pdf (1.2mb)

16
Ecological Monographs, 85(3), 2015, pp. 457–472 Ó 2015 by the Ecological Society of America Revisiting the hypothesis that fungal-to-bacterial dominance characterizes turnover of soil organic matter and nutrients JOHANNES ROUSK 1,3 AND SERITA D. FREY 2 1 Section of Microbial Ecology, Department of Biology, Lund University, Lund, Sweden 2 Department of Natural Resources and the Environment, University of New Hampshire, Durham, New Hampshire 03824 USA Abstract. Resolving fungal and bacterial groups within the microbial decomposer community is thought to capture disparate life strategies for soil microbial decomposers, associating bacteria with an r-selected strategy for carbon (C) and nutrient use, and fungi with a K-selected strategy. Additionally, food-web model-based work has established a widely held belief that the bacterial decomposer pathway in soil supports high turnover rates of easily available substrates, while the slower fungal-dominated decomposition pathway supports the decomposition of more complex organic material, thus characterizing the biogeochemistry of the ecosystem. We used a field experiment, the Detritus Input and Removal Treatments, or DIRT, experiment (Harvard Forest Long-Term Ecological Research Site, USA) where litter and root inputs (control, no litter, double litter, or no tree roots) have been experimentally manipulated during 23 years, generating differences in soil C quality. We hypothesized (1) that d 13 C enrichment would decrease with higher soil C quality and that a higher C quality would favor bacterial decomposers, (2) that the C mineralized in fungal-dominated treatments would be of lower quality and also depleted in d 13 C relative to bacterial-dominated high-quality soil C treatments, and (3) that higher C mineralization along with higher gross N mineralization rates would occur in bacterial- dominated treatments compared with more fungal-dominated treatments. The DIRT treatments resulted in a gradient of soil C quality, as shown by up to 4.5-fold differences between the respiration per soil C between treatments. High-quality C benefited fungal dominance, in direct contrast with our hypothesis. Further, there was no difference between the d 13 CO 2 produced by a fungal compared with a bacterial-dominated decomposer community. There were differences in C and N mineralization between DIRT treatments, but these were not related to the relative dominance of fungal and bacterial decomposers. Thus we find no support for the hypothesized differences between detrital food webs dominated by bacteria compared to those dominated by fungi. Rather, an association between high-quality soil C and fungi emerges from our results. Consequently there is a need to revise our basic understanding for microbial communities and the processes they regulate in soil. Key words: aboveground/belowground; biogeochemistry; carbon sequestration; decomposer ecology; food web; forest ecology; fungal and bacterial dominance; Harvard Forest Long-Term Ecological Research (LTER); soil C cycle. INTRODUCTION Soils contain vast and dynamic pools of carbon (C) that are fundamental for maintaining the balance of atmospheric CO 2 concentrations (Denman et al. 2007). Transformations of soil organic matter (SOM) mostly rely on the activity of soil decomposer microorganisms that control the release of greenhouse gases including CO 2 (Waksman and Gerretsen 1931, Schlesinger and Andrews 2000), the storage of SOM and subsequent maintenance of soil physical structure and fertility (Le Guillou et al. 2012), and nutrient regeneration essential for plant nutrition (Waksman and Starkey 1932). Despite their integral biogeochemical role, the explicit incorporation of microbial dynamics into C cycle models is limited by insufficient knowledge of if and how microbial community structure translates into functional relevance (Fierer et al. 2007, Keiser et al. 2014). Most C models treat the microbial community implicitly (e.g., Jenkinson et al. 1999), assuming that microbial activity is solely determined by abiotic factors, such as temperature, moisture, and pH, while commu- nity identity is not explicitly considered. This approach assumes that microbial communities, or their supported food webs, cannot be distinguished with regard to their functional impact in the soil system. Recent models have started to explicitly consider the qualitative characteris- tics of decomposer microbial communities (phylogenic or functional composition; Fontaine et al. 2003, Fontaine and Barot 2005, Allison 2012) finding im- proved model performances (Wieder et al. 2013). While merely resolving the microbial community into fungal Manuscript received 20 September 2014; revised 22 January 2015; accepted 28 January 2015. Corresponding Editor: J. B. Yavitt. 3 E-mail: [email protected] 457

Upload: ngotuyen

Post on 15-Dec-2016

237 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: PDF (1.2MB)

Ecological Monographs, 85(3), 2015, pp. 457–472� 2015 by the Ecological Society of America

Revisiting the hypothesis that fungal-to-bacterial dominancecharacterizes turnover of soil organic matter and nutrients

JOHANNES ROUSK1,3

AND SERITA D. FREY2

1Section of Microbial Ecology, Department of Biology, Lund University, Lund, Sweden2Department of Natural Resources and the Environment, University of New Hampshire, Durham, New Hampshire 03824 USA

Abstract. Resolving fungal and bacterial groups within the microbial decomposercommunity is thought to capture disparate life strategies for soil microbial decomposers,associating bacteria with an r-selected strategy for carbon (C) and nutrient use, and fungiwith a K-selected strategy. Additionally, food-web model-based work has established a widelyheld belief that the bacterial decomposer pathway in soil supports high turnover rates ofeasily available substrates, while the slower fungal-dominated decomposition pathwaysupports the decomposition of more complex organic material, thus characterizing thebiogeochemistry of the ecosystem. We used a field experiment, the Detritus Input andRemoval Treatments, or DIRT, experiment (Harvard Forest Long-Term EcologicalResearch Site, USA) where litter and root inputs (control, no litter, double litter, or notree roots) have been experimentally manipulated during 23 years, generating differences insoil C quality. We hypothesized (1) that d13C enrichment would decrease with higher soil Cquality and that a higher C quality would favor bacterial decomposers, (2) that the Cmineralized in fungal-dominated treatments would be of lower quality and also depleted ind13C relative to bacterial-dominated high-quality soil C treatments, and (3) that higher Cmineralization along with higher gross N mineralization rates would occur in bacterial-dominated treatments compared with more fungal-dominated treatments. The DIRTtreatments resulted in a gradient of soil C quality, as shown by up to 4.5-fold differencesbetween the respiration per soil C between treatments. High-quality C benefited fungaldominance, in direct contrast with our hypothesis. Further, there was no difference betweenthe d13CO2 produced by a fungal compared with a bacterial-dominated decomposercommunity. There were differences in C and N mineralization between DIRT treatments, butthese were not related to the relative dominance of fungal and bacterial decomposers. Thuswe find no support for the hypothesized differences between detrital food webs dominated bybacteria compared to those dominated by fungi. Rather, an association between high-qualitysoil C and fungi emerges from our results. Consequently there is a need to revise our basicunderstanding for microbial communities and the processes they regulate in soil.

Key words: aboveground/belowground; biogeochemistry; carbon sequestration; decomposer ecology;food web; forest ecology; fungal and bacterial dominance; Harvard Forest Long-Term Ecological Research(LTER); soil C cycle.

INTRODUCTION

Soils contain vast and dynamic pools of carbon (C)

that are fundamental for maintaining the balance of

atmospheric CO2 concentrations (Denman et al. 2007).

Transformations of soil organic matter (SOM) mostly

rely on the activity of soil decomposer microorganisms

that control the release of greenhouse gases including

CO2 (Waksman and Gerretsen 1931, Schlesinger and

Andrews 2000), the storage of SOM and subsequent

maintenance of soil physical structure and fertility (Le

Guillou et al. 2012), and nutrient regeneration essential

for plant nutrition (Waksman and Starkey 1932).

Despite their integral biogeochemical role, the explicit

incorporation of microbial dynamics into C cycle

models is limited by insufficient knowledge of if and

how microbial community structure translates into

functional relevance (Fierer et al. 2007, Keiser et al.

2014). Most C models treat the microbial community

implicitly (e.g., Jenkinson et al. 1999), assuming that

microbial activity is solely determined by abiotic factors,

such as temperature, moisture, and pH, while commu-

nity identity is not explicitly considered. This approach

assumes that microbial communities, or their supported

food webs, cannot be distinguished with regard to their

functional impact in the soil system. Recent models have

started to explicitly consider the qualitative characteris-

tics of decomposer microbial communities (phylogenic

or functional composition; Fontaine et al. 2003,

Fontaine and Barot 2005, Allison 2012) finding im-

proved model performances (Wieder et al. 2013). While

merely resolving the microbial community into fungal

Manuscript received 20 September 2014; revised 22 January2015; accepted 28 January 2015. Corresponding Editor: J. B.Yavitt.

3 E-mail: [email protected]

457

Page 2: PDF (1.2MB)

and bacterial components is a coarse division, there are

some emerging general patterns to note. In fact,

resolving fungal and bacterial decomposers has been

thought to capture disparate life strategies for soil

microbial decomposers, associating bacteria with an r-

selected strategy for C and nutrient use, and fungi with a

K-selected strategy (Fontaine et al. 2003, Fontaine and

Barot 2005, Blagodatskaya and Kuzyakov 2008, Fierer

et al. 2009). Moreover, food-web models predict very

different outcomes with regard to C sequestration,

nutrient turnover, and ecosystem stability depending

on microbial community composition, especially in

terms of the relative dominance of fungi and bacteria

(Moore et al. 2004, Six et al. 2006, Neutel et al. 2007,

Bezemer et al. 2010). In general, a widely held

assumption is that bacterial decomposer pathways in

soil support high turnover rates of easily available

substrates, while slower fungal-dominated decomposi-

tion pathways support the decomposition of more

complex organic material (Wardle et al. 2004, de Graaff

et al. 2010, de Vries et al. 2012).

