterioplankton. Analyses of dissolved organic matter compo-
sition and bacterial metabolism show that carbohydrates and proteins make roughly equal contri-
butions to bacterioplankton production on a global basis (e.g., Kirchman & White 1999, Kirchman
2003, Rosenstock & Simon 2003, Simon & Rosenstock 2007). If proteins and aminopolysaccha-
rides supply about half the C for production (Kirchman & White 1999, Kirchman 2003), a better
indicator of EEAC/Nfor bacterioplankton is (BG+LAP+NAG)/(LAP+NAG), which has a mean
value of 1.123 and yields a SC/Nvalue of 0.34.
Using a mean BC/Pof 106 (Sinsabaugh & Follstad Shah 2010), a mean LC/Pof 656 (Seitzinger
et al. 2005), and a mean EEAC/Pratio of 0.257 (Table 2;Figure 5), the SC/Pestimate for plankton
is 0.63. This estimate is similar to that for stream sediments, but in this case SC/Pis driven by
high phosphatase activity in relation to environmental P availability, whereas sediments have high
environmental P availability and low phosphatase activities. For plankton, phosphatase expression
is presumably facilitated by the tight coupling of C and N acquisition that comes from reliance
on protein and chitin for growth. Using KC/N=KC/P=0.5 and values of 0.34 and 0.63 for
SC/Nand SC/P, respectively, the mean GECfor plankton is 0.28 (Figure 8), which approximates
the mean GECof 0.26 estimated by del Giorgio & Cole (1998). Thus, Equation 13 produces
www.annualreviews.org Ecoenzymatic Stoichiometry 333
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
similar estimates for the mean growth efficiency of microbial communities in soils, sediments,
and plankton (0.29, 0.27, 0.28, respectively) despite large differences in resource availability and
biomass composition.
7.4. Model Synthesis
The biogeochemical equilibrium model (Equation 13) yields microbial community GEC
values that converge on GECmax/2 because resource allocation to substrate acquisition, measured
as potential EEA, generally scales with environmental substrate concentration such that enzymes
operate at about one-half their maximum catalytic capacity, i.e., the Michaelis-Menten parame-
ters Vmax, S, and apparent Km(Equation 2) are positively correlated (Williams 1973, Sinsabaugh
& Follstad Shah 2010). The Monod parameters KSand μmax (Equation 3) are also positively
correlated (Lobry et al. 1992). This condition sustains high substrate turnover while optimizing
responsiveness to fluctuations in substrate availability. Expending resources to produce additional
enzyme has marginal utility because additional enzyme depresses soluble substrate concentra-
tion, reducing the reaction velocity per enzyme. In addition, excess enzyme production depresses
growth efficiency by increasing maintenance costs (Wang et al. 2012a).
An analogous condition applies to insoluble substrates because the optimal substrate:enzyme
ratio is determined by the density of effective enzyme binding sites (B¨
unemann 2008, Lynd et al.
2002). This condition is modeled using adsorption isotherms or “reverse” Michaelis-Menten
kinetics (Schimel & Weintraub 2003). In this case, expressing more enzyme has marginal value
because substrate accessibility limits their effective use. This constraint can lead to C and nutrient
limitation even in environments with high organic matter concentrations.
The ecoenzymatic controls on microbial growth become less important as environmental nu-
trient availability increases to the point of saturating microbial demand (SC/Nand SC/P>1). In
Equations 5 and 6 the terms EEAC/Nand BC/N/LC/Nare linearly related to AN·BC/N/TER
C/N
and the terms EEAC/Pand BC/P/LC/Pare linearly related to AP·BC/P/TERC/P. In Equations 11
and 12 these linearities are limited to values of SC/Nand SC/P<1, corresponding to GEC<0.30
(Figure 8). As SC/Nand SC/Pincrease beyond the half-saturation constants KC/Nand KC/P, control
of growth efficiency progressively shifts from environmental resource availability to constraints
imposed by thermodynamics and cellular organization as predicted by metabolic theory.
Equation 13 defines a biogeochemical equilibrium or attractor for environmental resource
availability, EEA stoichiometry, biomass stoichiometry, and microbial community metabolism
(Figure 8). At low resource availability, microbial growth is controlled by substrate generation
via EEA. At high resource availability, microbial growth is controlled by membrane transport
capacity and metabolic constraints. Throughout this range, growth efficiency is modulated by the
stoichiometry of C, N, and P availability in relation to microbial growth requirements. This equi-
librium attractor provides a framework for comparative studies of organic matter decomposition
and microbial community metabolism.
8. FUTURE DIRECTIONS
Several simulation models that include ecoenzyme pools have been recently published (e.g.,
Davidson et al. 2012, Moorhead et al. 2012, Resat et al. 2012, Wang et al. 2012a). These models
highlight the uncertainties in our understanding of biogeochemical processes at various scales of
resolution. They also show that resource flows and microbial community metabolism depend on
(a) rates and priorities of ecoenzyme expression and (b) the effective activity and turnover rate of
ecoenzymes in the environment. Better estimates of these parameters, and their range of variation,
334 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
are needed to translate these models into applications that predict process rates, e.g., respiration
or production, from EEA monitoring data. The existence of a common EEA stoichiometry in
dynamic equilibrium with resource availability, biomass composition and community metabolism
suggests that such applications are feasible. One advantage of process models based on EEA stoi-
chiometry is the potential to include multiple resource drivers. Microbial metabolism in soils, for
example, is generally considered to be C limited. Some models include C and N, but EEA stoi-
chiometry suggests that P acquisition, rather than N, is often a constraint on microbial metabolism,
particularly for highly weathered soils.
Models that entrain EEA stoichiometry also facilitate monitoring of ecosystem responses to
environmental changes because EEA is easily measured at high throughput and relatively low cost
compared to most of the resource and metabolic variables to which it is dynamically linked. How-
ever, almost all empirical research is based on assays of potential activities most often conducted
under substrate saturating conditions at 20 C with pH buffering. Unfortunately, there are no
commercially available systems for continuous monitoring of in situ EEA at present, due largely
to lack of a market rather than technological impediments. Such systems could resolve a variety
of mechanistic questions and provide data to parameterize process models. Questions of interest
include the decomposition/nutrient mining activities of mycorrhizal fungi, signal transduction
between plants and rhizosphere microbial communities, pulse-response dynamics to precipitation
in arid land ecosystems or flooding in riparian ecosystems, and controls on soil organic C stores
in relation to fluctuating nutrient availability and temperature regimes.
In the biomics paradigm, ecoenzymes link microbial community structure to biogeochemi-
cal process. As the technology for extracting and sequencing these enzymes advances, it will be
increasingly possible to directly match enzymes to their producers. Eventually, comparisons of
metatransciptomes to metaproteomes will also show the relative turnover time of ecoenzymes
produced by various populations within the community, resolving legacy effects. These analyses
are the final link in the genes to ecosystems integration of biological organization.
DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
This work was supported by the NSF EaGER (DEB-0946288) and Ecosystem Studies programs
(DEB-0918718) and the Sevilleta LTER. We thank Steve Allison, Christine Hawkes, Brian Hill,
Kirsten Hofmockel, Rob Jackson, Daryl Moorhead, Eldor Paul, Kathleen Treseder, Bonnie War-
ing, and Michael Weintraub for providing comments on various drafts.
LITERATURE CITED
Abe T, Watanabe A. 2004. X-ray photoelectron spectroscopy of nitrogen functional groups in soil humic
acids. Soil Sci. 169:35–43
Acevedo F, Pizzul L, Castillo MD, Gonzalez ME, Cea M, et al. 2010. Degradation of polycyclic aromatic
hydrocarbons by free and nanoclay-immobilized manganese peroxidase from Anthracophyllum discolor.
Chemosphere 80:271–78
Aktas¸N,C¸ic¸ek H, ¨
Unal AT, Kibarer G, Kolankaya N, Tanyolac¸ A. 2001. Reaction kinetics for laccase-catalyzed
polymerization of 1-naphthol. Bioresour. Techonol. 80:29–36
www.annualreviews.org Ecoenzymatic Stoichiometry 335
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Allen AP, Gillooly JF. 2009. Towards an integration of ecological stoichiometry and the metabolic theory of
ecology to better understand nutrient cycling. Ecol. Lett. 12:369–84
Allison SD, Gartner T, Holland K, Weintraub M, Sinsabaugh RL. 2007. Soil enzymes: linking proteomics and
ecological process. In Manual of Environmental Microbiology, pp. 704–11. Washington, DC: ASM Press
Allison SD, Vitousek PM. 2004. Extracellular enzyme activities and carbon chemistry as drivers of tropical
plant litter decomposition. Biotropica 36:285–96
Allison SD, Vitousek PM. 2005. Responses of extracellular enzymes to simple and complex nutrient inputs.
