You are on page 1of 10

applied soil ecology 41 (2009) 239–248

available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/apsoil

Impact of biotic and abiotic interaction on soil microbial


communities and functions: A field study

Brajesh K. Singh *, Lorna A. Dawson, Catriona A. Macdonald 1, Sarah M. Buckland


Macaulay Institute, Craigiebuckler, Aberdeen AB15 8QH, United Kingdom

article info abstract

Article history: Interactions between plants, soils and microbes regulate terrestrial ecosystem functioning.
Received 18 April 2008 Biotic and abiotic interactions can strongly affect the community structure which in turn
Received in revised form will impact on ecological processes. Plant species with different ecophysiological traits can
7 October 2008 exert strong effects on soil biological properties. Our objective was to investigate and
Accepted 9 October 2008 identify the effects of different biotic and abiotic variables on soil microbial community
structure and functions and to examine if plant species with different physiological traits
support different microbial communities in soils. Here, we show that 3 years of the presence
Keywords: of plants had direct impacts on soil function in terms of total heterotrophic respiration and
Biotic and abiotic factors on microbial biomass and microbial community structure. However, the plant species-
Ecological interactions specific impact on bacterial community structure was weak, and differences were mainly
Plant physiological traits driven by sample field location. The fungal community analysis gave similar results, with
Fungal community soil location being the most important factor driving fungal community structure. The effect
Bacterial community of plant species on fungal community structure was weak but statistically significant. There
Plant species was a strong concordance between bacterial and fungal communities (P < 0.001) which
suggested that the bacteria and fungi have an influence on shaping the structure of each
other’s community. Among the abiotic factors, moisture had a comparatively higher impact
on bacterial communities compared to soil N and C. However, the fungal community was
not affected by the soil moisture but soil N and C had a stronger impact than on the bacterial
community. These results indicate that the microbial community structure in the natural
environment is influenced by interactions between both biotic and abiotic factors.
# 2008 Elsevier B.V. All rights reserved.

1. Introduction position, which may represent a positive association where


plants supply carbon for microbial growth and microorgan-
Microbial community structure and ecological functions are isms in return provide major elements such as N and P, as well
influenced by interactions between above and belowground as protection against pathogen and parasite attacks (Singh
biota (Wardle et al., 2002). These interactions can have et al., 2004). It is believed that the carbon source made
positive, negative or neutral impact on diversity and commu- available by plants drives many complex chemical and
nity structure of plants and soil microbes (Singh et al., 2004). It biological interactions in the soil (e.g. Jones et al., 2004). These
has been observed that many plants select beneficial groups of include sustaining a complex food web of prokaryotes and
microorganisms via rhizodeposition (loss of carbon due to root eukaryotes, the composition of which is thought to be
exudation, leakage or root decomposition), and litter decom- regulated by complex signalling (Phillips et al., 2003). However,

* Corresponding author. Tel.: +44 1224 498200; fax: +44 1224 498207.
E-mail address: b.singh@macaulay.ac.uk (B.K. Singh).
1
Present address: Rothamsted Research, Harpenden AL5 2JQ, UK.
0929-1393/$ – see front matter # 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsoil.2008.10.003
240 applied soil ecology 41 (2009) 239–248

it is not known if plants actively select for beneficial soil 2. Materials and methods
microbial communities. Significant change in soil microbial
community structures has been reported with change in 2.1. Site, soil properties and plant species
vegetative cover. These vegetation-induced changes have
been observed both at the universal bacterial group level This study was carried out at an experimental site at
(Grayston et al., 1998, 2004; Nusslein and Tiedje, 1999) as well Sourhope, Scotland (558280 3200 N 28140 4300 W), a semi-natural
as at the functional group level (Singh et al., 2007a). However, Festuca ovina–Agrostis capillaries–Galium saxatile grassland,
another study reported similar microbial community struc- determined as a Luzula multiflora–Rhytidiadelphus loreus sub-
ture in soils which shared similar agricultural management community (National Vegetation Classification (NVC) U4d),
practices despite differences in above-ground community also the location of the recent NERC Soil Biodiversity
composition (Buckley and Schmidt, 2001). Similarly, Girvan Programme (Usher et al., 2006). The site is a long-term
et al. (2003) found that soil type is the primary determinant of (>200 year old) grassland, and in 1998 a fence was erected
the composition of the total bacterial community in arable to exclude grazing animals (e.g. livestock, deer, rabbits and
soils. Most of the previous studies focused only on the hares). It is situated at 309 m above sea level and varies in slope
bacterial community (Buckley and Schmidt, 2001; Girvan from 88 at the top to 48 at the foot and has a northerly aspect,
et al., 2003; Nunan et al., 2005), while a few exclusively studied mean annual rainfall is 954 mm. The soils are developed on a
universal fungal communities (Smit et al., 1999; Marcial- drift locally derived from andesitic lavas of old red sandstone
Gomes et al., 2003). There are only few reports of investiga- and are characterised as acid brown forest soil belonging to
tions on the impact of both plant species and soil type on both the Sourhope series (SH 74711) (pH 4.5–5.0) (Grayston et al.,
fungal and bacterial communities (Mougel et al., 2006; Costa 2004).
et al., 2006; Singh et al., 2007b). To our knowledge none has In 1999, the site was fenced to prevent sheep grazing and
investigated the concordance between bacterial and free- set out in five replicate blocks, each block being placed along
living fungal communities. With the growing evidence for the the contour of the slope. Experimental plots (1 m2) were
linked ecological functions of bacterial and fungal commu- arranged in a randomised block design with nine treatments
nities in processes such as decomposition and nutrient in each of the five blocks. Nine treatments involved creating
cycling, it is important to understand whether both commu- monocultures, eight of which were common herbaceous
nities are influenced by the same environmental factors (Costa species at the Sourhope site: Agrostis capillaris (Ac), Anthox-
et al., 2006). anthum odoratum (Ao), Festuca ovina (Fo), Festuca rubra (Fr),
The influence of the plant community on ecosystem Nardus stricta (Ns), Luzula multiflora (Lm), Trifolium repens (Tr),
functions is well known (Horner-Devine et al., 2003; Tilman Rumex acetosa (Ra); and a ninth species which was present in
et al., 2006). Furthermore, it is well established that plant the improved areas of pasture, Lolium perenne (Lp). These
species which dominate different stages of succession have species also represent a broad range of physiological and
different sets of ecophysiological traits (Aber et al., 1990; Van functional types, illustrated by the contrasted values in key
Vuuren et al., 1992; Grime et al., 2007) and that these sets of plant traits (Table 1). The monoculture plots were created by
traits can exert strong effects on soil biological properties removing the surface 2–3 cm of turf and inserting vegetative
(Wardle et al., 1998; Bardgett et al., 1999; Hobbie, 2002; Garnier specimens collected from the surrounding natural vegetation
et al., 2004) such as selecting for decomposer food webs with contained in the Sourhope enclosure, washing and brushing
certain basic attributes (Wardle et al., 2002). There are off as much soil as possible, before transplanting into their
evidences that plant species differ considerably in the respective plots. Monocultures of L. perenne were established
composition of soil biota that they support (Grayston et al., by sowing in seed (60 g per plot). A 10th treatment included
1998; Wardle et al., 1998; Marilley et al., 1998; Bardgett et al., plots of undisturbed, natural (N) turf, to provide data on the
1999; Johnson et al., 2003; Vandenkoornhuyse et al., 2003). natural soil community in the absence of turf removal. A 11th
Other studies found mainly mycorrhizal fungi (Van der treatment comprised bare (B) plots, to provide data on the
Heijden et al., 1998a,b) or bacteria (Van der Heijden et al., effect of turf-stripping without reintroduction of transplants.
2006) to be key determinants of ecosystem diversity and The plots were maintained by regular hand-weeding to
functions. Although soil ecologists have long recognised the promote ongoing establishment of the monocultures and to
need to understand the mechanisms by which community prevent vegetation colonization on the bare plots. Plots were
structure is shaped and ecosystem functions are maintained, mown monthly between May and September, following the
study on multi-community and multi-trophic interactions management design implemented on the NERC soil biodiver-
between above and belowground communities and their sity experiment.
impact on ecosystem function is lacking. In the present study,
we studied the impact of grass species with differential 2.2. Measurement of soil respiration, microbial biomass,
physiological traits (in monoculture and as a natural mixed soil N, soil C and moisture
community) and soil abiotic factors (soil C, N and moisture) on
soil bacterial and fungal communities under field conditions. Three years after the treatments were established, soil
We also investigated the interactions between bacterial and samples were collected from each field-block by a (3.5 cm
fungal communities and the impacts on shaping microbial diameter  8 cm depth) corer. Roots were removed by hand
community structure. Soil functional capabilities under picking. From each plot, four cores were taken and stored in
different grass species were also measured in terms of plastic bags at 4 8C until analyses were completed. For
respiration rate and total microbial biomass. molecular work, sub-samples were frozen at 20 8C until
applied soil ecology 41 (2009) 239–248 241