The C-limited soil microbial community is dependent

on and regulated by the availability and quality of

organic matter (Demoling et al. 2007, Egli 2010, Hobbie

and Hobbie 2012, Inselbacher and Nasholm 2012). The

availability and quality of soil organic matter, in turn, is

regulated by plant inputs. The plant influence is

provided both from aboveground sources such as litter

input (Hieber and Gessner 2002, Romani et al. 2006,

Sayer et al. 2011, Rousk et al. 2013), and from

belowground sources, such as root exudates and

deposition (Farrar et al. 2003). It has been demonstrated

that nutrient content (Henriksen and Breland 1999) and

the C quality of litter (Rousk and Baath 2007)

determines which components of the microbial commu-

nity that respond. Similarly, the type and concentration

of root exudate additions to soil determine the

responsive fractions within the microbial community

(de Graaff et al. 2010, Eilers et al. 2010, Reischke et al.

2014, Rousk et al. 2015). Plant litter of high C quality

appears to favor bacterial more than fungal growth,

while lower-quality plant litter additions shows the

opposite relationship (Rousk and Baath 2007, Strick-

land et al. 2009). However, most experimental assess-

ments that have resolved microbial responses to date

have involved the influence of single additions of litter or

model root exudate compounds to soil, and most

assessments have been done in laboratory systems.

Consequently, the effects of plant inputs on microbial

responses have not yet been sufficiently validated in

intact ecosystems with semi-continuous supplies of litter

and root deposition. Moreover, how plant-input-in-

duced microbial responses translate into differences in

ecosystem function (e.g., C and N mineralization),

remains to be established.

Stable isotope probing (SIP) presents a means by

which the microbial use of C derived from plant material

of varying quality can be assessed. Biosynthetic processes

discriminate predictably against different C isotopes, and

the plant C produced by C3 plants, for example, have a

well-constrained d13C signal (Hayes 2001). After incor-

poration into microbial biomass, the C is slightly

enriched in 13C compared to the initial substrate. Thus,

after a cycle of microbial use, the C released back to the

soil environment as a result of microbial turnover has

become enriched in d13C (Abraham et al. 1998, Cifuentes

and Salata 2001). Since microbial derived C is of lower

quality compared to plant-derived C (Agren et al. 1996,

Schiff et al. 1997, Cotrufo et al. 2013), the gradual

enrichment in d13C signal in SOC that develops with each

microbial use cycle coincides with a gradual reduction in

SOC quality (Churchland et al. 2013). Although it is

theoretically possible that this relationship can be offset

by changes in the composition of compounds that make

up the SOC (e.g., an enrichment of lignin) due to

compound specific differences in d13C signals, the pattern

has been found to be robust within and between soils and

to overshadow compound specific differences (Sollins et

al. 2009, Gunina and Kuzyakov 2014). Consequently,

using differences in natural abundances of C isotopes, an

effective index presents itself from which we can infer the

quality of C left in the soil, as well as that used by

microorganisms for both catabolism (release of CO2) and

anabolism (incorporation into microbial biomarkers).

To test how microbial communities were affected by

SOC quality, an operational definition of SOC quality as

perceived by the microbial community is required. We

used two indices to achieve this. First, a direct measure

of the microbial ability to use the SOC as measured

respiration per SOC. While this measure only includes

one aspect of the microbial C use (catabolism; not

anabolism), it has been found to be a robust proxy for

microbial assimilability of C for both growth and

respiration under stable-state laboratory conditions

(Agren et al. 1996, Schiff et al. 1997, Cotrufo et al.

2013). Second, the d13C enrichment can be used to track

the quality of C in SOC, microorganisms, or respired

CO2.

We used a long-term field experiment where plant

inputs have been manipulated in a temperate forest

ecosystem over 23 years. This Detritus Input and

Removal Treatments, or DIRT, experiment has exper-

imentally adjusted litter input (no litter, double litter, or

control) and tree root input (no tree roots) in a

temperate, mixed-hardwood stand at the Harvard

Forest Long-Term Ecological Research (LTER) site in

Massachusetts, USA. This field experiment was com-

bined with (1) a natural abundance carbon isotope

(d13C) study to quantify the plant input manipulation

effects on the quality of the SOC, and its contribution to

respired CO2 and incorporation into the microbial

community as determined using microbial biomarkers

(phospholipid fatty acids, PLFAs); (2) growth rate and

biomass based assessments of fungal and bacterial

components resulting from the long-term plant-input

manipulations; and (3) C (respiration) and N mineral-

JOHANNES ROUSK AND SERITA D. FREY458 Ecological MonographsVol. 85, No. 3

Page 3: PDF (1.2MB)

ization (15N-pool dilutions) determinations to assess

microbial function. We addressed some long-standing

assumptions in soil microbial ecology by evaluating

three main hypotheses. (H1) SOM formed by 23 years of

increased plant litter inputs (e.g., double litter) should be

of higher quality and thus favor bacterial decomposers

compared to SOM resulting from 23-years of reduced

litter or root inputs (e.g., no litter, no tree roots), which

will be enriched in lower-quality C and thus favor fungal

decomposers. (H2) The C mineralized (d13CO2) in the

low-plant-input treatments (and presumably dominated

by low-quality C and fungi) will be enriched in d13Ccompared to the C mineralized from the high-C-input

treatments. (H3) Bacterial-dominated decomposition in

high SOC quality treatments will have a higher

mineralization rate of both C and N and a higher

potential for nutrient leakage (gross N mineralization

rates), and a lower potential to retain mineral N (N

assimilation rates) than the relatively more fungal-

dominated treatments.

METHODS

Site, field experiment, and sampling

The study site is located in Harvard Forest in north-

central Massachusetts, and has previously been de-

scribed in greater detail (Bowden et al. 1993). It is about

a 1-ha, century-old, mixed-hardwood stand dominated

by northern red oak (Quercus borealis Michx. F.; 43%basal area), red maple (Acer rubrum L.; 19% basal area),

and paper birch (Betula papurifera Marsh.; 15% basal

area), with an herbaceous ground cover dominated by

Canada mayflower (Maianthemum canadense Desf.) and

a variety of ferns. The soil is derived from glacial till

deposited (about 3 m deep) over granite, gneiss, and

schist bedrock, and is moderately well drained stony-

loam classified as an Inceptisol (USDA taxonomy). The

mean annual temperature is 78C and mean annual

precipitation about 1100 mm.

The DIRT experimental plots were established at the

site in 1990, as described in Bowden et al. (1993). The

plots are 3 3 3 m, free of trees and saplings, and with

aboveground and belowground plant C inputs manip-

ulated in a number of ways. The treatments include

control (Co, normal annual aboveground litter inputs),

double litter (2X, twice the litter inputs of the control

plots), no litter (0X, annual aboveground litter inputs

excluded), no tree roots (T, plots trenched and root

regrowth into plots prevented), no input (T0X, plots

trenched and annual aboveground litter excluded), and

no OA (OA, soil O, and A horizons removed and

replaced with B horizon material from nearby excava-

tions). The treatment codes are provided in Table 1. The

plots (n¼ 3) were set up in a stratified random approach

at upslope, midslope, and downslope positions within

the site (see Bowden et al. [1993] for more details), with

two randomly assigned control plots present in each of

the three blocks.

The litter interception treatments (0X, T0X) were

administered by placing plastic mesh shade cloth on the

plots from early September to late October (a period

during which 95% of all litter fall occurs). After this

period, the intercepted litter was returned to the plots

TABLE 1. Treatment codes and soil characteristics.

Treatment, horizon Soil pHSoil moisture

(% fm)SOM(%)

SOC(mg C/g soil)

TN(mg N g soil)

C:N(by mass)

NH4þ

(lg NH4þ-N/g soil)

Double litter, 2X

O 3.7 (0.1) 55.8 (4.0) 46.3 (4.8) 275 (20) 12.8 (1.4) 16.3 (2.7) 3.2 (1.0)M 4.2 (0.1) 32.2 (0.9) 13.0 (1.3) 67 (8) 3.7 (0.5) 15.6 (1.1) 3.8 (1.2)

Control, Co

O 3.8 (0.1) 52.3 (4.9) 42.9 (4.2) 248 (23) 11.7 (2.1) 16.3 (1.7) 5.6 (1.9)M 4.3 (0.3) 33.8 (0.7) 12.6 (0.7) 68 (4) 3.9 (0.4) 14.6 (2.0) 4.0 (1.7)

No litter, 0X

O 4.5 (0.2) 35.7 (2.6) 13.4 (4.6) 100 (13) 5.7 (0.6) 10.4 (5.6) 2.1 (0.3)M 4.4 (0.1) 31.0 (1.5) 10.3 (1.1) 47 (1) 3.0 (0.4) 15.9 (4.6) 3.2 (0.7)

No OA, OA

O 4.2 (0.1) 38.0 (1.5) 22.4 (1.0) 127 (9) 6.6 (0.5) 15.3 (0.7) 1.5 (0.2)M 4.6 (0.1) 19.6 (1.7) 4.6 (0.5) 19 (1) 1.6 (0.2) 12.7 (3.2) 2.4 (0.3)

No roots, T

O 4.0 (0.1) 51.0 (0.5) 33.0 (2.9) 192 (25) 9.5 (1.4) 15.9 (3.3) 1.8 (0.4)M 4.4 (0.1) 35.4 (0.8) 12.2 (0.9) 60 (10) 3.6 (1.1) 15.9 (2.7) 5.7 (2.2)

No input, T0X

O 4.4 (0.2) 39.3 (1.2) 18.4 (2.4) 95 (23) 5.2 (1.6) 16.6 (3.3) 4.8 (2.0)M 4.5 (0.0) 31.6 (1.3) 9.6 (0.4) 42 (2) 2.8 (0.4) 15.9 (3.6) 2.9 (0.8)

Notes: Values are means (with SE in parentheses) of six (Co) or three (other treatments) field replicates in the O and M horizons.Soil pH is the KCl (0.1 mol/L) extract (1:5, mass : volume). Soil moisture was determined using gravimentric loss at 1058C for 24 h,as is given as mass percentage of fresh mass (fm). Soil organic matter (SOM) content was determined using loss on ignition (6008Cfor 16 h). Soil organic carbon (SOC) and total nitrogen (TN) were determined using an elemental analyzer. The NH4

þ

concentration was determined using diffusion traps in a KCl extract (1 mol/L). See Methods for further details.