Soil Biol. Biochem. 37:937–44
Ambus P. 1993. Control of denitrification enzyme-activity in a streamside soil. FEMS Microbiol. Ecol. 102:225–
34
Annuar MSM, Adnan S, Vikineswary S, Chisti Y. 2009. Kinetics and energetics of azo dye decolorization by
Pycnoporus sanguineus.Water Air Soil Pollut. 202:179–88
Arnosti C. 2011. Microbial extracellular enzymes and the marine carbon cycle. Annu. Rev. Mar. Sci. 3:401–25
Azam F, Hodson RE. 1981. Multiphasic kinetics for D-glucose uptake by assemblages of natural marine
bacteria. Mar. Ecol. Prog. Ser. 6:213–22
Baldrian P. 2006. Fungal laccases—occurrence and properties. FEMS Microbiol. Rev. 30:215–42
Baltar F, Aristegui J, Sintes E, van Aken HM, Gasol JM, Herndl GJ. 2009. Prokaryotic extracellular enzymatic
activity in relation to biomass production and respiration in the meso- and bathypelagic waters of the
(sub)tropical Atlantic. Environ. Microbiol. 11:1998–2014
Beil S, Kehrli H, James P, Staudenmann W, Cook AM, et al. 1995. Purification and characterization of
the arylsulfatase synthesized by Pseudomonas aeruginosa PAO during growth in sulfate-free medium and
cloning of the arylsulfatase gene (atsA). Eur. J. Biochem. 29:385–94
Bellinger BJ, Gretz MR, Domozych DS, Kiemle SN, Hagerthey SE. 2010. Composition of extracellular
polymeric substances from periphyton assemblages in the Florida Everglades. J. Phycol. 46:484–96
Berg G, McClaugherty CM. 2010. Plant Litter: Decomposition, Humus Formation, Carbon Sequestration. Berlin:
Springer
Berman T. 1970. Alkaline phosphatases and phosphorus availability in Lake Kinneret. Limnol. Oceanogr.
15:663–74
Billen G. 1991. Protein degradation in aquatic environments. See Chr´
ost 1991, pp. 123–43
Blank RR. 2002. Amidohydrolase activity, soil nitrogen status, and the invasive crucifer Lepidium latifolium.
Plant Soil 239:155–63
Boucher D, Debroas D. 2009. Impact of environmental factors on couplings between bacterial community
composition and ectoenzymatic activities in a lacustrine ecosystem. FEMS Microbiol. Ecol. 70:66–78
Bremner JM, Mulvaney RL. 1978. Urease activity in soils. See Burns 1978, pp. 149–96
Brown JH, Gillooly JF, Allen AP, Savage VM, West GB. 2004. Toward a metabolic theory of ecology. Ecology
85:1771–89
Bugg TDH, Ahmad M, Hardiman EM, Singh R. 2011. The emerging role for bacteria in lignin degradation
and bio-product formation. Curr. Opin. Biotechnol. 22:394–400
B¨
unemann EK. 2008. Enzyme additions as a tool to assess the potential bioavailability of organically bound
nutrients. Soil Biol. Biochem. 40:2116–29
Burns RG, ed. 1978. Soil Enzymes. New York: Academic
Burns RG, Dick RP, eds. 2002. Enzymes in the Environment.NewYork:MarcelDekker
Button DK. 1993. Nutrient-limited microbial-growth kinetics—overview and recent advances. Antonie van
Leeuwenhoek Int. J. Gen. Mol. Microbiol. 63:225–35
Caruso G, Monticelli L, Azzaro F, Azzaro M, Decembrini F, et al. 2005. Dynamics of extracellular enzymatic
activities in a shallow Mediterranean ecosystem (Tindari ponds, Sicily). Mar. Freshwater Res. 56:173–88
Caruso G. 2010. Leucine aminopeptidase, beta-glucosidase and alkaline phosphatase activity rates and their
significance in nutrient cycles in some coastal Mediterranean sites. Mar. Drugs 8:916–40
Celussi M, Cataletto B, Umani SF, Del Negro P. 2009a. Depth profiles of bacterioplankton assemblages and
their activities in the Ross Sea. Deep-Sea Res. Part I: Oceanogr. Res. Pap. 56:2193–205
Celussi M, Paoli A, Crevatin E, Bergamasco A, Margiotta F, et al. 2009b. Short-term under-ice variability
of prokaryotic plankton communities in coastal Antarctic waters (Cape Hallett, Ross Sea). Estuar. Coast.
Shelf Sci. 81:491–500
336 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Characklis WG, Marshall K. 1990. Biofilms: A Basis for an Interdisciplinary Approach. New York: Wiley-Intersci.
Chen CY, Christensen ER. 1985. A unified theory for microbial growth under multiple nutrient limitation.
Water Res. 19:791–98
Cherif M, Loreau M. 2007. Stoichiometric constraints on resource use, competitive interactions, and elemental
cycling in microbial decomposers. Am. Nat. 169:709–24
Christian JR, Karl DM.1995. Bacterial ectoenzymes in marine waters: activity ratios and temperature responses
in three oceanographic provinces. Limnol. Oceanogr. 40:1042–49
Chr´
ost RJ, ed. 1991. Microbial Enzymes in Aquatic Environments. New York: Springer-Verlag
Chr´
ost RJ, Overbeck J. 1987. Kinetics of alkaline phosphatase and phosphorus availability for phytoplankton
and bacterioplankton in Lake Plussee. Microb. Ecol. 13:229–48
Cleveland CC, Liptzin D. 2007. C:N:P stoichiometry in soil: Is there a “Redfield ratio” for the microbial
biomass? Biogeochemistry 85:235–52
Cole JJ, Prairie YT, Caraco NF, McDowell WH, Tranvik LJ, et al. 2007. Plumbing the global carbon cycle:
integrating inland waters into the terrestrial carbon budget. Ecosystems 10:171–84
Collins SL, Sinsabaugh RL, Crenshaw C, Green LE, Porras-Alfaro A, et al. 2008. Pulse dynamics and microbial
processes in aridland ecosytems. J. Ecol. 96:413–20
Cunningham HW, Wetzel RG. 1989. Kinetic analysis of protein degradation by a freshwater wetland sediment
community. Appl. Environ. Microbiol. 55:1963–67
Dahiya N, Tewari R, Hoondal GS. 2006. Biotechnological aspects of chitinolytic enzymes: a review.
Appl. Microbiol. Biotechnol. 71:773–82
Danger M, Daufresne T, Lucas F, Pissard S, Lacroix G. 2008. Does Liebig’s law of the minimum scale up
from species to communities? Oikos 117:1741–51
Davidson EA, Jannsens IA. 2006. Temperature sensitivity of soil carbon decomposition and feedbacks to
climate change. Nature 440:165–73
Davidson EA, Samanta S, Caramori SS, Savage K. 2012. The dual Arrhenius and Michaelis-Menten kinetics
model for decomposition of soil organic matter at hourly to seasonal time scales. Glob. Chang. Biol.
18:371–84
DeAngelis KM, Lindow SW, Firestone MK. 2008. Bacterial quorum sensing and nitrogen cycling in rhizo-
sphere soil. FEMS Microbiol. Ecol. 66:197–207
Decho AW. 1990. Microbial exopolymer secretions in ocean environments: their role(s) in food webs and
marine processes. Oceanogr. Mar. Biol. Annu. Rev. 28:73–153
del Giorgio PA, Cole JJ. 1998. Bacterial growth efficiency in natural aquatic ecosystems. Annu. Rev. Ecol. Syst.
29:503–41
DeLong JP, Okie JG, Moses ME, Sibly RM, Brown JH. 2010. Shifts in metabolic scaling, production and
efficiency across major evolutionary transitions of life. Proc. Natl. Acad. Sci. USA 107:12941–45
Di Nardo C, Cinquegrana A, Papa S, Fuggi A, Fioretto A. 2004. Laccase and peroxidase isoenzymes during
leaf litter decomposition of Quercus ilex in a Mediterranean ecosystem. Soil Biol. Biochem. 36:1539–44
Dick R, ed. 2012. Methods of Soil Enzymology. Madison, WI: Soil Sci. Soc. Am.
Doi H, Cherif M, Iwabuchi T, Katano I, Stegen JC, Striebel M. 2010. Integrating elements and energy
through the metabolic dependencies of gross growth efficiency and the threshold elemental ratio. Oikos
119:752–65
Droop MR. 1977. Approach to quantitative nutrition of phytoplankton. J. Protozool. 24:528–32
Elser JJ, Acharya K, Kyle M, Cotner J, Makino W, et al. 2003. Growth rate-stoichiometry couplings in diverse
biota. Ecol. Lett. 6:936–43
Elsgaard L, Vinther FR. 2004. Modeling of the fine-scale temperature response of arylsulfatase activity in soil.