Table 1 – Plant species data for traits considered to have a significant influence on the rate of nutrient cycling.
Plant spp RGRa SLAb DMCb Leaf thicknessc Leaf toughnessc Leaf Nd Leaf Pd
(mm) (N mm1) (% d. wt) (% d. wt)

Agrostis capillaris 1.36 30.8 25.1 0.012 1.16 2.33 0.22


Anthoxanthum odoratum 0.94 27.3 22.0 0.098 5.86 2.04 0.15
Festuca ovina 1.00 10.9 41.0 0.300 7.79 1.46 0.09
Festuca rubra 1.18 17.7 26.7 0.179 6.32 1.69 0.22
Lolium perenne 1.30 26.4 21.2 0.134 1.15 3.27 0.44
Luzula multiflorae 1.01 21.6 21.4 0.222 0.97 1.95 0.20
Nardus stricta 0.71 10.2 38.8 0.256 11.72 1.59 0.17
Rumex acetosa 1.71 28.1 10.3 0.312 0.31 3.37 0.32
Trifolium repens 1.26 38.8 17.9 0.131 0.10 NA NA

Abbreviations: RGR, relative growth rate (unit: week1); SLA, specific leaf area (unit: mm2 mg1); DMC, dry matter contents (%); (N mm1),
Newton per millimeter.
a
Data source: Grime and Hunt (1975).
b
Data source: Grime et al. (2007).
c
Data source: Hodgson, personal communication.
d
Data source: Thompson et al. (1997).
e
Data have been taken from L. campestris, a closely related spp.