August 2015 459SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 4: PDF (1.2MB)

with normal rates of litter input (Co, T, OA), or

redistributed from the litter interception treatments (0X)

to the double litter treatment (2X). In the treatments

without root input (T and T0X), 70–100 cm trenches,

extending about 20 cm below the lower edge of the

rooting zone, were dug 0.5 m outside of the plot

boundaries. Corrugated fiberglass sheets (3 mm thick)

were installed to prevent roots from reestablishing, after

which the trenches were refilled. After noting signs of

root reestablishment, trenching treatments were re-

administered in 1999. This time, new trenches were

dug down to 1 m depth. The original plastic barrier was

removed and replaced with a seven-ply laminate

combining four layers of high density polyethylene and

three high-strength cord grids (Griffolyn; Reef Indus-

tries, Texas, USA), folded over twice. The new dividers

were extended 20–40 cm aboveground, and 30 cm below

the 1 m deep trenches. The OA treatments were

established by removing the O and A horizons and

replacing them with B horizon soils excavated from

nearby. The treatments have been maintained continu-

ously, and the no input (T0X) plots are kept free of

seedlings and herbaceous vegetation and falling wood

litter (more than twice each growing season).

The DIRT plots were sampled in late October 2013.

From each plot, four cores (8 cm wide 3 10 cm deep)

were separated into organic (O) and mineral (M)

horizons and then combined into composite samples

maintaining the two horizons and plot replication. The

treatment where mineral (B horizon) soil had replaced

the O and Oa horizons (treatment OA) at the field

experiment’s initiation had a poorly developed O

horizon (;1–2 cm depth). The samples were sieved (4

mm O horizon; 2 mm M horizon), pebbles and roots

were removed, and then mixed and homogenized into

press-seal polypropylene bags that were stored on ice in

styrofoam insulated boxes during shipping. Microbial

analyses were performed within five days of sampling.

Soil characterization

Water content was determined gravimetrically (1058C

for 24 h), organic matter content by loss on ignition

(6008C for 16 h), and soil pH in a 1:5 (mass : volume)

extraction with 0.1 mol/L KCl. Total soil organic C and

N concentrations were determined by dry combustion

using a Flash 2000 elemental analyzer (Thermo Fisher

Scientific, Bremen, Germany). Soil NH4 concentrations

were determined as part of the gross N mineralization

determination.

Bacterial and fungal growth, biomass and PLFA

composition, and total respiration

Bacterial growth was measured using the leucine

(Leu) incorporation method (Kirchman et al. 1985)

adapted for soil (Baath et al. 2001) with modifications

(Rousk et al. 2009), which estimates the rate of protein

synthesis as a measure of bacterial growth. We used a

mixture of radio-labeled Leu, [3H]Leu (37 MBq/mL,

5.74 TBq/mmol; Perkin Elmer, Waltham, Massachu-

setts, USA) and unlabeled Leu, resulting in 275 nmol/L

Leu in the bacterial suspension, and a 2 h incubation at

228C without light. Fungal growth and biomass

(ergosterol concentration) were determined using the

acetate-in-ergosterol incorporation method (Newell and

Fallon 1991) adapted for soil (Baath 2001), with

modifications (Rousk and Baath 2007, Rousk et al.

2009), which estimates the rate of ergosterol synthesis as

a measure of fungal growth. We used 1-[14C]acetic acid

(sodium salt, 2.04 GBq/mmol, 7.4 MBq/mL; Perkin

Elmer) and unlabeled acetate resulting in a final acetate

concentration of 220 lmol/L, and an incubation time of

4 h at 228C without light.

Respiration was measured by transferring 1 (O

horizon) or 2 (M horizon) g soil into a 20-mL glass

vial and purging the headspace with pressurized air. The

vial was closed with a crimp lid and incubated for 24 h at

228C without light. Afterward, the headspace CO2

concentration was analyzed using a gas chromatograph

(GC) equipped with a methanizer and a flame ionization

detector (FID; YL6500 GC, YL instruments, Gyeonggi-

do, Korea). Sub-samples of the headspace (1.5 mL) were

also transferred to He-purged gas-tight borosilicate

tubes (12 mL; Labco, Lampeter, UK) for subsequent

isotope ratio measurements of the produced CO2.

Microbial biomass was measured using the substrate

induced respiration (SIR) method (Anderson and

Domsch 1978). Glucose : talcum (4:1; 5 mg/g) was added

and mixed into the soil samples following the respiration

analyses. After 20 minutes, the atmosphere was purged

of CO2 with pressurized air, after which the vials were

again closed with crimp caps, and incubated for 2–3 h at

228C. The CO2 evolved was then determined as

described above. SIR-respiration was converted to

biomass using the relationship 1 mg CO2/h at 228C

corresponds to 20 mg biomass C (recalculated from

Anderson and Domsch 1978), and assuming a 45% C

content in microbial biomass.

The microbial phospholipid fatty acid (PLFA) com-

position was determined using 1 (O horizon) or 2 g (M

horizon) soil according to Frostegard et al. (1993) with

modifications described by Nilsson et al. (2007). An

internal standard (methyl nonadecanoate fatty acid,

19:0) was added before the methylation step. The

derived fatty acid method esters were quantified on a

GC-FID as previously described (Frostegard et al.

1993), and then analyzed for d13C content. The PLFAs

chosen to indicate bacterial biomass were i15:0, a15:0,

i16:0, 16:1x9, 16:1x7c, 10Me16:0, cy17:0, i17:0, a17:0,

18:1x7, and cy19:0, while PLFA 18:2x6,9 was used to

indicate fungal biomass (Frostegard and Baath 1996).

Taxonomic groups were designated based on a sub-

selection from Ruess and Chamberlain (2010).

Gross N transformation rates

The N mineralization and consumption rates were

determined using the 15N pool-dilution method. Each

JOHANNES ROUSK AND SERITA D. FREY460 Ecological MonographsVol. 85, No. 3

Page 5: PDF (1.2MB)

homogenized soil sample was divided into to subsamples

(each 10 g fresh mass) in 200-mL, gas-tight, polypropyl-

ene cups, to which 350 lL NH4Cl (45 lg N/mL) enrichedto 98% 15N (Sigma Aldrich, St. Louis, Missouri, USA)

were administered evenly using a pipette, followed by

vigorous mixing with a stainless steel spatula to ensureeven distribution, after which they were lidded. One set

of the subsamples was extracted using a 1 mol/L KCl

solution (1:5, mass : volume) 1 h after 15N addition, andthe second set was treated identically after about 24 h

incubation at 228C without light. The amounts of14NH4

þ-N and 15NH4þ-N were determined by isotope

ratio mass spectroscopy (IRMS) after diffusion to glass

fiber traps, according to standard procedures (IAEA

2001). The 14/15N-content of the glass fiber traps weredried at room temperature using a desiccator, transferred

to tin cups, and measured with a Flash 2000 elemental

analyzer coupled to a Delta V Plus via the ConFlow IVinterface (Thermo Fisher Scientific, Bremen, Germany)

at the Stable Isotope Facility at the Department of

Biology, Lund University, Sweden. The 15N addition tothese soils increased in the inorganic N concentration by

approximately 1–2 lg N/g, corresponding to rather

limited changes in mineral N concentrations, at ca. 50%increases in NH4

þ concentrations above initial levels. The

gross N transformation rates were estimated as described

in Bengtson et al. (2005).

d13C measurements

All results on 13C/12C ratios are reported as per mil(%) difference using the Pee Dee Belemnite (PDB)

standard in the ] notation:

d13Cð%Þ ¼ ½ðRsample=Rstd � 1Þ� 3 1000

where Rsample is the13C/12C ratio of the sample and Rstd

is the 13C/12C ratio in the PDB standard. All IRMSmeasurements were performed using various appliances

coupled to a Delta V Plus via the ConFlow IV interface

(Thermo Fisher Scientific, Bremen, Germany) at theStable Isotope Facility at the Department of Biology,

Lund University, Sweden. For measurements of d13C in

soil a Flash 2000 elemental analyzer was used, for d13Cin respired CO2 a GasBench II was used, and for the

d13C in FAMEs derived from extracted PLFAs a Trace

GC Ultra gas chromatograph was used. The reported

d13C values for the PLFAs were corrected for the Catom added during the methylation step as in Williams

et al. (2006). The d13C in the respired CO2 was corrected

for the background signal of d13C in CO2 in the sampledheadspace vials using standard procedures.