J. Plant Nutr. Soil Sci. 167:196–201
Enowashu E, Poll C, Lamersdorf N, Kandeler E. 2009. Microbial biomass and enzyme activities under reduced
nitrogen deposition in a spruce forest soil. Appl. Soil Ecol. 43:11–21
Ericksson K-E, Blanchette RA, Ander P. 1990. Microbial and Enzymatic Degradation of Wood Components. Berlin:
Springer. 407 pp.
Findlay S, Sinsabaugh RL, Fischer DT, Franchini P. 1998. Sources of dissolved organic carbon supporting
planktonic bacterial production in the tidal freshwater Hudson River. Ecosystems 1:227–39
www.annualreviews.org Ecoenzymatic Stoichiometry 337
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Foreman CM, Franchini P, Sinsabaugh RL. 1998. The trophic dynamics of riverine bacterioplankton: rela-
tionships among substrate availability, ectoenzyme kinetics and growth. Limnol. Oceanogr. 43:1344–52
Frankenberger WT, Tabatabai MA. 1991a. L-asparaginase activity of soils. Biol. Fertil. Soils 11:6–12
Frankenberger WT, Tabatabai MA. 1991b. L-glutaminase activity of soils. Soil Biol. Biochem. 23:869–74
Franklin O, Hall EK, Kaiser C, Battin TJ, Richter A. 2011. Optimization of biomass composition explains
microbial growth-stoichiometry relationships. Am. Nat. 177:E29–42
Freeman C, Ostle NJ, Fenner N, Kang H. 2004. A regulatory role for phenol oxidase during decomposition
in peatlands. Soil Biol. Biochem. 36:1663–67
Freeman C, Ostle N, Kang H. 2001. An enzymic ‘latch’ on a global carbon store—a shortage of oxygen locks
up carbon in peatlands by restraining a single enzyme. Nature 409:149–53
Frey SD, Gupta VVSR, Elliott ET, Paustian K. 2001. Protozoan grazing affects estimates of carbon utilization
efficiency of the soil microbial community. Soil Biol. Biochem. 33:1759–68
Frost PC, Benstead JP, Cross WF, Hillebrand H, Larson JH, et al. 2006. Threshold elemental ratios of carbon
and phosphorus in aquatic consumers. Ecol. Lett. 9:774–79
Fuhrman JA, Azam F. 1982. Thymidine incorporation as a measure of heterotrophic bacterial production in
marine surface waters—evaluation and field results. Mar. Biol. 66:109–20
Geisseler D, Doane TA, Horwath WR. 2009. Determining potential glutamine synthetase and glutamate
dehydrogenase activity in soil. Soil Biol. Biochem. 41:1741–49
Gfeller D, De Los Rios P, Caflisch A, Rao F. 2007. Complex network analysis of free-energy landscapes. Proc.
Natl. Acad. Sci. USA 104:1817–22
Gillooly JF, Allen AP, Brown JH, Elser JJ, del Rio CM, et al. 2005. The metabolic basis of whole-organism
RNA and phosphorus content. Proc. Natl. Acad. Sci. USA 102:11923–27
Gillooly JF, Brown JH, West GB, Savage VM, Charnov EL. 2001. Effects of size and temperature on metabolic
rate. Science 293:2248–51
Grossart HP, Simon M. 2002. Bacterioplankton dynamics in the Gulf of Aqaba and the Northern Red Sea in
early spring. Mar. Ecol. Prog. Ser. 239:263–76
Gusewell S, Gessner MO. 2009. N:P ratios influence litter decomposition and colonization by fungi and
bacteria in microcosms. Funct. Ecol. 23:211–19
Hall EK, Maixner F, Franklin O, Daims H, Richter A, Battin T. 2011. Linking microbial and ecosystem
ecology using ecological stoichiometry: a synthesis of conceptual and emprical approaches. Ecosystems
14:261–73
Hall-Stoodley L, Costerton JW, Stoodley P. 2004. Bacterial biofilms: from the natural environment to infec-
tious diseases. Nat. Rev. Microbiol. 2:95–108
Hasan HAH. 2000. Ureolytic organisms and soil fertility: a review. Comm. Soil Sci. Plant Anal. 31:2565–89
Hendel B, Marxsen J. 1997. Measurement of low-level extracellular enzyme activity in natural waters using
fluorigenic model substrates. Acta Hydrochim. Hydrobiol. 25:253–58
Hernandez DL, Hobbie SE. 2010. The effects of substrate composition, quantity and diversity on microbial
activity. Plant Soil 335:397–411
Herron PM, Stark JM, Holt C, Hooker T, Cardon ZG. 2009. Microbial growth efficiencies across a soil
moisture gradient assessed using 13C-acetic acid vapor and 15 N-ammonia gas. Soil Biol. Biochem. 41:1262–
69
Hill BH, Elonen CM, Jicha TM, Bolgrien DW, Moffett MF. 2010a. Sediment microbial enzyme activity as
an indicator of nutrient limitation in the great rivers of the Upper Mississippi River basin. Biogeochemistry
97:195–209
Hill BH, Elonen CM, Seifert LR, May AA, Tarquinio E. 2012. Ecoenzymatic stoichiometry and nutrient
limitation in US streams and rivers. Ecol. Indic. 18:540–55
Hill BH, McCormick FH, Harvey BC, Johnson SL, Warren ML, Elonen CM. 2010b. Microbial enzyme
activity, nutrient uptake and nutrient limitation in forested streams. Freshwater Biol. 55:1005–19
Hladyz S, Gessner MO, Giller PS, Pozo J, Woodward G. 2009. Resource quality and stoichiometric constraints
on stream ecosystem functioning. Freshw. Biol. 54:957–70
Hoch MP, Bronk DA. 2007. Bacterioplankton nutrient metabolism in the Eastern Tropical North Pacific.
J. Exp. Mar. Biol. Ecol. 349:390–404
338 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Hoch MP, Snyder RA, Jeffrey WH, Dillon KS, Coffin RB. 2006. Expression of glutamine synthetase and
glutamate dehydrogenase by marine bacterioplankton: assay optimizations and efficacy for assessing N
to carbon metabolic balance in situ. Limnol. Oceanogr. Methods 4:318–28
Hofmockel KS, Fierer N, Colman BP, Jackson RB. 2010. Amino acid abundance and proteolytic potential in
North American soils. Oecologia 163:1069–78
Hollibaugh JT, Azam F. 1983. Microbial degradation of dissolved proteins in seawater. Limnol. Oceanogr.
28:1104–16
Hoppe HG. 1983. Significance of exoenzymatic activities in the ecology of brackish water: measurements by
means of methylumbelliferyl substrates. Mar. Ecol. Prog. Ser. 11:299–308
Hoppe HG. 1991. Microbial extracellular enzyme activity: a new key parameter in aquatic ecology. See Chr ´
ost
1991, pp. 60–83
Hoppe HG, Kim SJ, Gocke K. 1988. Microbial decomposition in aquatic environments: combined processes
of extracellular enzyme activity and substrate uptake. Appl. Environ. Microbiol. 54:784–90
Hoppe HG, Ullrich S. 1999. Profiles of ectoenzymes in the Indian Ocean: phenomena of phosphatase activity
in the mesopelagic zone. Aquat. Microb. Ecol. 19:139–48
Houghton RA. 2007. Balancing the global carbon budget. Annu. Rev. Earth Planet. Sci. 35:313–47
Jackson C, Foreman C, Sinsabaugh RL. 1995. Microbial enzyme activities as indicators of organic matter
processing rates in a Lake Erie coastal wetland. Freshw. Biol. 34:329–42
Jones DL, Kielland K, Sinclair FL, Dahlgren RA, Newsham KK, et al. 2009. Soil organic ni-
trogen mineralization across a global latitudinal gradient. Glob. Biogeochem. Cycles 23:GB1016
(DOI:10.1029/2008GB003250)
Jones JG. 1972. Studies on freshwater microorganisms: phosphatase activity in lakes of differing degrees of
eutrophication. J. Ecol. 60:777–91
Jorgensen NOG, Kroer N, Coffin RB, Hoch MP. 1999. Relations between bacterial nitrogen metabolism and
growth efficiency in an estuarine and an open-water ecosystem. Aquat. Microb. Ecol. 18:247–61
Kahkonen MA, Wittmann C, Kurola J, Ilvesniemi H, Salkinoja-Salonen MS. 2001. Microbial activity of boreal
forest soil in a cold climate. Boreal Environ. Res. 6:19–28
Karner M, Rassoulzadegan F. 1995. Extracellular enzyme activity: indications for high short-term variability
in a coastal marine ecosystem. Microb. Ecol. 30:143–56
Kaziev FKH. 1975. Thermodynamic characteristics of enymic reactions in soil. Biol. Nauki 10:121–27
Kelley AM, Fay PF, Polley HW, Gill RA, Jackson RB. 2011. Atmospheric CO2and soil extracellular enzyme
activity: a meta-analysis and CO2gradient experiment. Ecosphere 2(8):art96 (DOI:10.1890/ES11-00117.1)
Kirchman DL. 2003. The contribution of monomers and other low-molecular weight compounds to the flux
of dissolved organic material in aquatic ecosystems. In Aquatic Ecosystems: The Interactivity of Dissolved
Organic Matter, ed. S Findlay, RL Sinsabaugh, pp. 218–39. San Diego: Academic
Kirchman DL, White J. 1999. Hydrolysis and mineralization of chitin in the Delaware estuary. Aquat. Microb.