DNA extraction. Gravimetric soil moisture content was 2.4. TRFLP—analysis for bacterial and fungal communities
determined on field fresh soil (ISO, 1997). The total soil C
and N contents of the each sample was determined by Microbial community fingerprint patterns (based on the 16S
automated Duma Combustion using a Carlo-Erba Elemental rRNA gene for bacterial communities and the ITS gene for fungal
Analyser (Carbo-Erba Instruments, Milan, Italy; Grayston communities) were generated by TRFLP (Terminal Restriction
et al., 2004). Fragment Length Polymorphism) analysis. Initially, a range of
Soil respiration was determined on 20 g soil samples (after restriction enzymes (RsaI, HaeIII, HhaI, MspI and TaqI; Promega,
sieving though a 4 mm mesh) following a conditioning period Southampton, UK) were used on randomly selected samples to
of 7 days in the dark at 25 8C and constant humidity and CO2 screen for the best enzyme for TRFLP analysis. HhaI and TaqI
concentration. The head space of the soil jars was sampled produced best results for bacterial and fungal communities,
(1 ml) at 0, 2, 4 and 6 h. The CO2 was determined by GC analysis respectively, on the basis of number, consistency and even
and respiration rates calculated (Sparling, 1981). Soil microbial distribution of terminal fragments and, therefore, all further
biomass was measured by chloroform-fumigation extraction analyses were carried out with these two restriction enzymes.
(Brookes et al., 1985). Approximately 1000 ng of purified DNA was digested either with
HhaI (for bacterial) or TaqI (for fungal). Each 20 ml reaction mix
2.3. DNA extraction and PCR consisted of the appropriate volume of DNA (for 1000 ng) mixed
with 2 ml of appropriate enzyme, 2 ml of 10 buffer and 0.2 ml of
DNA was extracted from 0.5 g of each soil sample by using soil BSA and the volume was made up by addition of sterile water.
DNA clean kit (Mo Bio, Carlsbad, California) according to the Samples were then incubated for 3 h at optimal temperature
manufacturers’ instructions. Extracted DNA was PCR amplified (37 8C for bacterial DNA and 65 8C for fungal DNA) as per enzyme
using universal bacterial 50 FAM (6-carboxyfluorescein) labelled supplier’s instructions on a PCR machine followed by a 15 min
forward primer 63F (50 -AGG CCT AAC ACA TGC AAG TC-30 ) and deactivation period at 95 8C. Aliquots (1 ml) of digested PCR
reverse primer 1494R (50 -TACGG YTACC TTGTT ACGAC-30 ) products were mixed with 12 ml of formamide loading dyes (ABI,
(Operon, Germany). Each 100 ml reaction mixture contained UK) and 1 ml of internal size standard (2500 TAMARA, ABI) before
10 ml 10 PCR buffer (Bioline, London, UK), 2 ml dNTP (0.2 mM), denaturing for 5 min at 95 8C. TRFLP analysis was carried out on
8 ml MgCl2 (25 mM), 2 ml BSA (Roche Diagnostic, Lewes, UK), 2 ml an automated sequencer, ABI PRISM1 3130xl Genetic Analyzer
of forward and reverse primers (20 mM), 1 ml of Taq polymerase (Applied Biosystems, Warrington, UK). Terminal restriction
(Bioline) and 1 ml of soil DNA template. PCR conditions consisted fragments (TRFs) generated were analysed using GeneMapper
of an initial denaturation at 95 8C for 5 min followed by 30 cycles 3.7 (ABI, UK). Only TRFs at position between 25 and 500 bp were
of 95 8C for 30 s, 55 8C for 30 s, 72 8C for 60 s, and a final extension considered for further analysis in order to avoid fragments
at 72 8C for 10 min. Fungal DNA was PCR amplified using formed by the primer dimmers. The overall structure of
primers specific to fungal ITS. A forward ITS1F-FAM labelled (50 - bacterial and fungal communities among the different habitats
CTT GGT CAT TTA GAG GAA GTA A-30 ) (White et al., 1990) and a (soils) and treatments (the presence of plants) were compared
reverse ITS4 (50 -TCC TCC GCT TAT TGA TAT GC-30 ) (Gardes and by analysing binary (presence/absence) data using a multi-
Bruns, 1993) were used. The PCR mixture and thermocycling variate approach (principal coordinate analysis).
conditions were the same as described above. To check purity
and size of PCR amplicons, 5 ml of each reaction mix was run on 2.5. Cloning and sequencing
1% agarose gel (w/v). The PCR products were purified using
GeneEluteTM PCR clean-up kit (Sigma–Aldrich, Dorset, UK) Bacterial 16S rDNA gene amplicons from samples (pooled
following the protocol provided by the manufacturer. DNA from selected treatments) were obtained using the
242 applied soil ecology 41 (2009) 239–248

same set of primers (unlabelled) described in Section 2.3 and Soils from the Fo treatment had the highest amount of
were cloned in E. coli using TOPO cloning kit (Invitrogen, microbial biomass. A similar observation was made for soil
Paisley, UK). About 25 clones were selected for each sample respiration values, where the presence of plants had a highly
for sequencing. The correct insert size in each clone was significant effect on the respiration rate (P < 0.01). Again, the
checked by vector targeted PCR (primer M13 F and M13 R) and rate of respiration was lowest in the bare soil and highest in
gel electrophoresis. PCR amplicons were purified by ethanol soil from the Fo treatment, which was significantly different
precipitation. Sequencing was performed with Big Dye from all the other treatments (P < 0.05; Fig. 1b). When plant
Terminator cycle sequencing reaction kit (ABI, UK) using was used as a treatment (considering all plant treatments
vector specific T3 or T7 primers on an automated DNA together), there was no apparent effect of the presence of
sequencer (Prism 3130xl Genetic Analyser, ABI, UK). The plants on soil moisture. However, the moisture in the bare soil
quality of sequence was checked by the Sequence analysis treatment was comparatively lower than in the soil under Fo,
software (ABI) and approximately 500 bp of good quality Ns and Tr plants (P < 0.05; Fig. 1c). Biological and physico-
sequence was obtained for each clone. Analysed sequences chemical properties of the field-blocks are listed in Table 2.
were compared with 16S rDNA using FASTA-3 search of the There was certain variation between different field-blocks for
EMBL database (http://www.ebi.ac.uk/fasta33). Fasta3 along all measured parameters but they were not statistically
with ribosomal database of the Michigan State University significant.
(http://rdp8.cme.msu.edu/html/) and NCBI (http://
www.ncbi.nlm.nih.gov) were used to classify individual
sequences into different groups. Sequences were then
grouped into different phyla and their distribution was
presented in a histogram.

2.6. Statistical analysis

Treatment effects (plant or field-block) on soil N, C and


moisture and on biological (respiration and microbial bio-
mass) were analysed using analysis of variance (ANOVA) for
randomised blocks. Significant differences between treat-
ments were established using least significant difference (LSD)
values. The overall structure of the bacterial and fungal
communities in the different locations (field-blocks) and
treatments (plant species) was compared by analysing binary
data (presence/absence) using principal coordinates (PCO)
analysis and a Euclidian/Jaccard similarity matrix. Principal
coordinates analysis uses a matrix of similarities or dissim-
ilarities between the samples to produce a low-dimensional
representation of the data in such a way that the distances
between points are as close as possible to the original
dissimilarities. For each data set we carried out an ANOVA
on PCO scores for the first five dimensions as a randomised
complete block design. To identify overall pattern of commu-
nity distribution, PCO scores from the first 10 PCO dimensions
were used for canonical variate analysis (CVA). An ordination
diagram was obtained from the scores from CV axis 1 and CV
axis 2 to elucidate the effect of plant species or field-block on
soil microbial communities. To examine the correlation
between different abiotic and biotic factors and microbial
community pattern, regression analysis was carried out. All
statistical analyses were carried out using GenStat version 8
(VSN International Ltd.).