Statistical analyses

Treatment effects on all measured variables were

tested for using two-way analysis of variance (ANOVA)

with the field treatment as one factor (Co, 2X, 0X, T,T0X, and OA) and soil horizon (O or M) as the other.

Treatment comparisons were done using Tukey’s pair-

wise comparisons at the a¼ 0.05 level. The relationship

between C quality (respiration per unit SOC) and d13Ccontent of SOC and respired CO2 and N transformationrates were tested using linear type II major axis

regressions. The PLFA composition (molecular percent-age of the 30 most abundant PLFAs) was analyzed with

a principal component analysis (PCA) after standardiz-ing to unit variance. To directly test microbial responses

to a C-quality effect, as measured with respiration perunit SOC, type II major axis regressions were used. Toconservatively estimate independence between field

treatments, these regressions were done on the meansof each treatment, rather than individual field replicates

(i.e., assuming n¼ 12 rather than n¼ 42), to avoid falsepositives.

RESULTS

Soil characteristics

The SOM content was affected by both treatment(F5,30¼ 14.0, P , 0.0001) and horizon (F1,30¼ 116, P ,

0.0001) with an interaction between the factors (F5,30 ¼7.9, P , 0.0001). There was a higher concentration of

SOM in the organic than in the mineral horizon, whilethe treatments separated clearly between those thatreceived litter and those did not (Table 1). The SOC and

N concentrations behaved very similarly to the SOMconcentrations, leading to a stable C:N ratio, indistin-

guishable between horizons and treatments, at about 15(Table 1). Soil moisture was also affected by both

treatment (F5,30¼ 8.45, P , 0.0001) and horizon (F1,30¼69, P , 0.0001) with an interaction between the factors

(F5,30 ¼ 2.8, P ¼ 0.04; Table 1). The O horizon hadhigher moisture than the M horizon, and the treatments

with litter had higher moisture than those without litterinput (Table 1). However, the observed differences in

soil moisture were related to variation in the water-holding capacity, driven by the SOM concentration,

rather than differences in microbial availability to water.The NH4

þ concentration was not distinguishable be-

tween horizons or treatments, averaging about 3.4 lg N/g soil (Table 1). Soil pH was affected by treatment (F5,30

¼11.7, P , 0.001) and horizon (F1,30¼38.4, P , 0.0001)with an interaction between the two factors (F5,30¼ 4.0,

P ¼ 0.007). It was marginally higher in the M than Ohorizon, while the treatments without plant litter input(0X, T0X) had marginal higher pH than those with litter

input (Co, 2X; Table 1).

Fungal and bacterial growth

There were small but systematic effects of treatments

and horizon, with an interaction between the factors forfungal growth per unit SOC (Fig. 1a, b). Higher litter

input appeared to stimulate fungal growth; however, aclear effect was the high rate of fungal growth in the OA

treatment, with fungal growth rates being higher overallin the O compared with the M horizon. The bacterial

growth per unit SOC was affected by treatment but notby horizon (Fig. 2c, d). Like for fungal growth, high

rates of bacterial growth were observed in the OA

August 2015 461SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 6: PDF (1.2MB)

treatment, but the treatments with litter additions (2X,

Co, T) tended to be only marginally lower than those

without (0X, T0X). This yielded a ratio between fungal

and bacterial growth that was affected by treatment but

not horizon (Fig. 1e, f ). Again, we observed a trend

where treatments with litter inputs (2X, Co, OA, T) were

higher than those without litter input (0X, T0X)

suggesting a correlation between fungal dominance

and litter input.

Respiration and gross N transformation

The respiration rate per unit SOC was affected by

both treatment and horizon (Fig. 2a, b). Rates were

higher overall in the O compared to the M horizon and

were particularly high in the OA treatment. However,

FIG. 1. (a, b) The fungal growth rate as determined using the acetate incorporation into ergosterol method (treatment F5,30¼12.4, P , 0.0001; horizon F1,30¼ 17.5, P , 0.0001; treatment 3 horizon F5,30¼ 3.0, P¼ 0.02), (c, d) the bacterial growth rate asdetermined using the leucine incorporation into extracted bacteria method (treatment F5,30¼ 3.6, P¼ 0.011; horizon F1,30¼ 0.64,P ¼ 0.43; treatment 3 horizon F5,30 ¼ 2.7, P ¼ 0.041), and (e, f ) the logarithm of the ratio between fungal-to-bacterial growth(treatment F5,30¼3.3, P¼0.018; horizon F1,30¼0.25, P¼0.62; treatment3horizon F5,30¼1.2, P¼0.33). The O (a, c, and e) and M(b, d, and f ) horizon results are presented separately. Values represent the mean 6 SE of six (Co) or three (other treatments) fieldreplicates, and are presented per soil organic carbon (SOC). Statistics are from two-way ANOVAs. See Table 1 for treatmentdefinitions.

JOHANNES ROUSK AND SERITA D. FREY462 Ecological MonographsVol. 85, No. 3

Page 7: PDF (1.2MB)

there was also a tendency for higher respiration in

treatments with litter input (2X, Co, OA, T) compared

to those without litter input (0X, T0X); a trend observed

in the O but not in the M horizon. The NH4þ production

per unit SOC estimates gave a very similar, and yet

clearer picture (Fig. 2c, d). Rates were higher in the O

than M horizons, and treatments with litter input (2C,

Co, OA, and T) were all clearly higher than those

without litter input (0X and T0X) in the O horizon, with

no obvious treatment differences in the M horizon. No

significant treatment or horizon differences were ob-

served for NH4þ consumption per unit SOC estimates

(Fig. 2e, f ).

C and N mineralization relationships with SOC quality

The d13C enrichment of SOC was distinguishable

between horizons F1,30 ¼ 145, P , 0.0001) as well as

between treatments (F5,30 ¼ 6.4, P ¼ 0.0004) where an

interaction (F5,30 ¼ 2.6, P ¼ 0.046) revealed larger

treatment effects in the O horizon. The SOC pool was

FIG. 2. The microbial rates of (a and b) C (respiration; treatment F5,30¼ 10.7, P , 0.0001; horizon F1,30¼ 31.2, P , 0.0001;treatment3horizon F5,30¼ 0.99, P¼ 0.49) and (c–f ) N (NH4

þ production, panels c and d [treatment F5,30¼ 2.3, P¼ 0.067; horizonF1,30¼8.1, P¼0.0079; treatment3horizon F5,30¼0.59, P¼0.70], and consumption, panels e and f [treatment F5,30¼1.27, P¼0.30;horizon F1,30¼ 0.02, P¼ 0.86; treatment3horizon F5,30¼ 1.6, P¼ 0.18]) transformation. The O (panels a, c, and e) and M (panelsb, d, and f ) results are presented separately. Values represent the mean 6 SE of six (Co) or three (other treatments) field replicates,and are presented per soil organic carbon (SOC). Statistics are from two-way ANOVAs.

August 2015 463SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 8: PDF (1.2MB)

more enriched in 13C in the M than the O horizon, and

arranged in order from least to most depleted as 0X (a),

T0X (ab), Co (bc), T (bc), OA (bc), and 2X (c)

(treatments sharing the same lowercase letter are not

significantly different per Tukey’s test; Fig. 3a). This

variation in d13C enrichment of SOC correlated strongly

with C quality, as indicated by respiration per unit SOC

(Fig. 3a), where the most depleted SOC (OA and 2X

treatments in the O horizon with a d13C ,�28.0%) had

a fourfold higher quality than the most enriched SOC

(2X and T0X treatments in the M horizon with a d13C .

�26.5%; Fig. 3a). In contrast, the ]13C enrichment in the

respired CO2 revealed no systematic horizon or treat-

ment dependence, and thus C quality had no influence

on the CO2 d13C enrichment (Fig. 3b).

The NH4þ production showed a clear dependence on

the quality of SOC as indicated by both respiration per

SOC (Fig. 3c) and by the d13C enrichment in SOC (Fig.

3d). The slope between the rate of mineralization of C

per h and N per day (Fig. 3c) suggested that the C:N

ratio of the SOM mineralized was ;25, and thus higher

than the average C:N of the SOM (;15; Table 1). In

contrast, the NH4þ consumption rate did not show any

clear relationship with the quality of SOC. Instead,

strong positive relationships between NH4þ consump-

tion rates and both fungal (type II major axis regression:

P ¼ 0.02, R2 ¼ 0.43, n ¼ 12), and particularly with

bacterial growth (type II major axis regression: P ¼0.0006, R2 ¼ 0.71, n ¼ 12) could be substantiated.

Microbial biomass

The microbial biomass per unit SOC, as measured by

SIR, depended on treatment (F5,30¼ 10.1, P , 0.0001),

but only marginally on soil horizon (F1,30 ¼ 4.1, P ¼

FIG. 3. The dependence of carbon (C) quality (as indicated by respiration per SOC) on the d13C in (a) soil organic C (y¼�414� 16.1x; P¼ 0.004, R2¼ 0.58) or (b) in respired CO2, and the dependence of nitrogen (N) quality (as indicated by NH4

þ productionper SOC) on (c) the C quality (y¼ 0.96x� 10.5; P¼ 0.0007, R2¼ 0.69) and (d) the d13C in soil organic C (y¼�15.4x� 404; P¼0.0003, R2¼0.74). The fitted relationship is a linear type II major axis regression, and is only shown when significant. The O and Mhorizons are shown as solid circles and open circles, respectively. Values represent the mean 6 SE of six (Co) or three (othertreatments) field replicates.