Ecol. 18:187–96
Knorr M, Frey SD, Curtis PS. 2005. Nitrogen additions and litter decomposition: a meta-analysis. Ecology
86:3252–57
Kocabas DS, Bakir U, Phillips SEV, McPherson MJ, Ogel ZB. 2008. Purification, characterization, and iden-
tification of a novel bifunctional catalase-phenol oxidase from Scytalidium thermophilum.Appl. Microbiol.
Biotechnol. 79:407–15
Leppard GG. 1995. The characterization of algal and microbial mucilages and their aggregates in aquatic
ecosystems. Sci. Total Environ. 165:103–31
Lo SC, Ho YS, Buswell JA. 2001. Effect of phenolic monomers on the production of laccases by the edi-
ble mushroom Pleurotus sajor-caju, and partial characterization of a major laccase component. Mycologia
93:413–21
Lobry JR, Flandrois JP, Carret G, Pave A. 1992. Monod’s microbial growth model revisited. Bull. Math. Biol.
54:117–21
Lynd LR, Weimer PJ, van Zyl WH, Pretorius IS. 2002. Microbial cellulose utilization: fundamentals and
biotechnology. Microbiol. Mol. Biol. Rev. 66:506–77
Maire V, Alvarez G, Colombet J, Comby A, Despinasse R, et al. 2012. An unknown respiration pathway
substantially contributes to soil CO2emissions. Biogeosci.Discuss. 6:8663–91
www.annualreviews.org Ecoenzymatic Stoichiometry 339
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Manzoni S, Porporato A. 2009. Soil carbon and nitrogen mineralization: theory and models across scales. Soil
Biol. Biochem. 41:1355–79
Manzoni S, Taylor P, Richter A, Porporato A, ˚
Agren GI. 2012. Environmental and stoichiometric controls
on microbial carbon-use efficiency in soils. New Phytol. 196:79–91
Manzoni S, Trofymow JA, Jackson RB, Porporato A. 2010. Stoichiometric controls on carbon, nitrogen, and
phosphorus dynamics in decomposing litter. Ecol. Monogr. 80:89–106
Marklein AR, Houlton BZ. 2012. Nitrogen inputs accelerate phosphorus cycling rates across a wide variety
of terrestrial ecosystems. New Phytol. 193(3):696–704
Marzluf GA. 1997. Genetic regulation of nitrogen metabolism in the fungi. Microbol. Mol. Biol. Rev. 61:17–32
McClaugherty CA, Linkins AE. 1990. Temperature responses of enzymes in two forest soils. Soil Biol. Biochem.
22:29–33
McLaren AD. 1978. Kinetics and consecutive reactions of soil enzymes. See Burns 1978, pp. 97–116
Menezes-Blackburn D, Jorquera M, Gianfreda L, Rao M, Greiner R, et al. 2011. Activity stabilization of
Aspergillus niger and Escherichia coli phytases immobilized on allophanic synthetic compounds and mont-
morillonite nanoclays. Bioresour. Technol. 102:9360–67
Meyer-Reil LA. 1987. Seasonal and spatial distribution of extracellular enzymatic activities and microbial
incorporation of dissolved organic substrates in marine sediments. Appl. Environ. Microbiol. 53:1748–55
Michaelis L, Menten MI. 1913. Die Kinetik der Invertinwirkung. Biochemische Z. 49:333–69
Mobley HLT, Island MD, Hausinger RP. 1995. Molecular biology of microbial ureases. Microbiol. Rev. 59:451–
80
Monod J. 1949. The growth of bacterial cultures. Annu. Rev. Microbiol. 3:371–94
Monticelli LS, La Ferla R, Maimone G. 2003. Dynamics of bacterioplankton activities after a summer phyto-
plankton bloom period in Terra Nova Bay. Antarct. Sci. 15:85–93
Moorhead DL, Lashermes G, Sinsabaugh RL. 2012. A theoretical model of C- and N-acquiring exoenzyme
activities balancing microbial demands during decomposition. Soil Biol. Biochem. 53:133–41
Moorhead DL, Sinsabaugh RL. 2006. A theoretical model of litter decay and microbial interaction. Ecol.
Monogr. 76:151–74
Mudryk ZJ, Sk´
orczewski P. 2004. Extracellular enzyme activity at the air-water interface of an estuarine lake.
Estuar. Coast. Shelf Sci. 59:59–67
M¨
unster U. 1991. Extracellular enzyme activity in eutrophic and polyhumic lakes. See Chr ´
ost 1991, pp. 96–122
M¨
unster U, Nurminen J, Einio P, Overbeck J. 1992. Extracellular enzymes in a small polyhumic lake—origin,
distribution and activities. Hydrobiologia 243:47–59
Olander LP, Vitousek PM. 2000. Regulation of soil phosphatase and chitinase activity by N and P availability.
Biogeochemistry 49:175–90
Oshrain RL, Wiebe WJ. 1976. Arylsulfatase activity of salt marsh soils. Appl. Environ. Microbiol. 38:337–40
Overbeck J. 1975. Distribution pattern of uptake kinetic responses in a stratified eutrophic lake. Verh. Int.
Vereinigung Theor. Angew. Limnol. 19:2600–15
Overbeck J. 1991. Early studies on ecto- and extracellular enzymes in aquatic environments. See Chr ´
ost 1991,
pp. 1–5
Pamer E, Vujovic G, Knezevic P, Kojic D, Prvulovic D, et al. 2011. Water quality assessment in lakes of
Vojvodina. Int. J. Environ. Res. 5:891–900
Pereira S, Zille A, Micheletti E, Moradas-Ferreira P, De Philippis R, Tamagnini P. 2009. Complexity of
cyanobacterial exopolysaccharides: composition, structures, inducing factors and putative genes involved
in their biosynthesis and assembly. FEMS Microbiol. Rev. 33:917–41
Peterjohn WT. 1991. Denitrification—enzyme content and activity in desert soils. Soil Biol. Biochem. 23:845–55
Piontek J, Handel N, De Bodt C, Harlay J, Chou L, Engel A. 2011. The utilization of polysaccharides by
heterotrophic bacterioplankton in the Bay of Biscay (North Atlantic Ocean). J. Plankton Res. 33:1719–35
Rabinovich ML, Bolobova AV, Vasilchenko LG. 2004. Fungal decomposition of natural aromatic structures
and xenobiotics: a review. Appl. Biochem. Microbiol. 40:1–17
Ramirez KS, Craine JM, Fierer N. 2012. Consistent effects of nitrogen amendments on soil microbial com-
munities and processes across biomes. Glob. Change Biol. 18:1918–27
Ramirez-Martinez JR, McClaren AD. 1966. Some factors influencing the determination of phosphatase activity
in native soils and in soils sterilized br irradiation. Enzymologia 31:23–38
340 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Reichardt W, Overbeck J, Steubing L. 1967. Free dissolved enzymes in lake water. Nature 216:1345–47
Resat H, Bailey V, McCue LA, Konopka A. 2012. Modeling microbial dynamics in heterogeneous environ-
ments: growth on soil carbon sources. Microb. Ecol. 63:883–97
Rosenstock B, Simon M. 2003. Consumption of dissolved amino acids and carbohydrates by limnetic bacteri-
oplankton according to molecular weight fractions and proportions bound to humic matter. Microb. Ecol.