3. Results
Fig. 1 – Microbial biomass (a), respiration (b), moisture of
3.1. Effects of plant species on soil moisture, respiration, fresh soil (c) and C:N ratio (d) in soils with different plant
microbial biomass, soil N and soil C species. Error bars represent W standard errors. Treatment
acronyms are Agrostis capillaris (Ac), Anthoxanthum
The effect of the presence of plants on microbial biomass was odoratum (Ao), bare soil (B). Festuca ovina (Fo), Festuca rubra
highly significant (P < 0.001; Fig. 1a). Bare soil had the lowest (Fr), Nardus stricta (Ns), Lolium perenne (Lp), natural turf (N)
microbial biomass compared to all planted soils (P < 0.05). and Trifolium repens (Tr).
applied soil ecology 41 (2009) 239–248 243

Table 2 – Biological and physico-chemical properties of the field-blocks. Values in brackets represents standard errors of
replicates.
Unit Block 2 Block 3 Block 4 Block 6 Block 7

Soil N (%) 1.1 (0.09) 1.0 (0.06) 1.1 (0.03) 1.3 (0.08) 1.1 (0.08)
Soil C (%) 13.8 (1.1) 13.2 (0.64) 14.4 (0.4) 17.6 (1.2) 13.9 (0.93)
Moisture (%) 453.6 (2.6) 47.8 (0.99) 44.4 (3.2) 50.5 (1.9) 47.4 (3.93)
Microbial biomass mg/g of soil 1894.4 (315.9) 2093.8 (256.8) 2064.4 (152.3) 1475.4 (97.4) 1947.2 (434.7)
Respiration mg/g of soil 2.3 (0.28) 2.2 (0.16) 2.1 (0.16) 2.3 (0.23) 2.4 (0.35)

Regression analysis on these characteristics revealed an variation within replicates was large. For example, Fr was
interactive response. Moisture showed a slight positive separated from all other soils except Ac. Similarly, Tr was
correlation with respiration (P < 0.001, R2 = 0.27) and microbial different from all except N (Fig. 2c).
biomass (P < 0.05, R2 = 0.15). This relationship was strength- However, when the TRFLP data were analysed for the
ened when multiple regression was employed using plant type effects of field-blocks on community structure, all first three
as a grouping factor. The relationship between soil respiration PCO dimensions were significantly affected by the block
and microbial biomass was strong and statistically highly location (Table 3). Further analysis of the data by CVA
significant (P < 0.001, R2 = 0.55). When both moisture and suggested that the bacterial community of field-block 1 was
biomass were used as explanatory variates to identify impact different from all other field-blocks. Similarly field-blocks 3
on soil respiration, it gave a very high correlation value and 4 were different from field-blocks 6 and 7 (Fig. 2d).
(P < 0.001, R2 = 0.63). Similarly, moisture alone had compara- Regression analysis on the first five PCO dimensions, sug-
tively smaller impacts on microbial biomass and the relation- gested that bacterial community structure was not correlated
ship was weak (R2 = 0.15), however this relationship was with respiration, but the third dimension was significantly
substantially increased when regression was carried out with influenced by the moisture content (Table 3). There was a
plants as a grouping factor (P < 0.001, R2 = 0.40). The C:N ratio strong correlation between microbial biomass and bacterial
in soils planted with different species tended to be different, communities in PCO dimension 2 (P < 0.001). PCO dimension 4
although this effect was not statistically significant. of the bacterial community was significantly influenced by soil
N (P < 0.05) and soil C (P < 0.01) (Table 3).
3.2. TRFLP analysis of the bacterial community
3.3. Cloning and sequencing of the bacterial community
TRFLP profiles of the bacterial community were obtained for
all samples. On average between 20 (for the natural mixed Clone libraries for 16S rRNA genes were constructed for the
community) and 38 (for Ao) well-resolved TRFs were detected bacterial community from soil treatments B, Ac, Ao, N, Tr and
(Fig. 2a). A one-way ANOVA on TRFs had suggested that there Lp. For each library, 25 clones were selected randomly for
was no effect of plant species on the number of TRFs (P > 0.05). sequencing of approximately 500 bases. In all, 107 good
When the number of TRFs was analysed for presence in each quality partial sequences were obtained and analysed. Based
field-block, between 19 (for field-block 3) and 38 (for field-block on their FASTA 3, RDP II and NCBI database matches, all
7) TRFs were detected. ANOVA suggested that field-block had a sequences were grouped under 11 phyla; alpha-, beta-,
significant effect on the number of TRFs (P < 0.01). Further gamma- and delta-Proteobacteria, Firmicutes, Actinobac-
analysis of the data suggested that field-block 3 had teria, Acidobacteria, Gemmatimonadetes, Bacteroidetes,
significantly lower numbers of TRFs than all the other field- Chloroflexi, and unclassified domains. Gemmatimonadetes
blocks while field-blocks 2 and 4 had significantly lower and Bacteroidetes were present in only one library (N and Ao,
numbers of TRFs than field-block 7 (Fig. 2b). respectively). The distribution of other phyla of bacteria in
The TRFLP data from all samples were analysed by PCO each sample is presented in Fig. 3. All libraries contained
analysis. Scores of the first 10 dimensions were obtained. representative from Acidobacteria which varied from 14% (B
ANOVA on the first five PCO scores suggested that none of the soil) to 53% (Tr soil) of the total clones. Similarly alpha-
PCO dimensions were significantly influenced by plant species Proteobacteria were present in all samples and their relative
(Table 3). However, when the ANOVA was carried out proportions varied from 29% (in Ao soil) to 56% (in Lp soil).
including the location as a block effect, the bacterial commu- gamma-Proteobacteria were present in all libraries except Tr.
nity in the bare soil was significantly different from all other The Firmicutes were present only in the B soil but they
soils in PCO dimension 4 (data not shown). However there was constituted a substantial proportion of clones (14%) in this
no difference in the bacterial communities between soils with library. Actinobacteria were another major group which were
different plant species. PCO scores from the first 10 dimen- present in all samples except B and Lp libraries. A substantial
sions were further analysed by CVA, which confirmed that the proportion of the unclassified bacteria was also detected in B
soil bacterial community in the bare soil treatment was and Lp soils. Overall, the soil without any plants (B) was
different from all other soil treatments (Fig. 2c). However, CVA dominated by Bacilli, Clostridia (Firmicutes), alpha-Proteo-
analysis also suggested that there was a weak effect of plant bacteria and an unclassified group. It has the lowest number
species on soil bacterial communities. Here some soils, with of Acidobacteria in comparison with other libraries and was
particular plant species treatments were separated from all devoid of beta-Proteobacteria, Actinobacteria and Chlorflexi
other soils on CV 2, although the data were dispersed and groups. Soil samples with growing plants were dominated by
244 applied soil ecology 41 (2009) 239–248

Proteobacteria and Acidobacteria. However the difference in


clone distribution between plant species was smaller (Fig. 3).