JOHANNES ROUSK AND SERITA D. FREY464 Ecological MonographsVol. 85, No. 3

Page 9: PDF (1.2MB)

0.051). The OA treatment stood out with an approxi-

mately twofold higher SIR biomass than the other

treatments, which were not distinguishable (Table 2).

The microbial biomass per unit SOC, as measured by

total PLFA, depended on both treatment (F5,30¼ 4.4, P

¼ 0.0042) and horizon (F1,30 ¼ 4.4, P ¼ 0.044), with an

interaction showing a larger treatment influence in the

M horizon (F5,30 ¼ 3.4, P ¼ 0.016). The total PLFA

concentration per unit SOC was higher in the O than M

horizon, and ranked from high to low as OA, T0X, 0X,

Co, 2X, and T (Table 2). The bacterial PLFA

concentration per unit SOC was unaffected by treatment

and horizon (Table 2). The fungal PLFA concentration

per unit SOC depended on both treatment (F5,30¼ 22.1,

P , 0.0001) and horizon (F1,30¼ 106, P , 0.0001), with

an interaction (F5,30 ¼ 3.8, P ¼ 0.0083) showing that

treatment effects were larger in the O horizon.

The fungal PLFA concentrations were higher in the O

horizon, and the concentration in the different treat-

ments ordered from high to low concentration as OA,

2X, Co, T, T0X, 0X, (Table 2). Thus fungal PLFA

concentration per unit SOC was higher with litter input

(OA, 2X, Co, T) than without (T0X, 0X). The

ergosterol concentration per unit SOC was affected by

both treatments (F5,30 ¼ 6.5, P ¼ 0.0003) and horizon

(F1,30¼ 47.9, P , 0.0001). The concentration was higher

in O than in M horizons, and the treatments ranked

from high to low concentrations as OA, T, 2X, Co, T0X,

0X (Table 2). Thus, treatments with litter input had

higher concentrations of ergosterol per SOC than those

without litter input, consistent with the fungal PLFA

results.

Microbial PLFA composition and d13C content

The first principal component (PC1) of the PLFA

pattern explained a substantial fraction (30.9%) of the

variation in the data, while PC2 explained 13.8% (Fig.

4a). The C-quality effect on the PLFA composition

aligned closely along PC1, with high C-quality treat-

ments aligning toward high PC1 values, and vice versa.

Thus, the O horizon samples clustered toward high PC1

values and M horizon samples toward low PC1 values.

The litter input treatments (2X, OA, Co) within the

horizons clustered toward high PC1 values, with the

opposite being true for treatments with no litter input

(0X, T0X) within each horizon. This led to significant

effects of both treatment (F5,30¼ 19.8, P , 0.0001) and

horizon (F1,30 ¼ 224, P , 0.0001) on PC1.

The d13C enrichment of PLFA markers did not

depend on treatment or horizon (Fig. 5). There were

systematic differences between the enrichment of differ-

ent PLFA markers (Fig. 5; one-way ANOVA; F18, 779¼13.3, P , 0.0001), where the d13C enrichment ranged

between relatively enriched markers such as 10Me18:0,

16:1x5, a17:0 and i14:0 at�26% to�27%, to relatively

depleted markers, such as 18:1x7, cy19:0 and 18:2x6,9at �31% to �29% (Fig. 5). When aggregated into the

taxonomic categories listed in Fig. 5, there were

significant differences between groups (F6,30 ¼ 14.4, P

, 0.0001), with d13C enrichment ranging from highest to

TABLE 2. Soil microbial characteristics.

TreatmentSIR-biomass(mg C/g SOC)

Total PLFAs(nmol/g SOC)

Bacterial PLFAs(nmol/g SOC)

Fungal PLFA 18:2x6,9(nmol/g SOC)

Ergosterol(lg/g SOC)

Double litter, 2X

O 11.7 (0.73) 1493 (53) 560 (22) 203 (25) 253 (64)M 9.7 (0.38) 1708 (89) 766 (56) 93 (10) 129 (8)

Control, Co

O 13.1 (0.90) 1614 (75) 638 (24) 185 (12) 224 (29)M 9.5 (0.72) 1652 (96) 743 (41) 88 (11) 109 (11)

No litter, 0X

O 10.2 (1.16) 2011 (186) 768 (98) 74 (19) 111 (16)M 7.0 (0.81) 1353 (139) 569 (79) 40 (5) 73 (8)

No OA, OA

O 19.3 (1.26) 2072 (113) 717 (55) 293 (13) 361 (71)M 23.5 (11.49) 2283 (292) 938 (133) 121 (19) 132 (16)

No roots, T

O 12.1 (1.86) 1631 (271) 680 (119) 153 (36) 280 (16)M 7.5 (0.75) 1367 (175) 629 (81) 46 (6) 136 (26)

No input, T0X

O 13.7 (1.15) 2033 (317) 672 (40) 98 (2) 160 (9)M 8.9 (2.02) 1357 (140) 595 (38) 37 (3) 70 (6)

Notes: Values are means (with SE in parentheses) of six (Co) or three (other treatments) field replicates in the O and M horizons.Microbial variables are given per g soil organic carbon (SOC). SIR-biomass is the total microbial biomass as estimated using thesubstrate-induced respiration (SIR) method. Total PLFAs (phospholipid fatty acids) is a measure of the total microbial biomass,while bacterial PLFAs and fungal PLFA represent measures of the subcomponents bacteria and fungi. Ergosterol is a measure oftotal fungal biomass (see methods for further detail).

August 2015 465SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 10: PDF (1.2MB)

lowest as G�/AMF, Actin, Gþ, G�/F, F and G�(taxonomic groups defined in Fig. 5).

C-quality effects on microbial variables

There was a significant influence on the fungal growth

per unit SOC by C-quality, as estimated by respiration

per SOC (Fig. 6a). In contrast, bacterial growth per unit

SOC (Fig. 6b) and the fungal to bacterial growth ratio

(Fig. 6c) were both unaffected by C quality. The

biomass concentration of fungi (Fig. 7a, d) and bacteria

(Fig. 7b) per SOC along with the PLFA composition did

not significantly depend on the C quality. However, we

note clear positive trends with the C quality for

concentrations of fungal PLFA 18:2x6,9 per unit SOC

(Fig. 7a; P¼ 0.06), ergosterol per unit SOC (Fig. 7c; P¼0.16), and for a systematic variation in the PLFA

composition (Fig. 7c; P ¼ 0.09), while the bacterial

biomass per unit SOC contrasted with this pattern by

being highly independent of the C quality (Fig. 7b).

DISCUSSION

DIRT treatment induced effects on SOC quality

The two metrics used to assess SOC quality as

perceived by the microbial decomposer community,

respiration per SOC and d13C in SOM, appeared to be

responsive to the field-experiment manipulations, and

yielded corroborative results. Consequently, our hy-

potheses could be addressed. We hypothesized (H1) that

the decadal plant input manipulations of the DIRT

experiment would result in differences in the SOC

quality, and that relatively high-quality SOC would

benefit bacteria relative to fungi, with the opposite being

true for relatively low-quality SOC. Our results clearly

show that the first part of the hypothesis is supported

with 4.5-fold differences in C quality (as shown by

respiration per unit SOC) that had resulted from the

field treatments. Even within the O horizon, the field

treatments had resulted in threefold differences in SOC

quality (respiration per unit SOC in OA compared with

0X) and, more generally, the treatments with litter (OA,

2X, Co, T) had resulted in SOC with on average a

twofold higher SOC-quality than that resulting from

treatments without litter input (0X, T0X; Fig. 2a). We

also note that the OA treatment (OA removal and

replacement with B horizon soils from nearby) generated

a particularly high C quality. This result is consistent

with early observations a few years after the initiation of

this DIRT experiment (Bowden et al. 1993), along with

recent observations (Lajtha et al. 2014); however, the

underlying mechanisms are neither clear nor possible to

evaluate in the current assessment. One possible reason

could be elevated pH during the past decades, but this

can at present only be substantiated with a marginally

elevated pH in the M horizon compared with other

treatments (Table 1). Interesting follow-up questions

that emerge are whether the physical, chemical, or biotic

mechanisms cause these pronounced differences

(Schmidt et al. 2011). Specifically, it would be valuable

to explore whether the chemistry of annual C inputs is

different between the OA and control or litter addition

treatments, or if biophysical accessibility, related to the

nature of C movement down the soil profile, was

radically affected in the OA treatment. The treatments

where belowground plant C allocation was manipulated

by trenching, in contrast, did not clearly affect the SOC

quality, suggesting that aboveground plant C inputs

dominated the long-term effects on SOC quality in the

studied system relative to belowground plant C inputs.

Although the determination of mechanisms for a C-

quality gradient of SOM was not an objective of the

FIG. 4. The microbial phospholipid fatty acids (PLFA) composition variation. The treatment variation can be seen in a plot ofthe scores of the two first components from a principal component analysis on (a) the microbial PLFA composition and (b) themarkers that drive this sample separation can be seen in a loading plot of the two first components. The O and M horizons areshown as solid circles and open circles, respectively. Values represent the mean 6 SE of six (Co) or three (other treatments) fieldreplicates.