45:433–43
Schimel JP, Weintraub MN. 2003. The implications of exoenzyme activity on microbial carbon and nitrogen
limitation in soil: a theoretical model. Soil Biol. Biochem. 35:549–63
Seitzinger SP, Harrison JA, Dumont E, Beusen AHW, Bouwman AF. 2005. Sources and delivery of carbon, ni-
trogen and phosphorus to the coastal zone: an overview of global nutrient export from watersheds (NEWS)
models and their application. Glob. Biogeochem. Cycles 19:GB4S01 (DOI:10.1029/2005GB002606)
Shukla G, Varma A, eds. 2011. Soil Enzymology. Berlin: Springer-Verlag
Sierra CA. 2012. Temperature sensitivity of organic matter decomposition in the Arrhenius equation: some
theoretical considerations. Biogeochemistry 108:1–15
Simon M, Rosenstock B. 2007. Different coupling of dissolved amino acid, protein, and carbohydrate turnover
to heterotrophic picoplankton production in the southern ocean in austral summer and fall. Limnol.
Oceanogr. 52:85–95
Sinsabaugh RL. 2010. Phenol oxidase, peroxidase and organic matter dynamics of soil. Soil Biol. Biochem.
42:391–404
Sinsabaugh RL, Antibus RK, Linkins AE, McClaugherty CA, Rayburn L, et al. 1992. Wood decomposition
over a first-order watershed: mass loss as a function of lignocellulase activity. Soil Biol. Biochem. 24:743–49
Sinsabaugh RL, Antibus RK, Linkins AE, Rayburn L, Repert D, Weiland T. 1993. Wood decomposition:
nitrogen and phosphorus dynamics in relation to extracellular enzyme activity. Ecology 74:1586–93
Sinsabaugh RL, Benfield EF, Linkins AE. 1981. Cellulase actvity associated with the decomposition of leaf
litter in a woodland stream. Oikos 36:184–90
Sinsabaugh RL, Carreiro MM, Alvarez S. 2002. Enzyme and microbial dynamics during litter decomposition.
See Burns 2002, pp. 249–66
Sinsabaugh RL, Findlay S, Franchini P, Fischer D. 1997. Enzymatic analysis of riverine bacterioplankton
production. Limnol. Oceanogr. 42:29–38
Sinsabaugh RL, Follstad Shah JJ. 2010. Integrating resource utilization and temperature in metabolic scaling
of riverine bacterial production. Ecology 91:1455–65
Sinsabaugh RL, Follstad Shah JJ. 2011. Ecoenzymatic stoichiometry of recalcitrant organic matter decompo-
sition: the growth rate hypothesis in reverse. Biogeochemistry 102:31–43
Sinsabaugh RL, Follstad Shah JJ, Hill BH, Elonen CM. 2011. Ecoenzymatic stoichiometry of stream sediments
with comparison to terrestrial soils. Biogeochemistry. Houten, The Neth.: Springer (DOI:10.1007/s10533-
011-9676-x)
Sinsabaugh RL, Foreman CM. 2001. Activity profiles of bacterioplankton in a eutrophic river. Freshw. Biol.
46:1–12
Sinsabaugh RL, Hill BH, Follstad Shah JJ. 2009. Ecoenzymatic stoichiometry of microbial organic nutrient
acquisition in soil and sediment. Nature 462:795–98
Sinsabaugh RL, Lauber CL, Weintraub MN, Ahmed B, Allison SD, et al. 2008. Stoichiometry of soil enzyme
activity at global scale. Ecol. Lett. 11:1252–64
Sinsabaugh RL, Linkins AE. 1993. Statistical modeling of litter decomposition from integrated cellulase
activity. Ecology 74:1594–97
Sinsabaugh RL, Moorhead DL. 1994. Resource allocation to extracellular enzyme production: a model for
nitrogen and phosphorus control of litter decomposition. Soil Biol. Biochem. 26:1305–11
Sinsabaugh RL, Osgood MP, Findlay S. 1994. Enzymatic models for estimating decomposition rates of
particulate detritus. J. North Am. Benthol. Soc. 13:160–69
Sinsabaugh RL, Van Horn DJ, Follstad Shah JJ, Findlay S. 2010. Ecoenzymatic stoichiometry in relation to
productivity for freshwater biofilm and plankton communities. Microb. Ecol. 60:885–93
Skujin¸ˇ
s J. 1978. History of abiontic soil enzyme research. See Burns 1978, pp. 1–49
www.annualreviews.org Ecoenzymatic Stoichiometry 341
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Snajdr J, Cajthaml T, Valaskova V, Merhautova V, Petrankova M, et al. 2011. Transformation of Quercus petraea
litter: successive changes in litter chemistry are reflected in differential enzyme activity and changes in
the microbial community composition. FEMS Microbiol. Ecol. 75:291–303
Someville M, Billen G. 1983. A method for determining exoproteolytic activity in natural waters. Limnol.
Oceanogr. 28:190–93
Speir TW, Ross DJ. 1978. Soil phosphatases and sulphatase. See Burns 1978, pp. 197–250
Sterner RW, Elser JJ. 2002. Ecological Stoichiometry: The Biology of Elements from Molecules to the Biosphere.
Princeton: Princeton Univ.
Stone MM, Weiss MS, Goodale CL, Adams MB, Fernandez IJ, et al. 2012. Temperature sensitivity of soil
enzyme kinetics under N-fertilization in two temperate forests. Glob. Chang. Biol. 18(3):1173–84
Strong PJ, Claus H. 2011. Laccase: a review of its past and its future in bioremediation. Crit. Rev. Environ.
Sci.Technol. 41:373–434
Stursova M, Crenshaw C, Sinsabaugh RL. 2006. Microbial responses to long-term N deposition in a semiarid
grassland. Microb. Ecol. 51:90–98
Tabatabai MA, Bremner JM. 1970. Arylsulfatase activity of soils. Soil Sci. Soc. Am. Proc. 34:225–29
Taylor GT, Way J, Yu Y, Scranton MI. 2003. Ectohydrolase activity in surface waters of the Hudson River
and western Long Island Sound estuaries. Mar. Ecol. Prog. Series 263:1–15
Theuerl S, Buscot F. 2010. Laccases: toward disentangling their diversity and functions in relation to soil
organic matter cycling. Biol. Fertil. Soils 46:215–25
Trasar-Cepeda C, Gil-Sotres F, Leiros MC. 2007. Thermodynamic parameters of enzymes in grassland soils
from Galicia, NW Spain. Soil Biol. Biochem. 39:311–19
Trasar-Cepeda C, Hern´
andez T, Garcia C, Rad C. 2011. Soil Enzymology in the Recycling of Organic Wastes and
Environmental Restoration. Berlin: Springer
Treseder KK. 2008. Nitrogen additions and microbial biomass: a meta-analysis of ecosystem studies.
Ecol. Lett. 11:1111–20
Valtcheva E, Veleva S, Radeva G, Valtchev I. 2003. Enzyme action of the laccase-mediator system in the pulp
delignification process. React. Kinet. Catal. Lett. 78:183–91
Van Horn DJ, Sinsabaugh RL, Takacs-Vesbach CD, Mitchell KR, Dahm CN. 2011. The response of het-
erotrophic stream biofilm communities to a resource gradient. Aquat. Microb. Ecol. 64:149–61
Vrba J, Callier C, Bittl T, Simek K, Bertoni R, et al. 2004. Are bacteria the major producers of extracellular
glycolytic enzymes in aquatic environments? Int. Rev. Hydrobiol. 89:102–17
Wanek W, Mooshammer M, Blochl A, Hanreich A, Richter A. 2010. Determination of gross rates of amino
acid production and immobilization in decomposing leaf litter by a novel N-15 isotope pool dilution
technique. Soil Biol. Biochem. 42:1293–302
Wang G, Post WM, Mayes MA. 2012a. Development of microbial-enzyme-mediated decomposition model
parameters through steady-state and dynamic analyses. Ecol. Appl. In press. (DOI:10.1890/12-0681.1)
Wang G, Post WM, Mayes MA, Frerichs JT, Sindhu J. 2012b. Parameter estimation for models of ligninolytic
and cellulolytic enzyme kinetics. Soil Biol. Biochem. 48:28–38
Warren RAJ. 1996. Microbial hydrolysis of polysaccharides. Annu. Rev. Microbiol. 50:183–212
Wetzel RG. 1981. Longterm dissolved and particulate alkaline phosphatase activity in a hardwater lake in
relation to lake stability and phosphorus enrichments. Verhandlungen Int. Vereinigung Theor. Angwandte
Limnol. 21:337–39
Wetzel RG. 1991. Extracellular enzymatic interactions: storage, redistribution and interspecific communica-
tion. See Chr´
ost 1991, pp. 6–28
Williams CJ, Scott AB, Wilson HF, Xenopoulos MA. 2012. Effects of land use on water column bacterial
activity and enzyme stoichiometry in stream ecosystems. Aquat. Sci. 74(3):483–94
Williams PJL. 1973. The validity of the application of simple kinetic analysis to heterogeneous microbial
populations. Limnol. Oceanogr. 18:159–65
Zeglin LH, Kluber LA, Myrold DD. 2012. The importance of amino sugar turnover to C and N cycling in
organic horizons of old growth Douglas-fir forest soils colonized by ectomycorrhizal mats. Biogeochemistry.