3.4. TRFLP analysis of fungal communities

Like the bacterial community, TRFLP profiles for fungal


communities were obtained successfully from all soil sam-
ples. ANOVA suggested that there was no effect of plant
species on the number of TRFs (P > 0.05), which varied from 46
(for Fr) to 79 (for Lp) (Fig. 4a). However, the average TRFs for
different field-blocks varied from 46 (for field-block 2) to 71 (for
field-block 7) (Fig. 4b) and the effect of field-blocks on the
number of TRFs was significant (P < 0.05).
PCO analysis of fungal binary data of TRFs suggested that
the first three dimensions of the PCO scores were not
affected by plant species. However, the fourth dimension
was significantly influenced by plant species (P < 0.05)
(Table 4). In this dimension, fungal communities in B soil
were different from all other samples except Lm, Ns, Ra and
Tr (P < 0.05). However, in this dimension, the fungal
communities from Ac and N samples were also different
from Ra and Lm (P < 0.05), suggesting an impact of plant
species on the soil fungal community. Data were further
analysed by CV analysis on the PCO scores from the first 10
dimensions and an ordination diagram was obtained (Fig. 4c
and d) which confirmed that the fungal community in B soil
was different from all other samples. B soils were separated
from other samples on CVA axis 1 which accounted for 42%
of variability in the dataset. Fungal communities from Fo, Fr,
Lm, Lp and Tr treatments were separated from other plant
species treatments on axis 2 suggesting some impact of
plant species on fungal community structure in soil (Fig. 4c).
When the data were analysed for the impact of field-block
on the fungal communities, the first two dimensions of the
PCO score showed a significant correlation (P < 0.01 and
<0.001, respectively). This observation was further sup-
ported by the CVA analysis of the first 10 PCO dimensions
where fungal community from field-blocks 2, 3 and 4 were
separated on CVA axis 1. Field-blocks 6 and 7 were
separated from Field-block 2 on CVA axis 2, while from 3
and 4 on both axes 1 and 2 (Fig. 4d).
Regression analysis on PCO scores from the first five
dimensions suggested a comparatively weak but significant
correlation between soil respiration and fungal communities
in PCO dimension 5 and a strong correlation between
microbial biomass and fungal communities in PCO dimension
2 (P < 0.001). Further analysis of data suggested a strong
Fig. 2 – Number of bacterial terminal restriction fragments influence of soil N (P < 0.05) and soil C (P < 0.01) on fungal
(TRFs) obtained from soils (a) under different vegetation communities in PCO dimension 1 (Table 4).
(plant species) and (b) from different field-blocks.
Ordination plots of the first and second dimensions of 3.5. Interaction between bacterial and fungal
canonical variate analysis (CVA) scores for the soil communities
bacterial community obtained for different plant species
(c) and field-blocks (d). Error bars are Wstandard errors. To investigate the interactions between soil bacterial and
Numbers in parentheses are percentage variation fungal communities, regression analysis of the first five
explained by the CV. Treatment acronyms are Agrostis dimensions of bacterial communities was carried out
capillaris (Ac), Anthoxanthum odoratum (Ao), Festuca ovina against the first five dimensions of fungal communities.
(Fo), Festuca rubra (Fr), Nardus stricta (Ns), Luzula multiflora With the exception of dimension five, all dimensions of
(Lm), Trifolium repens (Tr), Rumex acetosa (Ra), Lolium bacterial communities showed a significant correlation
perenne (Lp), natural turf (N) and bare soil (B). between bacterial and fungal communities (P < 0.05). PCO
applied soil ecology 41 (2009) 239–248 245

Table 3 – Significance (P-values) of the effects of location (field-blocks), plant species and selected environmental variates
on the soil bacterial community. The number in parentheses under the PC dimension column represents % variation
explained by individual dimensions.
PC dimensions Plant Field-block Respiration Moisture Biomass Soil N Soil C

1 (29.3%) 0.72 0.001 0.22 0.19 0.12 0.09 0.16


2 (8.13%) 0.71 0.04 0.09 0.45 <0.001 0.75 0.93
3 (5.9%) 0.94 0.004 0.15 0.023 0.41 0.3 0.1
4 (4.9%) 0.12 0.059 0.199 0.084 0.33 0.03 0.005
5 (4.37%) 0.2 0.85 0.22 0.78 0.84 0.38 0.5

dimension one of the bacterial community was significantly 4. Discussion


correlated with PCO dimensions 1, 2 and 5 of the fungal
communities. PCO dimensions 2, 3, and 4 of bacterial Plants create positive feedbacks to nutrient cycling because of
communities were also correlated with PCO dimensions, 2, 4 species differences in carbon deposition and competition with
and 1 of the fungal communities. When PCO scores of microbes for nutrients. It has been argued that plant
dimension 2 of bacterial and fungal communities were physiological trait effects can be as or more important than
plotted against each other a significant correlation was abiotic factors such as climate, in controlling ecosystem
observed (P < 0.001) (Fig. 5). The degree of concordance functions (Verville et al., 1998). Physiological traits such as
between the two microbial communities was further relative growth rate (RGR) and specific leaf area (SLA) have
examined by Procrustean analysis of the first five dimen- been linked to the success of plant growth (Zou et al., 2007),
sions of fungal and bacterial communities. Procrustean evolutionary specialisation (Hunt and Cornelissen, 1997), and
analysis was chosen over the more commonly used to soil biota (Wardle et al., 1998). Data from this study revealed
Mantel test because it is more powerful in detecting that the presence of plants resulted in a significant increase in
matrix association (Peres-Neto and Jackson, 2001). A soil moisture content, microbial biomass and respiration. This
permutation procedure for Procrustean rotation (PROTEST; observation suggests that plant presence not only supports
Jackson, 1995) was used to evaluate the significance of the growth in soil microbial communities but also influences soil
association between fungal and bacterial data matrices abiotic properties. Although not presented in this manuscript,
obtained by T-RFLP. Here too, a significant correlation was we also found different plant species produced different
found (M1/2 = 0.39; P < 0.001) between bacterial and fungal quality (C:N ratio, dry weight/wet weight ratio) and quantity of
communities. When procrustean residuals were plotted plant biomass (shoot, root and litter). However, plant biomass
against soil moisture, N, and C no significant correlation did not correlate with microbial biomass, respiration or soil
was found. moisture content. Multivariate analysis of the microbial
community suggested that the bacterial communities are
highly influenced by sample location (field-block) and to a
lesser extent, by the presence of plants (Fig. 2). The plant
species-specific effect on the bacterial community was weak
(Fig. 2; Table 3). However, CVA analysis did suggest a weak
impact of a few plant species on soil microbial communities
after the 3 years of growth. Species-specific impacts of plants
on soil biota have been attributed to specific plant physiolo-
gical traits, independent of environmental factors. It has been
argued that plants with differential ecophysiological traits will
exert an impact on soil biota through differential rhizodeposi-
tion and quality of litters (Bardgett et al., 1999). However,
several other studies suggested that soil characteristics
(Girvan et al., 2003; Singh et al., 2006) and spatial distribution
(Nunan et al., 2005; Ritz et al., 2004) are more important factors
in shaping the microbial community structure. Our data
supports the later observation. Among the abiotic factors,
multivariate analysis of data suggested that the bacterial
community was influenced by soil moisture (dimension 3) and
Fig. 3 – Frequency distribution of phylogenetic taxa at the impact of soil C and N appeared only in dimension 4
class-level in clone libraries of amplified 16S rRNA genes (Table 3).
of bacteria recovered from bare soil and soils with The objective of cloning and sequencing data was to
different plant species. Treatment acronyms are Agrostis confirm our observation of TRFLP for the bacterial community
capillaris (Ac), Anthoxanthum odoratum (Ao), Festuca ovina and that is why only limited clones were sequenced for this
(Fo), Festuca rubra (Fr), Nardus stricta (Ns), Luzula multiflora study. These data support our TRFLP results in the sense that
(Lm), Trifolium repens (Tr), Rumex acetosa (Ra), Lolium the presence of plants causes change in the bacterial
perenne (Lp), natural turf (N) and bare soil (B). community but the plant species-specific effect on the
246 applied soil ecology 41 (2009) 239–248