JOHANNES ROUSK AND SERITA D. FREY466 Ecological MonographsVol. 85, No. 3

Page 11: PDF (1.2MB)

present study, it is reassuring that our findings are

consistent with previously reported results on litter-

manipulation experiments based on SOC quality esti-

mations from a previous sampling of the same field

experiment (estimated as time until 1% SOC loss by

respiration in laboratory incubations; Lajtha et al.

2014). Additionally, although the belowground plant

C input has been estimated to contribute substantially to

field measurements of soil respiration in DIRT experi-

ments, its contribution to SOM build-up has been

thought to be small in relation to the contribution by

litter (Sulzman et al. 2005). Moreover, in a recent

assessment of a parallel DIRT experiment in deciduous

forest stands in Oregon, root-derived C had a higher

resistance to decomposition than C derived from litter

(Crow et al. 2009). Irrespective of the underlying

mechanisms by which a gradient in SOC quality was

generated, we do see unambiguous evidence for one

(Fig. 2a, b). Added to this, both indices used to assess

SOC quality, the natural abundance of d13C in SOM

and respiration per SOC, provide corroborative evi-

dence, thus matching previous findings (e.g., Abraham

et al. 1998, Cifuentes and Salata 2001, Sollins et al. 2009,

Gunina and Kuzyakov 2014). We can consequently

proceed to evaluate whether high-quality SOC favors

bacteria, with the reverse being true for fungi (H1), and

track the quality of used SOC with the d13C.

Microbial responses to SOC quality

Fungal growth per unit SOC increased with higher

SOC quality (Fig. 6a), while both bacterial growth per

unit SOC (Fig. 6b) and the fungal-to-bacterial growth

FIG. 5. The d13C in extracted PLFA markers in the (a) O and (b) M horizons. The treatments are color coded according tolegend. Markers associated with specific microbial groups are noted as Gþ (gram positive bacteria), G� (gram negative bacteria),Actin (actinomycetes), AMF (arbuscular mycorrhizal fungi), F (saprotrophic fungi), and Univ (universal). Values represent themean 6 SE of six (Co) or three (other treatments) field replicates.

August 2015 467SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 12: PDF (1.2MB)

ratio (Fig. 6c) showed no systematic pattern with soil C

quality. While the biomass measures of fungi and

bacteria did not show strong patterns with the SOC

quality gradient resulting from the DIRT treatments, we

note clear and reproducible suggestions for higher

fungal biomass per unit SOC with higher-quality C

(Fig. 7a, c), while bacterial biomass appeared completely

independent (Fig. 7c). The d13C microbial PLFA was

not affected by the DIRT treatments (Fig. 5). There

were also very limited differences of d13C among most

microbial PLFA markers (Fig. 5). In minor deviations

from this general pattern, we note a trend for a use of

particularly low-quality SOC by one of two actinomy-

cete markers (i.e., enriched values in 10Me18:0), and a

use of higher-quality SOC suggested by somewhat

depleted values in the fungal marker (PLFA 18:2x6,9;Fig. 5). Added to this, the relative abundance of fungal

PLFAs (PLFA 18:2x6,9) was higher in the high-C-

quality treatments (Fig. 4). It should be noted that

mycorrhizal fungal biomass normally is more depleted

(by 2–3%) than saprotrophic fungal biomass (Hogberg

et al. 1999), which could also have contributed to both

the relative abundance PLFA 18:2x6,9 and its d13Cenrichment. However, together with the results from

saprotrophic fungal growth, the d13C of fungal PLFA

suggested use of higher-quality SOC than that used for

most other PLFAs, further strengthening a consistent

pattern where fungi were benefited relative to bacteria

by high-quality SOC. Thus, several lines of evidence add

up to show that high-quality SOC benefited fungal

dominance over bacterial, in direct contrast with our

hypothesis (H1).

SOC use by fungal and bacterial decomposers

We hypothesized that the CO2 produced in treatments

that had resulted in lower-quality SOC with relatively

higher dominance of fungi would have a more enriched

d13C, indicative of more recalcitrant C being mineralized

(H2). In contrast, the d13C in respired CO2 appeared to

be independent of the SOC quality (Fig. 3d). This

suggests that in treatments with a reduction in SOC

quality (i.e., a lower relative abundance of high-quality

C), less C was mineralized without any change in the

quality of the C used. This result is consistent with

results from previous assessments, which have found

similar d13C in respired CO2 in spite of d13C differences

in substrates respired (Fernandez et al. 2003, Church-

land et al. 2013), and undistinguishable d13C in CO2

produced from different DIRT treatments in short term

assessments of fresh soil (Crow et al. 2006). Thus we see

evidence for a selective microbial use of high-quality

components of the SOC, with a resulting respiration

dominated by the abundance of high-quality C remain-

ing, as suggested by the stable d13C (Fig. 3b) but

variable rate (Fig. 2a, b) of respired CO2. As previously

pointed out (e.g., Fernandez et al. 2003), the selective

microbial processing of SOC with a relatively depleted

d13C can also explain how the pattern between apparent

FIG. 6. The dependence of (a) fungal growth (y ¼ 8.1x þ0.34; R2 ¼ 0.38, P ¼ 0.034, F ¼ 6.01, h ¼ 12), (b) bacterialgrowth, and (c) the fungal-to-bacteria growth ratio on carbon(C) quality (as indicated by respiration per SOC). The fittedrelationship is a linear type II major axis regression, and is onlyshown when significant. The O and M horizons are shown assolid circles and open circles, respectively. Values represent themean 6 SE of six (Co) or three (other treatments) fieldreplicates.

JOHANNES ROUSK AND SERITA D. FREY468 Ecological MonographsVol. 85, No. 3

Page 13: PDF (1.2MB)

SOC quality and its d13C content (Fig. 3a) has formed.

To evaluate our second hypothesis, we see no difference

between the SOC-quality used by a fungal compared

with a bacterial-dominated decomposer community.

Nutrient cycling of fungal and bacterial decomposers

We hypothesized that a relatively more bacterial-

dominated decomposer system should be characterized

by a nutrient cycle of lower retention, as shown by

higher C and N mineralization and a higher potential for

mineral N leaching (Wardle et al. 2004, de Vries et al.

2012). Since mineralized organic matter is composed of

both C and N, it should naturally be expected that C

mineralization and gross N mineralization should be

well correlated (Hart et al. 1994). Our results confirm

this expectation (R2¼ 0.69, P¼ 0.0007, n¼ 12; Fig. 3c).

We observed treatment related differences in the NH4þ

production, and thus the potential for mineral N

leaching, that were positively related with SOC quality.

This contrasted with previous assessments of other

DIRT experiments where no treatment differences were

discerned (Holub et al. 2005). However, those previous

assessments only included the mineral horizons, where

we also observed less-pronounced treatment-induced

differences. The rather higher C:N ratio of mineralized

SOM (;25; Fig. 3c) relative to the C:N of the SOM

(;15; Table 1) corroborates the evidence from the

depleted d13C signal in the respired CO2 (Fig. 3b); both

suggesting that plant-like material dominates the micro-

bial SOM use (Hogberg et al. 1999, Gunina and

Kuzyakov 2014).

As has been previously noted, estimating NH4þ

consumption with the presently used approach is biased

by the increased NH4þ-concentration that addition of

15N tracers induces. Thus, it likely results in overesti-

mated consumption rates (Hart et al. 1994, Holub et al.

2005), which could have compromised the sensitivity of

NH4þ consumption estimates to capture treatment

differences (Fig. 2e, f ). While no relationship between

NH4þ consumption and any of our used SOC quality

indices could found, clear positive relationships were

seen to microbial growth; particularly to bacterial

growth. So, did the detected differences in N transfor-

mation rates coincide with shifts in bacterial dominance?

There was no relationship of N or C mineralization with

measured bacterial variables (Figs. 1 and 2). Respiration

(C mineralization) was closely related to fungal mea-

sures, with a corroborative tendency also found for

FIG. 7. The dependence of (a) the fungal 18:2x6,9 concentration (b) bacterial PLFA concentration and (c) the PLFAcomposition (PC1 case scores), and (d) fungal ergosterol concentration on carbon (C) quality (as indicated by respiration perSOC). The O and M horizons are shown as solid circles and open circles, respectively. Values represent the mean 6 SE of six (Co)or three (other treatments) field replicates.

August 2015 469SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 14: PDF (1.2MB)

NH4þ production. Thus, the propensity for the ecosys-

tem to loose both C and N increased with a higher

fungal dominance, in contrast with our hypothesis (H3).

Moreover, the propensity for the system to retain

mineral N (NH4þ consumption rate) was more positively

related to bacterial than to fungal growth, again in

contrast with the hypothesis (H3).

Ecosystem implications of fungal-to-bacterial

dominance revised

To summarize and to evaluate our set of hypotheses

together, we found no support for a positive relationship

between the relative dominance of bacteria and high

SOC quality (H1). The SOC decomposed in relatively

more fungal-dominated soils is not different from that

degraded in relatively more bacterial-dominated systems

(H2). The C and N mineralization, and thus the

potential for C and N leakage, in relatively fungal-

dominated systems was not lower than that occurring in

relatively bacterial-dominated ones; and fungal-domi-

nated systems were not better than bacterial-dominated

systems at retaining mineral N (H3). These results

suggest structural differences between soil microbial

characteristics driven by the age and quality of C input,

that also carry implications for system biogeochemistry.