In press (DOI:10.1007/s10533-012-9746-8)
Zeglin LH, Stursova M, Sinsabaugh RL, Collins SL. 2007. Microbial responses to nitrogen addition in three
contrasting grassland ecosystems. Oecologia 296:65–75
342 Sinsabaugh ·Follstad Shah
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43CH15-Sinsabaugh ARI 26 September 2012 11:56
Zhang JB, Liu XP, Xu ZQ, Chen H, Yang YX. 2008. Degradation of chlorophenols catalyzed by laccase. Int.
Biodeterior. Biodegrad. 61:351–56
Zhang XD, Amelung W. 1996. Gas chromatographic determination of muramic acid, glucosamine, man-
nosamine, and galactosamine in soils. Soil Biol. Biochem. 28:1201–6
Zinn M, Witholt B, Egli T. 2004. Dual nutrient limited growth: models, experimental obsevations, and
applications. J. Biotechnol. 113:263–79
ZoBell CE. 1943. The effect of solid surfaces upon bacterial activity. J. Bacteriol. 46:39–56
www.annualreviews.org Ecoenzymatic Stoichiometry 343
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43-FrontMatter ARI 1 October 2012 13:46
Annual Review of
Ecology, Evolution,
and Systematics
Volume 43, 2012
Contents
Scalingy Up in Ecology: Mechanistic Approaches
Mark Denny and Lisandro Benedetti-Cecchi ppppppppppppppppppppppppppppppppppppppppppppppppp1
Adaptive Genetic Variation on the Landscape: Methods and Cases
Sean D. Schoville, Aur´elie Bonin, Olivier Fran¸cois, St´ephane Lobreaux,
Christelle Melodelima, and St´ephanie Manel pppppppppppppppppppppppppppppppppppppppppppppp23
Endogenous Plant Cell Wall Digestion: A Key Mechanism
in Insect Evolution
Nancy Calder´on-Cort´es, Mauricio Quesada, Hirofumi Watanabe,
Horacio Cano-Camacho, and Ken Oyama ppppppppppppppppppppppppppppppppppppppppppppppppp45
New Insights into Pelagic Migrations: Implications for Ecology
and Conservation
Daniel P. Costa, Greg A. Breed, and Patrick W. Robinson ppppppppppppppppppppppppppppppppp73
The Biogeography of Marine Invertebrate Life Histories
Dustin J. Marshall, Patrick J. Krug, Elena K. Kupriyanova, Maria Byrne,
and Richard B. Emlet ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp97
Mutation Load: The Fitness of Individuals in Populations Where
Deleterious Alleles Are Abunduant
Aneil F. Agrawal and Michael C. Whitlock pppppppppppppppppppppppppppppppppppppppppppppp115
From Animalcules to an Ecosystem: Application of Ecological Concepts
to the Human Microbiome
Noah Fierer, Scott Ferrenberg, Gilberto E. Flores, Antonio Gonz´alez,
Jordan Kueneman, Teresa Legg, Ryan C. Lynch, Daniel McDonald,
Joseph R. Mihaljevic, Sean P. O’Neill, Matthew E. Rhodes, Se Jin Song,
and William A. Walters ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp137
Effects of Host Diversity on Infectious Disease
Richard S. Ostfeld and Felicia Keesing ppppppppppppppppppppppppppppppppppppppppppppppppppppp157
Coextinction and Persistence of Dependent Species in a Changing World
Robert K. Colwell, Robert R. Dunn, and Nyeema C. Harris ppppppppppppppppppppppppppppp183
Functional and Phylogenetic Approaches to Forecasting Species’ Responses
to Climate Change
Lauren B. Buckley and Joel G. Kingsolver pppppppppppppppppppppppppppppppppppppppppppppppp205
v
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
ES43-FrontMatter ARI 1 October 2012 13:46
Rethinking Community Assembly through the Lens of Coexistence Theory
J. HilleRisLambers, P.B. Adler, W.S. Harpole, J.M. Levine, and M.M. Mayfield ppppp227
The Role of Mountain Ranges in the Diversification of Birds
Jon Fjelds˚a, Rauri C.K. Bowie, and Carsten Rahbek ppppppppppppppppppppppppppppppppppppp249
Evolutionary Inferences from Phylogenies: A Review of Methods
Brian C. O’Meara pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp267
A Guide to Sexual Selection Theory
Bram Kuijper, Ido Pen, and Franz J. Weissing pppppppppppppppppppppppppppppppppppppppppp287
Ecoenzymatic Stoichiometry and Ecological Theory
Robert L. Sinsabaugh and Jennifer J. Follstad Shah ppppppppppppppppppppppppppppppppppppp313
Origins of New Genes and Evolution of Their Novel Functions
Yun Ding, Qi Zhou, and Wen Wang ppppppppppppppppppppppppppppppppppppppppppppppppppppp345
Climate Change, Aboveground-Belowground Interactions,
and Species’ Range Shifts
Wim H. Van der Putten pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp365
Inflammation: Mechanisms, Costs, and Natural Variation
Noah T. Ashley, Zachary M. Weil, and Randy J. Nelson pppppppppppppppppppppppppppppppp385
New Pathways and Processes in the Global Nitrogen Cycle
Bo Thamdrup pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp407
Beyond the Plankton Ecology Groug (PEG) Model: Mechanisms Driving
Plankton Succession
Ulrich Sommer, Rita Adrian, Lisette De Senerpont Domis, James J. Elser,
Ursula Gaedke, Bas Ibelings, Erik Jeppesen, Miquel L¨urling, Juan Carlos Molinero,
Wolf M. Mooij, Ellen van Donk, and Monika Winder ppppppppppppppppppppppppppppppppp429
Global Introductions of Crayfishes: Evaluating the Impact of Species
Invasions on Ecosystem Services
David M. Lodge, Andrew Deines, Francesca Gherardi, Darren C.J. Yeo,
Tracy Arcella, Ashley K. Baldridge, Matthew A. Barnes, W. Lindsay Chadderton,
Jeffrey L. Feder, Crysta A. Gantz, Geoffrey W. Howard, Christopher L. Jerde,
Brett W. Peters, Jody A. Peters, Lindsey W. Sargent, Cameron R. Turner,
Marion E. Wittmann, and Yiwen Zeng pppppppppppppppppppppppppppppppppppppppppppppppp449
Indexes
Cumulative Index of Contributing Authors, Volumes 39–43 ppppppppppppppppppppppppppp473
Cumulative Index of Chapter Titles, Volumes 39–43 ppppppppppppppppppppppppppppppppppp477
Errata
An online log of corrections to Annual Review of Ecology, Evolution, and Systematics
articles may be found at http://ecolsys.annualreviews.org/errata.shtml
vi Contents
Annu. Rev. Ecol. Evol. Syst. 2012.43:313-343. Downloaded from www.annualreviews.org
by University of New Mexico (UNM) on 12/04/12. For personal use only.
... Higher temperatures are expected to increase rates of microbial respiration in predictable ways, based on metabolic scaling theory (MST) that describes how the rate of most cellular reactions increases exponentially with temperature (Allen et al., 2005;Arrhenius, 1915;Michaletz & Garen, 2024). Microbial respiration under nutrient limited conditions may require more enzymatic steps and possibly yield a greater temperature dependence (i.e., the activation energy of wood microbial respiration: E [in electron volts]) under equal conditions of humidity (Sinsabaugh & Follstad Shah, 2012). However, recent analyses of global wood decomposition showed that nutrient limitation instead yielded a lower temperature dependence, with a range of E ≈ 0.16-0.43 ...
... According to the enzyme vector model (Moorhead et al., 2016;Sinsabaugh & Follstad Shah, 2012), we calculated the activity vector lengths and angles for all enzyme data using Microsoft Excel (Office for Windows 2019). C/(C + N) versus C/(C + P) could quantify relative C versus nutrient investments as the vector length, and P versus N investment was determined as the vector angle. ...
... E for microbial exoenzymes that degrade organic macromolecules, including lignin and cellulose, are between~0.31 and 0.56 eV (Sinsabaugh & Follstad Shah, 2012;Wang et al., 2012). Wood P concentration (in milligrams of phosphorus per gram of dry wood mass) and P:C ratio can be used as an additional line of evidence for wood P availability effects on microbial respiration in the models (Appendix S1: Equations S1 and S2). ...