bacterial community is not prominent. The results showed


that the bare soil was dominated by Firmicutes and Proteo-
bacteria, while soils with different plant species were mainly
dominated by the Acidobacteria and Proteobacteria. However,
this observation should be interpreted with caution as the
number of clones on which this finding was based was limited.
Nonetheless, agreeing with this, a few reports suggest that the
Proteobacteria and Acidobacteria dominate in rhizosphere
locations at the expense of other Gram-positive bacteria
(Singh et al., 2007b; Lu et al., 2006). In a study of the grassland
soil adjacent to the present site at Sourhope, Griffiths et al.
(2006) found the upper organic layer dominated by alpha-
Proteobacteria and the mineral layer by Acidobacteria. McCaig
et al. (1999) showed that clone libraries from the adjacent plot
of the same site comprised approximately 40% alpha-
Proteobacteria and 8% Acidobacteria. Our results are in
accordance with these findings.
As with the bacterial community, the fungal community
was strongly influenced by block position and only weakly
influenced by plant species. However, PCO dimension 4
showed a significant impact of plant species on fungal
community structure, suggesting a relatively stronger effect
in comparison to the bacterial community. Further analysis of
data by CVA suggested that soils with Fo, Tr, Lp and Ac
harboured distinctly different fungal communities than the
other soils. This may suggest that plant species with faster
growth rates have distinct effects on the fungal community in
comparison to slower growing plant species such as Ns. The
fungal community responded differently to abiotic factors as
compared to the bacterial community. Soil moisture did not
have any influence on the fungal community while soil N and
soil C had a comparatively stronger influence in shaping the
fungal community structure; PCO dimension 1 was signifi-
cantly co-related with these properties.
The most important finding of this work is the interaction
between the fungal and bacterial community structure.
Although fungi and bacteria have been studied extensively
in isolation in various habitats, few studies have examined
them simultaneously (Singh et al., 2007b; Costa et al., 2006),
and none of them attempted to investigate the interaction
between these two microbial communities. Bacteria and fungi
live in close association in the soil environment. Due to this
close proximity, fungal–bacterial interactions could be nega-
tive, positive or neutral (Boer et al., 2005). Our data suggest a
strong correlation between fungal and bacterial communities.
The degree of concordance between these two communities
Fig. 4 – Number of fungal TRFs obtained from soils (a) under was further investigated by the Procrustean analysis of PCO
different vegetation (plant species) and (b) from different scores. This was also done to factor out plant-specific
field-blocks. Ordination plots of the first and second influences. Our results suggest a significant concordance
dimensions of CVA scores for the soil fungal community between these two microbial communities (P < 0.001). The
obtained for different plant species (c) and field-blocks (d). procrustean residuals were not correlated with any soil abiotic
Error bars are Wstandard errors. Numbers in parentheses properties, suggesting that the relationship between bacterial
are percentage variation explained by the CV. Treatment and fungal communities was independent of soil moisture
acronyms are Agrostis capillaris (Ac), Anthoxanthum content, pH, C and N. However, our results should be
odoratum (Ao), Trifolium repens (Tr), Lolium perenne (Lp), interpreted with caution as the present observation was
natural turf (N) and bare soil (B). made for just one sampling regime. Nonetheless, the relation-
ships between AM fungi and bacteria have been reported
previously. In a laboratory study, a single AM fungus, Glomus
mosseae, was reported to have a direct impact in shaping the
bacterial community found on plant roots (Artursson et al.,
applied soil ecology 41 (2009) 239–248 247

Table 4 – Significance (P-values) of the effects of location (field-blocks), plant species and selected environmental variates
on the soil fungal community. The number in parentheses under the PC dimension column represents % variation
explained by individual dimensions.
PC dimensions Plant Field-block Respiration Moisture Biomass Soil N Soil C

1 (12.3%) 0.42 0.005 0.4 0.83 0.31 0.002 0.01


2 (6.9%) 0.17 <0.001 0.16 0.75 <0.001 0.67 0.34
3 (4.9%) 0.85 0.35 0.28 0.34 0.62 0.92 0.64
4 (4.4%) 0.03 0.5 0.31 0.16 0.7 0.83 0.89
5 (3.7%) 0.76 0.07 0.04 0.11 0.09 0.4 0.31