Specifically, increased fresh plant C influx will promote a

shift toward a fungal dominance of SOM turnover. Such

changes could be induced by increased forest produc-

tivity (e.g., in warmer conditions) or induced by biomass

harvest and subsequent recolonization (as indicated by

the OA treatment). A shift toward fungal dominance

will also lead to a faster mineralization of C and N, and

a somewhat lower microbial retention of mineral N.

These findings contrast with widely held beliefs within

soil microbial ecology. Detrital food webs dominated by

bacteria are thought to support high turnover rates of

easily available substrates, while slower fungal-domi-

nated decomposition pathways support the decomposi-

tion of more complex organic material (Wardle et al.

2004, de Graaff et al. 2010, de Vries et al. 2012). The

crude phylogenetic resolution of fungal and bacterial

decomposers has been thought to capture disparate life

strategies for soil decomposers, associating bacteria with

an r-selected strategy for C and nutrient use, and fungi

with a K-selected strategy (Fontaine et al. 2003,

Fontaine and Barot 2005, Fierer et al. 2009). These

assumptions have been incorporated into food-web

models, where the different properties of fungi and

bacteria with regard to C sequestration, nutrient

turnover, and ecosystem stability have been concluded

(Moore et al. 2004, Six et al. 2006, Neutel et al. 2007,

Bezemer et al. 2010). This influential work has

manifested a widely held belief that bacterial decom-

poser pathways in soil support high turnover rates of

easily available substrates, while slower fungal-domi-

nated decomposition pathways support the decomposi-

tion of more complex organic material (Wardle et al.

2004, de Graaff et al. 2010, de Vries et al. 2012). We note

that these assumptions about the microbial processing

of SOM rarely have been tested, or, when they have,

either the microbial use of SOM has been inferred from

the colonization of plant litter (e.g., de Graaff et al.

2010) rather than SOM or biomass concentrations of

fungi and bacteria have been assumed to reflect their

contribution to SOM turnover (e.g., Wardle et al. 2004,

de Vries et al. 2012), an assumption that is not met in

soil (Rousk and Baath 2007, 2011).

We find no evidence in favor of the supposition that

bacterial-dominated food webs cycle faster and leak

more C and N than fungal-dominated ones, and

convincing suggestions for the opposite, in the studied

forest soil. Our results highlight the fact that many of the

important dogma held as true in soil microbial ecology

have limited empirical support (Strickland and Rousk

2010). In one of the few empirical validations of another

dogma, that fungi have a higher C-use efficiency than

bacteria (Six et al. 2006), Thiet et al. (2006) found that

in fact the C-use efficiency of bacterial and fungal-

dominated decomposer systems could not be distin-

guished. Taken together with the results we report here,

the lack of support for central assumptions about the

properties of the soil decomposer food web can only be

regarded as humbling. Assumptions about the soil

decomposer community (e.g., C-use efficiency, nutrient

retention, etc.) influence predictions of ecosystem

models (Allison 2012, Wieder et al. 2013) and the

empirical basis for these assumptions is still weak.

ACKNOWLEDGMENTS

We thank Mel Knorr for field, technical, and laboratoryassistance, and Dr. Per Bengtson for stable isotopes methodadvice. This study was supported by the Swedish ResearchCouncil (Vetenskapsradet) (grant no 621-2011-5719), the LundUniversity Infrastructure Grant, the Royal PhysiographicSociety of Lund, and by the U.S. National Science FoundationLong-Term Ecological Research (LTER) Program.

LITERATURE CITED

Abraham, W. R., C. Hesse, and O. Pelz. 1998. Ratios of carbonisotopes in microbial lipids as an indicator of substrate usage.Applied and Environmental Microbiology 64:4202–4209.

Agren, G., E. Bosatta, and J. Balesdent. 1996. Isotopediscrimination during decomposition of organic matter: atheoretical analysis. Soil Science Society of America Journal60:1121–1126.

Allison, S. D. 2012. A trait-based approach for modellingmicrobial litter decomposition. Ecology Letters 15:1058–1070.

Anderson, J. P. E., and K. H. Domsch. 1978. A physiologicalmethod for quantitative measurement of microbial biomassin soils. Soil Biology & Biochemistry 10:215–221.

Baath, E. 2001. Estimation of fungal growth rates in soil using14C-acetate incorporation into ergosterol. Soil Biology &Biochemistry 33:2011–2018.

Baath, E., M. Pettersson, and K. H. Soderberg. 2001.Adaptation of a rapid and economical microcentrifugationmethod to measure thymidine and leucine incorporation bysoil bacteria. Soil Biology & Biochemistry 33:1571–1574.

Bengtson P, U. Falkengren-Grerup, and G. Bengtsson. 2005.Relieving substrate limitation—soil moisture and tempera-ture determine gross N transformation rates. Oikos 111:81–90.

JOHANNES ROUSK AND SERITA D. FREY470 Ecological MonographsVol. 85, No. 3

Page 15: PDF (1.2MB)

Bezemer, T. M., et al. 2010. Divergent composition but similarfunction of soil food webs of individual plants: plant speciesand community effects. Ecology 91:3027–3036.

Blagodatskaya, E., and Y. Kuzyakov. 2008. Mechanisms ofreal and apparent priming effects and their dependence onsoil microbial biomass and community structure: criticalreview. Biology and Fertility of Soils 45:115–131.

Bowden, R. D., K. J. Nadelhoffer, R. D. Boone, J. M. Meillo,and J. B. Garrison. 1993. Contributions of abovegroundlitter, belowground litter, and root respiration to total soilrespiration in a temperate mixed hardwood forest. CanadianJournal of Forest Research 23:1402–1407.

Churchland, C., S. J. Grayston, and P. Bengtson. 2013. Spatialvariability of soil fungal and bacterial abundance: conse-quences for carbon turnover along a transition from aforested to clear-cut site. Soil Biology & Biochemistry 63:5–13.

Cifuentes, L. A., and G. C. Salata. 2001. Significance of carbonisotope discrimination between bulk carbon and extractedphospholipid fatty acids in selected terrestrial and marineenvironments. Organic Geochemistry 32:613–621.

Cotrufo, M. F., M. D. Wallenstein, C. M. Boot, K. Denef, andE. Paul. 2013. The Microbial Efficiency-Matrix Stabilization(MEMS) framework integrates plant litter decompositionwith soil organic matter stabilization: Do labile plant inputsform stable soil organic matter? Global Change Biology19:988–995.

Crow, S. E., K. Lajtha, T. R. Filley, C. W. Swanson, R. D.Bowden, and B. A. Caldwell. 2009. Sources of plant-derivedcarbon and stability of organic matter in soil: implications forglobal change. Global Change Biology 15:2003–2019.

Crow, S. E., E. W. Sulzman, W. D. Rugh, R. D. Bowden, andK. Lajtha. 2006. Isotopic analysis of respired CO2 duringdecomposition of separated soil organic matter pools. SoilBiology & Biochemistry 38:3279–3291.

de Graaff, M.-A., A. T. Classen, H. F. Castro, and C. W.Schadt. 2010. Labile soil carbon inputs mediate the soilmicrobial community composition and plant residue decom-position rates. New Phytologist 188:1055–1064.

de Vries, F. T., J. Bloem, H. Quirk, C. J. Stevens, R. Bol, andR. D. Bardgett. 2012. Extensive management promotes plantand microbial nitrogen retention in temperate grassland.PLoS ONE 7:e51201.

Demoling, F., D. Figueroa, and E. Baath. 2007. Comparison offactors limiting bacterial growth in different soils. SoilBiology & Biochemistry 39:2485–2495.

Denman, K. L., et al. 2007. Couplings between changes in theclimate system and biogeochemistry. Pages 663–754 in S.Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K.B.Averyt, M. Tignor and H .L. Miller, editors. Climate change2007: the physical science basis. Contribution of WorkingGroup I to the Fourth Assessment Report of the Intergov-ernmental Panel on Climate Change. Cambridge UniversityPress, Cambridge, UK.

Egli, T. 2010. How to live at very low substrate concentrations.Water Research 44:4826–4837.

Eilers, K. G., C. L. Lauber, R. Knight, and N. Fierer. 2010.Shifts in bacterial community structure associated withinputs of low molecular weight carbon compounds to soil.Soil Biology & Biochemistry 42:896–903.

Farrar, J. E., M. Hawes, D. L. Jones, and S. Lindow. 2003.How roots control the flux of carbon to the rhizosphere.Ecology 84:827–837.

Fernandez, I., N. Mahieu, and G. Cadisch. 2003. Carbonisotope fractionation during decomposition of plant materi-als of different quality. Global Biogeochemical Cycles17:1075–1085.

Fierer, N., M. A. Bradford, and R. B. Jackson. 2007. Towardan ecological classification of soil bacteria. Ecology 88:1354–1364.

Fierer, N., M. S. Strickland, D. Liptzin, M. A. Bradford, andC. C. Cleveland. 2009. Global patterns in belowgroundcommunities. Ecology Letters 12:1238–1249.

Fontaine, S., and S. Barot. 2005. Size and functional diversityof microbe populations control plant persistence and carbonaccumulation. Ecology Letters 8:1075–1087.

Fontaine, S., A. Mariotti, and L. Abbadie. 2003. The primingeffect of soil organic matter: a question of microbialcompetition? Soil Biology & Biochemistry 35:837–843.

Frostegard, A., and E. Baath. 1996. The use of phospholipidfatty acid analysis to estimate bacterial and fungal biomass insoil. Biology and Fertility of Soils 22:59–65.