Article
Full-text available
Fungi are key decomposers of deadwood, but the impact of anthropogenic changes in nutrients and temperature on fungal community and its consequences for wood microbial respiration are not well understood. Here, we examined how nitrogen and phosphorus additions (field experiment) and warming (laboratory experiment) together influence fungal composition and microbial respiration from decomposing wood of angiosperms and gymnosperms in a subtropical forest. Nutrient additions significantly increased wood microbial respiration via fungal composition, but effects varied with nutrient types and taxonomic groups. Specifically, phosphorus addition significantly increased wood microbial respiration (65%) through decreased acid phosphatase activity and increased abundance of fast-decaying fungi (e.g., white rot), while nitrogen addition marginally increased it (30%). Phosphorus addition caused a greater increase in microbial respiration in gymnosperms than in angiosperms (83.3% vs. 46.9%), which was associated with an increase in Basidiomycota:Ascomycota operational taxonomic unit abundance in gymnosperms but a decrease in angiosperms. The temperature dependencies of microbial respiration were remarkably constant across nutrient levels, consistent with metabolic scaling theory hypotheses. This is because there was no significant interaction between temperature and wood phosphorus availability or fungal composition, or the interaction among the three factors. Our results highlight the key role of tree identity in regulating nutrient response of wood microbial respiration through controlling fungal composition. Given that the range of angiosperm species may expand under climate warming and forest management, our data suggest that expansion will decrease nutrient effects on forest carbon cycling in forests previously dominated by gymnosperm species.
... Higher temperatures are expected to increase rates of microbial respiration in predictable ways, based on metabolic scaling theory (MST) that describes how the rate of most cellular reactions increases exponentially with temperature (Allen et al., 2005;Arrhenius, 1915;Michaletz & Garen, 2024). Microbial respiration under nutrient limited conditions may require more enzymatic steps and possibly yield a greater temperature dependence (i.e., the activation energy of wood microbial respiration: E [in electron volts]) under equal conditions of humidity (Sinsabaugh & Follstad Shah, 2012). However, recent analyses of global wood decomposition showed that nutrient limitation instead yielded a lower temperature dependence, with a range of E ≈ 0.16-0.43 ...
... According to the enzyme vector model (Moorhead et al., 2016;Sinsabaugh & Follstad Shah, 2012), we calculated the activity vector lengths and angles for all enzyme data using Microsoft Excel (Office for Windows 2019). C/(C + N) versus C/(C + P) could quantify relative C versus nutrient investments as the vector length, and P versus N investment was determined as the vector angle. ...
... E for microbial exoenzymes that degrade organic macromolecules, including lignin and cellulose, are between~0.31 and 0.56 eV (Sinsabaugh & Follstad Shah, 2012;Wang et al., 2012). Wood P concentration (in milligrams of phosphorus per gram of dry wood mass) and P:C ratio can be used as an additional line of evidence for wood P availability effects on microbial respiration in the models (Appendix S1: Equations S1 and S2). ...
Article
Full-text available
Fungi are key decomposers of deadwood, but the impact of anthropogenic changes in nutrients and temperature on fungal community and its consequences for wood microbial respiration are not well understood. Here, we examined how nitrogen and phosphorus additions (field experiment) and warming (laboratory experiment) together influence fungal composition and microbial respiration from decomposing wood of angiosperms and gymnosperms in a subtropical forest. Nutrient additions significantly increased wood microbial respiration via fungal composition, but effects varied with nutrient types and taxonomic groups. Specifically, phosphorus addition significantly increased wood microbial respiration (65%) through decreased acid phosphatase activity and increased abundance of fast‐decaying fungi (e.g., white rot), while nitrogen addition marginally increased it (30%). Phosphorus addition caused a greater increase in microbial respiration in gymnosperms than in angiosperms (83.3% vs. 46.9%), which was associated with an increase in Basidiomycota:Ascomycota operational taxonomic unit abundance in gymnosperms but a decrease in angiosperms. The temperature dependencies of microbial respiration were remarkably constant across nutrient levels, consistent with metabolic scaling theory hypotheses. This is because there was no significant interaction between temperature and wood phosphorus availability or fungal composition, or the interaction among the three factors. Our results highlight the key role of tree identity in regulating nutrient response of wood microbial respiration through controlling fungal composition. Given that the range of angiosperm species may expand under climate warming and forest management, our data suggest that expansion will decrease nutrient effects on forest carbon cycling in forests previously dominated by gymnosperm species.
... Enzymes such as β-1,4-glucosidase (BG), β-1,4-N-acetylglucosaminidase (NAG), leucine aminopeptidase (LAP), and acid phosphatase (ACP)-which are vital components of the C, N, and P cycles-have been extensively employed in investigations into the stoichiometry of extracellular enzyme activities [21][22][23][24]. Sinsabaugh et al. reported that the ecological enzyme stoichiometry of soils and sediments in geoecosystems is approximately 1:1:1 [25]. However, Luo et al. observed differences in enzyme C:N:P ratios in peatlands on the Ruoergai Plateau compared to the expected 1:1:1 ratio, with ratios closer to soil C:N:P ratios, indicating phosphorus (P) limitation in the soils of this region [26]. ...
... The stoichiometric characterization of the extracellular enzyme activities C:N, C:P, and N:P are denoted as C:Ne, C:Pe, and N:Pe, respectively. These ratios were calculated using the following equations [25]: ...
... The stoichiometry of extracellular enzyme activity can serve as an indicator of the dynamic equilibrium between nutrient availability [53,54]. In a meta-analysis conducted by Sinsabaugh, it was discovered that the stoichiometry of C, N, and P acquisition enzyme activities at a global scale, as represented by ln(BG):ln(LAP + NAG):ln(ACP) ratios, was approximately 1:1:1 [25]. Peng et al. found this pattern in tropical ecosystems [55]. ...
Article
Full-text available
Rational application of N fertilizer is essential for maintaining the long-term productivity of Moso bamboo forests. Microbial activity is a crucial indicator of soil quality. Changes in soil nutrient resources due to N addition can lead to microbial nutrient limitations, thereby impeding the maintenance of soil quality. Currently, there is limited research on the effects of N application on microbial nutrient limitations in Moso bamboo forest soils. To examine the changes in extracellular enzyme activity and microbial nutrient limitations in Moso bamboo forest soils following N application, we conducted an N application experiment in northern Guizhou. The findings revealed that the N3 treatment (726 kg·N·hm−2·yr−1) significantly reduced β-glucosidase (BG) activity by 27.61% compared to the control group (no fertilization). The N1 (242 kg·N·hm−2·yr−1), N2 (484 kg·N·hm−2·yr−1), and N3 treatments notably increased the activities of leucine aminopeptidase (LAP) and N-acetyl-β-D-glucosidase (NAG) by 11.45% to 15.79%. Acid phosphatase (ACP) activity remained unaffected by fertilization. N application treatments significantly decreased the C:Ne and C:Pe ratios, while the N:Pe ratio was less influenced by N fertilizer application. Scatter plots and vector characteristics of enzyme activity stoichiometry suggested that microorganisms in the study area were limited by C and N, and N fertilizer application reduced the vector length and increased the vector angle, indicating that N application alleviated the C and N limitation of microorganisms in Moso bamboo forests. Redundancy Analysis (RDA) demonstrated that microbial biomass phosphorus (MBP) was the most critical factor affecting extracellular enzyme activity and stoichiometry. Furthermore, Random Forest Regression analysis identified MBP and the N:Pm ratio as the most significant factors influencing microbial C and N limitation, respectively. The study demonstrated that N application modulates the microbial nutrient acquisition strategy by altering soil nutrient resources in Moso bamboo forests. Formulating fertilizer application strategies based on microbial nutrient requirements is more beneficial for maintaining soil quality and sustainably managing Moso bamboo forests. Additionally, our study offers a theoretical reference for understanding carbon cycling in bamboo forest ecosystems in the context of substantial N inputs.
... Microbial CUE was assessed using the updated biogeochemical equilibrium model (Cui et al., 2021;Sinsabaugh & Follstad Shah, 2012). We determined the relevant soil parameters required in this model, i.e. dissolved nutrients (dissolved organic carbon where S C:N and S C:P is a scaler that represents the extent to which the allocation of extracellular enzyme activities (EEA) offsets the disparity between the elemental composition of available resources and the composition of microbial biomass (Sinsabaugh et al., 2016). ...
... Labile substrate availability for C, N and P was measured as the quantity of DOC, DON and Olsen P. Moreover, K C:N and K C:P , which are half-saturation constants for CUE based on C, N, and P availability, were assumed to be 0.5 (Sinsabaugh & Follstad Shah, 2012). The value of CUE max was assigned as 0.6, which is the upper limit for microbial growth efficiency based on thermodynamic constraints (Sinsabaugh et al., 2016). ...