metagenome due to Glomus mosseae inoculation or plant


species. Environ. Microbiol. 7, 1952–1966.
Bardgett, R.D., Mawdsley, J.L., Edwards, S., Hobbs, P.J., Rodwell,
J.S., Davies, W.J., 1999. Plant species and nitrogen effects on
soil biological properties of temperate upland grasslands.
Funct. Ecol. 13, 650–660.
Boer, E.D., Folman, L.B., Summerbell, R.C., Boddy, L., 2005. Living
in a fungal world: impact of fungi on soil bacterial niche
development. FEMS Microbiol. Rev. 29, 795–811.
Brookes, P.C., Landman, A., Pruden, G., Jenkinson, D.S., 1985.
Chloroform fumigation and release of soil nitrogen—a rapid
direct extraction method to measure microbial biomass
nitrogen in soil. Soil Biol. Biochem. 17, 837–842.
Buckley, D.H., Schmidt, T.M., 2001. The structure of microbial
communities in soil and the lasting impact of cultivation.
Microbial Ecol. 42, 11–21.
Fig. 5 – Relationship between fungal and bacterial Costa, R., Gotz, M., Mrotzek, N., Lottmann, J., Berg, G., Smalla, K.,
communities based on principal coordinate analysis 2006. Effects of site and plant species on rhizosphere
dimension 2 (PCO-2) scores associated with all field plots. community structure as revealed by molecular analysis of
microbial guilds. FEMS Microbiol. Ecol. 56, 236–249.
Gardes, M., Bruns, T.D., 1993. ITS primers with enhanced
specificity for basidiomycetes: application to the
2005). Later, Singh et al. (2008) reported that that AM fungal
identification of mycorrhiza and rusts. Mol. Ecol. 2, 118–133.
community influenced the bacterial community structure on Garnier, E., Cortez, J., Billes, G., Navas, M., Roumet, C.,
grass rhizoplanes. However, this study is the first time a Debussche, M., Laurent, G., Blanchard, A., Aubry, D.,
concordance between fungi and bacteria at the community Bellman, A., Niell, C., Toussant, J., 2004. Plant functional
level and field scale has been shown. This observation markers capture ecosystem properties during secondary
suggests that a complex multi-trophic interaction occurs succession. Ecology 85, 2630–2637.
Girvan, M.S., Bullimore, J., Pretty, J.N., Osborn, A.M., Ball, A.S.,
between above and belowground communities and the
2003. Soil type is the primary determinant of the
microbial community structures are shaped both by several
composition of the total and active bacterial communities
abiotic and biotic interacting factors. in arable soils. Appl. Environ. Microbiol. 69, 1800–1809.
Grayston, S.J., Wang, S.Q., Campbell, C.D., Edwards, A.C., 1998.
Selective influence of plant species on microbial diversity in
Acknowledgements the rhizosphere. Soil Biol. Biochem. 30, 369–378.
Grayston, S.J., Campbell, C.D., Bardgett, R.D., Mawdsley, J.L.,
Clegg, C., Ritz, K., Griffiths, B.S., Rodwell, J.S., Edwards, S.J.,
We thank Dr Naoise Nunan (CNR, France) for help in statistical
Davies, W.J., Elston, D.J., Millard, P., 2004. Assessing shifts in
analysis, Jasmine Ross, Eileen Reid, Ruth Primrose, Brian Ord, microbial community structure across a range of grasslands
Stacey Munro for various analyses and Dr Dave Johnson of differing management intensity using CLPP PLFA and
(University of Aberdeen) for critical reviewing of the manu- community DNA techniques. Appl. Soil Ecol. 25, 2563–2584.
script. This project was funded by the Scottish Executive Griffiths, R.I., Bailey, M.J., McNamara, N.P., Whiteley, A.S., 2006.
Environment and Rural Affairs Department. The functions and components of the Sourhope soil
microbiota. Appl. Soil Ecol. 33, 114–126.
Grime, J.P., Hunt, R., 1975. Relative growth rate—its range and
adaptive significance in a local flora. J. Ecol. 63, 393–422.
references
Grime, J.P., Hodgson, J.G., Hunt, R., 2007. Comparative Plant
Ecology: A Functional Approach to Common British Species.
Castlepoint Press, UK.
Aber, J.D., Mellilo, J.M., MacClaugherty, C.A., 1990. Predicting Hobbie, S.E., 2002. Effects of plant species on nutrient cycling.
long-term patterns of mass loss, N dynamics and soil Trend Ecol. Evol. 7, 336–339.
organic matter formation from initial fine litter chemistry in Horner-Devine, M.C., Leibold, M.A., Smith, V.H., Bohannan,
temperate forest ecosystems. Can. J. Bot. 68, 2202–2208. B.J.M., 2003. Bacterial diversity patterns along a gradient of
Artursson, V., Finlay, R.D., Jansson, J.K., 2005. Combined primary productivity. Ecol. Lett. 6, 613–622.
bromodeoxyuridine immunocapture and terminal- Hunt, R., Cornelissen, J.H.C., 1997. Components of relative
restriction fragment length polymorphism analysis growth rate and their interrelation in 59 temperate plant
highlights differences in the active soil bacterial species. New Phytol. 135, 395–417.
248 applied soil ecology 41 (2009) 239–248