Frostegard, A., A. Tunlid, and E. Baath. 1993. Phospholipidfatty-acid composition, biomass, and activity of microbialcommunities from 2 soil types experimentally exposed todifferent heavy-metals. Applied and Environmental Micro-biology 59:3605–3617.

Gunina, A., and Y. Kuzyakov. 2014. Pathways of litter C byformation of aggregates and SOM density fractions: Impli-cations from 13C natural abundance. Soil Biology &Biochemistry 71:95–104.

Hart, S. C., G. E. Nason, D. D. Myrold, and D. A. Perry. 1994.Dynamics of gross nitrogen transformation in an old-growthforest: the carbon connection. Ecology 75:880–891.

Hayes, J. M. 2001. Fractionation of carbon and hydrogenisotopes in biosynthetic processes. Stable Isotope Geochem-istry 16:1115–1128.

Henriksen, T. M., and T. A. Breland. 1999. Evaluation ofcriteria for describing crop residue degradability in a modelof carbon and nitrogen turnover in soil. Soil Biology &Biochemistry 31:1135–1149.

Hieber, M., and M. O. Gessner. 2002. Contribution of streamdetrivores, fungi, and bacteria to leaf breakdown based onbiomass estimates. Ecology 83:1026–1038.

Hobbie, J. E., and E. A. Hobbie. 2012. Amino acid cycling inplankton and soil microbes studied with radioisotopes:measured amino acids in soil do not reflect bioavailability.Biogeochemistry 107:339–360.

Hogberg, P., A. G. Plamboeck, A. F. S. Taylor, and P. M. A.Fransson. 1999. Natural 13C abundance reveals trophicstatus of fungi and host-origin of carbon in mycorrhizalfungi in mixed forests. Proceedings of the National Academyof Sciences USA 96:8534–8539.

Holub, S. M., K. Lajtha, J. D. H. Spears, J. A. Toth, S. E.Crow, B. A. Caldwell, M. Papp, and P. T. Nagy. 2005.Organic matter manipulations have little effect on gross andnet nitrogen transformations in two temperate forest mineralsoils in the USA and central Europe. Forest Ecology andManagement 214:320–330.

IAEA. 2001. Use of isotope and radiation methods in soil andwater management and crop nutrition. IAEA TrainingCourse Series No. 14. International Atomic Energy Agency,Vienna, Austria.

Inselbacher, E., and T. Nasholm. 2012. The below-groundperspective of forest plants: soil provides mainly organicnitrogen for plants and mycorrhizal fungi. New Phytologist195:329–334.

Jenkinson, D. S., J. Meredith, J. I. Kinyamario, G. P. Warren,M. T. F. Wong, D. D. Harkness, R. Bol, and K. Coleman.1999. Estimating net primary production from measurementsmade on soil organic-matter. Ecology 80:2762–2773.

Keiser, S. D., D. A. Keiser, M. S. Strickland, and M. A.Bradford. 2014. Disentangling the mechanisms underlyingfunctional differences between decomposer communities.Journal of Ecology 102:603–609.

Kirchman, D., E. Knees, and R. Hodson. 1985. Leucineincorporation and its potential as a measure of protein-synthesis by bacteria in natural aquatic systems. Applied andEnvironmental Microbiology 49:599–607.

Lajtha, K., R. D. Bowden, and K. Nadelhoffer. 2014. Litterand root manipulations provide insights into soil organic

August 2015 471SOIL C CONTROLS ON FUNGI AND BACTERIA

Page 16: PDF (1.2MB)

matter dynamics and stability. Soil Science Society ofAmerica Journal 78:S261–S269.

Le Guillou, C., D. A. Angers, P. A. Maron, P. Leterme, and S.Menasseri-Aubry. 2012. Linking microbial community to soilwater-stable aggregation during crop residue decomposition.Soil Biology & Biochemistry 50:126–133.

Moore, J. C., et al. 2004. Detritus, trophic dynamics andbiodiversity. Ecology Letters 7:584–600.

Neutel, A. M., J. A. P. Heesterbeek, J. van de Koppel, G.Hoenderboom, A. Vos, C. Kaldeway, F. Berendse, and P. C.de Ruiter. 2007. Reconciling complexity with stability innaturally assembling food webs. Nature 449:599–602.

Newell, S. Y., and R. D. Fallon. 1991. Toward a method formeasuring instantaneous fungal growth-rates in field sam-ples. Ecology 72:1547–1559.

Nilsson, L. O., E. Baath, U. Falkengren-Grerup, and H.Wallander. 2007. Growth of ectomycorrhizal mycelia andcomposition of soil microbial communities in oak forest soilsalong a nitrogen deposition gradient. Oecologia 153:375–384.

Reischke, S., J. Rousk, and E. Baath. 2014. The effects ofglucose loading rates in bacterial and fungal growth in soil.Soil Biology & Biochemistry 70:88–95.

Romani, A. M., H. Fischer, C. Mille-Lindblom, and L. J.Tranvik. 2006. Interactions of bacteria and fungi ondecomposing litter: differential extracellular enzyme activi-ties. Ecology 87:2559–2569.

Rousk, J., and E. Baath. 2007. Fungal and bacterial growth insoil with plant materials of different C/N ratios. FEMSMicrobiology Ecology 62:258–267.

Rousk, J., and E. Baath. 2011. Growth of saprotrophic fungiand bacteria in soil. FEMS Microbiology Ecology 78:17–30.

Rousk, J., P. C. Brookes, and E. Baath E. 2009. Contrastingsoil pH effects on fungal and bacterial growth suggestsfunctional redundancy in carbon mineralization. Applied andEnvironmental Microbiology 75:1589–1596.

Rousk, J., P. W. Hill, and D. L. Jones. 2015. Priming of thedecomposition of ageing soil organic matter: concentrationdependence and microbial control. Functional Ecology29:285–296.

Rousk, K., J. Rousk, D. L. Jones, O. Zackrisson, and T. H.DeLuca. 2013. Feather moss nitrogen acquisition acrossnatural fertility gradients in boreal forests. Soil Biology &Biochemistry 61:86–95.

Ruess, L., and P. M. Chamberlain. 2010. The fat that matters:soil food web analysis using fatty acids and their carbonstable isotope signal. Soil Biology & Biochemistry 42:1898–1910.

Sayer, E. J., M. S. Heard, H. K. Grant, T. R. Marthews, andE. V. J. Tanner. 2011. Soil carbon release enhanced byincreased tropical forest litterfall. Nature Climate Change1:304–307.

Schiff, S., R. Aravena, S. Trumbore, M. Hinton, R. Elgood,and P. Dillon. 1997. Export of DOC from forestedcatchments on the Precambrian shield of central Ontario:clues from C-13 and C-14. Biogeochemistry 36:43–65.

Schlesinger, W. H., and J. A. Andrews. 2000. Soil respirationand the global carbon cycle. Biogeochemistry 48:7–20.

Schmidt, M. W. I., et al. 2011. Persistence of soil organic matteras an ecosystem property. Nature 473:49–56.

Six, J., S. D. Frey, R. K. Thiet, and K. M. Batten. 2006.Bacterial and fungal contributions to carbon sequestration inagroecosystems. Soil Science Society of America Journal70:555–569.

Sollins, P., M. G. Kramer, C. Swanston, K. Lajtha, T. Filley,A. K. Aufdenkampe, R. Wagai, and R. D. Bowden. 2009.Sequential density fractionation across soils of contrastingmineralogy: evidence for both microbial- and mineral-controlled soil organic matter stabilization. Biogeochemistry96:209–231.

Strickland, M. S., C. Lauber, N. Fierer, and M. A. Bradford.2009. Testing the functional significance of microbialcommunity composition. Ecology 90:441–451.

Strickland, M. S., and J. Rousk. 2010. Considering fungal : bac-terial dominance in soils—methods, controls, and ecosystemimplications. Soil Biology & Biochemistry 42:1385–1395.

Sulzman, E. W., J. B. Brant, R. D. Bowden, and K. Lajtha.2005. Contribution of aboveground litter, belowground litter,and rhizosphere respiration to total soil CO2 efflux in an oldgrowth coniferous forest. Biogeochemistry 73:231–256.

Thiet, R. K., S. D. Frey, and J. Six. 2006. Do growth fieldefficiencies differ between microbial communities infungal : bacterial ratios? Reality check and methodologicalissues. Soil Biology & Biochemistry 38:837–844.

Waksman, S. A., and F. C. Gerretsen. 1931. Influence oftemperature and moisture upon the nature and extent ofdecomposition of plant residues by microorganisms. Ecology12:33–60.

Waksman, S. A., and R. L. Starkey. 1932. Partial sterilizationof soil, microbiological activities and soil fertility. II. SoilScience 16:247–268.

Wardle, D. A., R. D. Bardgett, J. N. Klironomos, H. Setala,W. H. van der Putten, and D. H. Wall. 2004. Ecologicallinkages between aboveground and belowground biota.Science 304:1629–1633.

Wieder, W. R., G. B. Bonan, and S. D. Allison. 2013. Globalsoil carbon predictions are improved by modelling microbialprocesses. Nature Climate Change 3:909–912.

Williams, M. A., D. D. Myrold, and P. J. Bottomley. 2006.Carbon flow from 13C-labelled straw and root residues intothe phospholipid fatty acids of a soil microbial communityunder field conditions. Soil Biology & Biochemistry 38:759–768.

JOHANNES ROUSK AND SERITA D. FREY472 Ecological MonographsVol. 85, No. 3