Article
Full-text available
Enhanced rock weathering (ERW) has been proposed as a measure to enhance the carbon (C)‐sequestration potential and fertility of soils. The effects of this practice on the soil phosphorus (P) pools and the general mechanisms affecting microbial P cycling, as well as plant P uptake are not well understood. Here, the impact of ERW on soil P availability and microbial P cycling functional groups and root P‐acquisition traits were explored through a 2‐year wollastonite field addition experiment in a tropical rubber plantation. The results show that ERW significantly increased soil microbial carbon‐use efficiency and total P concentrations and indirectly increased soil P availability by enhancing organic P mobilization and mineralization of rhizosheath carboxylates and phosphatase, respectively. Also, ERW stimulated the activities of P‐solubilizing ( gcd , ppa and ppx ) and mineralizing enzymes ( phoADN and phnAPHLFXIM ), thus contributing to the inorganic P solubilization and organic P mineralization. Accompanying the increase in soil P availability, the P‐acquisition strategy of the rubber fine roots changed from do‐it‐yourself acquisition by roots to dependence on mycorrhizal collaboration and the release of root exudates. In addition, the direct effects of ERW on root P‐acquisition traits (such as root diameter, specific root length, and mycorrhizal colonization rate) may also be related to changes in the pattern of belowground carbon investments in plants. Our study provides a new insight that ERW increases carbon‐sequestration potential and P availability in tropical forests and profoundly affects belowground plant resource‐use strategies.
... 토양 효소는 유기물 분해와 질소와 인과 같은 양분의 무기화 등 토양 물질순환에 매우 중요하다 (Sinsabaugh and Shah, 2012). 자연상태의 토양과 달리 인위적으로 관리되는 농경지 토양에서 효소 활성도는 영농관리 요인에 영향을 받는다 (Jian et al., 2016;Uwituze et al., 2022). ...
... Because there was no change in either MBC or MBN, there was also consequently no change after fertilization in microbial stoichiometry, which can sometimes be used to suggest microbial limitation by specific resources (Sinsabaugh and Shah, 2012). Classic N-saturation models suggest that in an N-limited system, N added to the system will move from vegetation, to litter, to soil, and, given that enough N is added to the system, to N losses typically through leaching losses but potentially through denitrification and ammonia volatilization (Aber et al., 1998). ...
Article
Full-text available
Although the negative consequences of increased nitrogen (N) supply for plant communities and soil chemistry are well known, most studies have focused on mesic grasslands, and the fate of added N in arid and semi-arid ecosystems remains unclear. To study the impacts of long-term increased N deposition on ecosystem N pools, we sampled a 26-year-long fertilization (10 g N m−2 yr−1) experiment in the northern Chihuahuan Desert at the Sevilleta National Wildlife Refuge (SNWR) in New Mexico. To determine the fate of the added N, we measured multiple soil, microbial, and plant N pools in shallow soils at three time points across the 2020 growing season. We found small but significant increases with fertilization in soil-available NO3--N and NH4+-N, yet the soil microbial and plant communities do not appear to be taking advantage of the increased N availability, with no changes in biomass or N content in either community. However, there were increases in total soil N with fertilization, suggesting increases in microbial or plant N earlier in the experiment. Ultimately, the majority of the N added in this multi-decadal experiment was not found in the shallow soil or the microbial or plant community and is likely to have been lost from the ecosystem entirely.
Article
Microplastics (MPs) are emerging pollutants of terrestrial ecosystems. The impacts of MP particle size on terrestrial systems remain unclear. The current study aimed to investigate the effects of six particle sizes (i.e., 4500, 1500, 500, 50, 5, and 0.5 μm) of polyethylene (PE) and polyvinyl chloride (PVC) on soil respiration, enzyme activity, bacteria, fungi, protists, and seed germination. MPs significantly promoted soil respiration, and the stimulating effects of PE were the strongest for medium and small-sized (0.5–1500 μm) particles, while those of PVC were the strongest for small particle sizes (0.5–50 μm). Large-sized (4500 μm) PE and all sizes of PVC significantly improved soil urease activity, while medium-sized (1500 μm) PVC significantly improved soil invertase activity. MPs altered the soil microbial community diversity, and the effects were especially pronounced for medium and small-sized (0.5–1500 μm) particles of PE and PVC on bacteria and fungi and small-sized (0.5 μm) particles of PE on protists. The impacts of MPs on bacteria and fungi were greater than on protists. The seed germination rate of Brassica chinensis decreased gradually with the decrease in PE MPs particle size. Therefore, to reduce the impact of MPs on soil ecosystems, effective measures should be taken to avoid the transformation of MPs into smaller particles in soil environmental management.
Article
The effects of several phenolic monomers on the growth and laccase production by the edible mushroom Pleurotus sajor-caju have been compared. Of the phenols tested, 4-hydroxybenzaldehyde and vanillin were most inhibitory to fungal growth, and cultures supplemented with these two compounds also exhibited the largest increases in laccase specific activity compared with unsupplemented controls. Up to five laccase isoforms were detected in culture fluids when P. sajor-caju was grown in submerged culture in the presence of several phenolic monomers. A major laccase component (laccase IV) present in P. sajor-caju culture supernatants was purified 152-fold with an overall recovery of 4.9%. Laccase IV was shown to be homogeneous by SDS-PAGE and isoelectric focusing, and had a molecular weight of 55 kD and an isoelectric point of 3.6. It displayed a maximum reaction velocity at pH 2.1 and 45 C, and the activation energy of the enzyme reaction between 25 and 45 C as determined by an Arrhenius plot was 12.4 kJ mol⁻¹. The Km of laccase IV, as determined using ABTS as the substrate, was 0.092 mM. Laccase IV was glycosylated with mannose as the major sugar component together with galactose, glucosamine and fucose. Comparison of up to thirteen N-terminal amino acids of laccase IV with those of other fungal laccases revealed the highest similarity (85%) with a laccase from Pleurotus ostreatus. Considerable similarity (46–50%) was also observed with laccases from Coriolus versicolor, Pycnoporus cinnabarinus, Phlebia radiata, Coriolus hirsutus and Ceriporiopsis subvermispora but not Agaricus bisporus.
Book
Soil enzymes are one of the vital key mediators involved in nutrient recycling and the decomposition of organic matter and thereby in maintaining soil quality and fertility. This Soil Biology volume covers the various facets of soil enzymes, such as their functions, biochemical and microbiological properties and the factors affecting their activities. Enzymes in the rhizosphere, in forest soils, and in volcanic ash-derived soils are described. Soil enzymes covered include phosphohydrolases, lignocellulose-degrading enzymes, phenol oxidases, fungal oxidoreductases, keratinases, pectinases, xylanases, lipases and pectinases. Several chapters treat the soil enzymatic activities in the bioremediation of soils contaminated with pesticides and pollutants such as oil, chlorinated compounds, synthetic dyes and aromatic hydrocarbons. The role of soil enzymes as bioindicators is a further important topic addressed.
Book
Soil enzymes play a fundamental role in many soil processes such as the mineralization of organic matter, the synthesis of humic substances, the degradation of xenobiotics or the mechanisms involved in the biocontrol of plant pathogens. Their direct link with soil microorganisms gives them a key role as biomonitors of the evolution of soil quality or in the monitoring of the application of organic amendments to degraded soils. As a consequence of the importance of soil enzymes on soil processes, there is an increasing interest in their study, as well as in the application of molecular techniques as diagnostic tools.
Chapter
Organic matter degradation in aquatic environments has been studied for a long time by sanitary engineers and geologists concerned with predicting the rate of this process, either in polluted rivers, in sewage treatment plants or in sediment. Models have been established for this purpose, and most of them derive from the early Streeter and Phelps (1925) model where the rate of organic matter degradation is simply assumed to be proportional to the organic load. In order to take into account the differing susceptibilities to bacterial attack of the various classes of compounds making up the overall organic matter, Jorgensen (1978), Berner (1980) and Westrich and Berner (1984) suggested the use of “multi G’s-first order kinetics,” considering a number of organic matter fractions, each with its own first-order degradation constant.
Article
For many years, the synthesis, secretion, and activity of extracellular enzymes of different organisms have been intensively studied by workers in many different disciplines, such as applied microbiology, biotechnology, biochemistry, and medicine (Pollock, 1962; Priest, 1977; 1984; Kreutzberg et al., 1986; Chaloupka and Krumphanzl, 1987). In aquatic sciences, the first reports on extracellular enzymes were written in middle of the 1960s (Overbeck and Babenzien, 1964; Reichardt et al., 1967; Kim and ZoBell, 1974). However, in the last decade, interest has increased in the role of microbial extracellular enzymes in natural aquatic environments.