International Standard Organisation, 1997. Soil Quality— Singh, B.K., Munro, S., Potts, J., Millard, P., 2007b. Influence of
Determination of Water Content on a Volume Basis— grass species and soil type on rhizosphere microbial
Gravimetric Method. ISO 11461. ISO, Geneva. community structure in grassland soils. Appl. Soil Ecol. 36,
Jackson, D.A., 1995. PROTEST—a procrustean randomization 147–155.
test of community environment concordance. Ecoscience 2, Singh, B.K., Nunan, N., Ridgway, K.P., McNicol, J., Young, J.P.W.,
297–303. Daniell, T.J., Prosser, J.I., Millard, P., 2008. Relationship
Johnson, D., Booth, R.E., Whiteley, A.S., Bailey, M.J., Read, D.J., between assemblages of mycorrhizal fungi and bacteria on
Grime, J.P., Leak, J.R., 2003. Plant community composition grass roots. Environ. Microbiol. 10, 534–541.
affected the biomass, activity and diversity of soil Smit, E., Leeflang, P., Glandorf, B., van Elsas, J.D.V., Wernars, K.,
microorganisms in reconstituted calcareous grassland. Eur. 1999. Analysis of fungal diversity in the wheat rhizosphere
J. Soil Sci. 54, 671–678. by sequencing of cloned PCR-amplified gene encoding 18S
Jones, D.L., Hodge, A., Kuzyakov, Y., 2004. Plant and mycorrhizal rRNA and temperature gradient gel electrophoresis. Appl.
regulation of hizodeposition. New Phytol. 163, 459–480. Environ. Microbiol. 65, 2614–2621.
Lu, Y., Rosencrantz, D., Liesack, W., Conrad, R., 2006. Structure Sparling, G.P., 1981. Microcalorimetry and other methods to
and activity of bacterial community inhabiting rice roots assess biomass and activity in soils. Soil Biol. Biochem. 13,
and the rhizosphere. Environ. Microbiol. 8, 1351–1360. 93–98.
Marcial-Gomes, N.C., Fagbola, O., Costa, R., Rumjanek, N.G., Thompson, K., Parkinson, J.A., Band, S.R., Spencer, R.E., 1997. A
Buchner, A., Mendona-Hagler, L., Smalla, K., 2003. comparative study of leaf nutrient concentration in a
Dynamics of fungal communities in bulk and maize regional herbaceous flora. New Phytol. 136, 679–689.
rhizosphere soil in the tropics. Appl. Environ. Microbiol. 69, Tilman, D., Reich, P.B., Knops, J.M.H., 2006. Biodiversity and
3758–3766. ecosystem stability in a decade-long grassland experiment.
Marilley, L., Vogt, G., Blanc, M., Aragno, M., 1998. Bacterial Nature 441, 629–632.
diversity in the bulk soil and rhizosphere of Lolium perenne Usher, M.B., Sier, A.R.J., Hornung, M., Millard, P., 2006.
and Trifolium repens as revealed by PCR restriction analysis Understanding biological diversity in soil: The UK’s Soil
of 16S rDNA. Plant Soil 198, 219–224. Biodiversity Research Programme. Appl. Soil Ecol. 33,
McCaig, A.E., Glover, L.A., Prosser, J.I., 1999. Molecular analysis 101–113.
of bacterial community structure and diversity in Vandenkoornhuyse, P., Ridgway, K.P., Watson, I.J., Fitter, A.H.,
unimproved and improved upland grassland pastures. Young, J.P.W., 2003. Co-existing grass species have
Appl. Environ. Microbiol. 65, 1721–1730. distinctive arbuscular mycorrhizal communities. Mol. Ecol.
Mougel, C., Offre, P., Ranjard, L., Corbeand, T., Gamalero, E., 12, 3085–3095.
Robin, C., Lemanceau, P., 2006. Dynamics of the genetic Van der Heijden, M.G.A., Klironomos, J.N., Ursic, M., Moutoglis,
structure of bacterial and fungal communities at different P., Streitwolf-Engel, R., Boller, T., Wiemken, A., Sanders, I.R.,
developmental stages of Medicago truncatula 1998a. Mycorrhizal fungal diversity determines plant
Gaertn.cv.jemalong line J5. New Phytol. 170, 165–175. biodiversity, ecosystem variability and productivity. Nature
Nunan, N., Daniell, T.J., Singh, B.K., Papert, A., McNicol, J.W., 396, 69–72.
Prosser, J.I., 2005. Links between plant and rhizoplane Van der Heijden, M.G.A., Boller, T., Wiemken, A., Sanders, I.R.,
bacterial communities in grassland soils characterized 1998b. Different arbuscular mycorrhizal fungal species are
using molecular techniques. Appl. Environ. Microbiol. 71, potential determinants of plant community structure.
6784–6792. Ecology 79, 2082–2091.
Nusslein, K., Tiedje, J.M., 1999. Soil bacterial community shift Van der Heijden, M.G.A., Bakker, R., Verwaal, J., Scheublin, T.R.,
correlated with change from forest to pasture vegetation in Rutten, M., van Logtestijn, R., Staehelin, C., 2006. Symbiotic
a tropical soil. Appl. Environ. Microbiol. 65, 3622–3626. bacteria as a determinant of plant community structure and
Peres-Neto, P.R., Jackson, D.A., 2001. How well do multivariate plant productivity in dune grassland. FEMS Microbiol. Ecol.
data sets match? The advantages of a Procrustean 56, 178–187.
superimposition approach over the Mantel test. Oecologia Van Vuuren, M.M.I., Aerts, R.F., Berendse, F., De Visser, W., 1992.
129, 169–178. N mineralization in heathland ecosystems dominated by
Phillips, D.A., Ferris, H., Cook, D.R., Strong, D.R., 2003. Molecular different plant species. Biogeochemistry 16, 151–166.
control points in rhizosphere food webs. Ecology 84, Verville, J.H., Hobbie, S.E., Chapin, F.S., hooper, D.U., 1998.
816–826. Response of tundra Ch4 and Co2 flux to manipulation of
Ritz, K., McNicol, W., Nunan, N., Grayston, S., Millard, P., temperature and vegetation. Biogeochemistry 41, 215–235.
Atkinson, D., Gollotte, A., Habeshaw, D., Boag, B., Clegg, Wardle, D.A., Barker, G.M., Bonner, K.I., Nicholson, K.S., 1998.
C.D., Griffiths, B.S., Wheatley, R.E., Glover, L.A., McCaig, A.E., Can comparative approaches based on plant
Prosser, J.I., 2004. Spatial structure in soil chemical and ecophysiological traits predict the nature of biotic
microbiological properties in an upland grassland. FEMS interactions and individual plant species effects in
Microbiol. Ecol. 49, 191–205. ecosystems? J. Ecol. 86, 405–420.
Singh, B.K., Millard, P., Whiteley, A.S., Murrell, J.C., 2004. Wardle, D.A., Bardgett, R.D., Klironomos, J.N., Setala, H., van der
Unravelling rhizosphere–microbial interactions: Putten, W.H., Wall, D.H., 2002. Ecological linkages between
opportunities and limitations. Trend Microbiol. 12, 386–393. aboveground and belowground biota. Science 304, 1629–
Singh, B.K., Nazarie, L., Munro, S., Anderson, I., Campbell, C.D., 1633.
2006. Use of multiplex terminal restriction fragment length White, T.J., Bruns, T.D., Lee, S., Taylor, J., 1990. Analysis of
polymorphism for rapid and simultaneous analysis of phylogenetic relationship by amplification and direct
different components of the soil microbial community. sequencing of ribosomal RNA genes. In: Innis, M.A., Gelfond,
Appl. Environ. Microbiol. 72, 7278–7285. D.H., Sainsky, J.J., White, T.J. (Eds.), PCR Protocol: A Guide to
Singh, B.K., Tate, K.R., Kapadia, G., Hedley, C.B., Macdonald, Method and Applications. Academic Press, New York, NY.
C.A., Millard, P., Murrell, J.C., 2007a. Effect of afforestation Zou, J., Roger, W.E., Siemann, E., 2007. Differences in
and reforestation of pastures on the activity and population morphological and physiological traits between native and
dynamics of methanotrophic bacteria. Appl. Environ. invasive populations of Sapium sebiferum. Funct. Ecol. 21,
Microbiol. 73, 5153–5161. 721–730.

You might also like