You are on page 1of 14

Environmental Microbiology (2007) 9(2), 308–321 doi:10.1111/j.1462-2920.2006.01139.

X-ray absorption spectroscopy (XAS) of toxic metal


mineral transformations by fungi

Marina Fomina,1 John Charnock,2 Andrew D. Bowen1 et al., 2003). In terms of metal pollution, both processes of
and Geoffrey M. Gadd1* metal mobilization (e.g. in situ soil washing, extraction and
1
Division of Environmental and Applied Biology, filtration techniques) and immobilization (e.g. in situ sta-
Biological Sciences Institute, School of Life Sciences, bilization techniques) may be applied to remediate con-
University of Dundee, Dundee, DD1 4HN, UK. taminated matrices (Cotter-Howells and Caporn, 1996;
2
SRS Daresbury Laboratory, Daresbury, Warrington, Knox et al., 2000; Prasad and De Oliveira Freitas, 2003;
WA4 4AD, UK. Brown et al., 2004). Fungi are a major component of the
biota in soils and mineral substrates, and can be very
tolerant to toxic metals: under certain environmental con-
Summary
ditions (e.g. low pH, pronounced toxic metal pollution)
Fungi can be highly efficient biogeochemical agents they can become the dominant microbial group (Fomina
and accumulators of soluble and particulate forms of et al., 2005a). The involvement of fungi in mutualistic
metals. This work aims to understand some of the symbiotic associations with plants (mycorrhizas), algae
physico-chemical mechanisms involved in toxic and cyanobacteria (lichens) has major consequences for
metal transformations focusing on the speciation of the biogeochemical cycling of elements (Gorbushina
metals accumulated by fungi and mycorrhizal et al., 1993; Smith and Read, 1997; Meharg and Cairney,
associations. The amorphous state or poor crystallin- 2000; Sterflinger, 2000; Gadd et al., 2005). Free-living
ity of metal complexes within biomass and relatively and mycorrhizal fungi can be efficient biogeochemical
low metal concentrations make the determination of agents and bioaccumulators of soluble and particulate
metal speciation in biological systems a challenging forms of metals (Gadd, 1993; 2004; Burford et al., 2003;
problem but this can be overcome by using Fomina et al., 2004; 2005a,b). More than 95% of all land
synchrotron-based element-specific X-ray absorption plants depend on symbiotic mycorrhizal fungi (Smith and
spectroscopy (XAS) techniques. In this research, we Read, 1997) which make fungal biogeochemical activity
have exposed fungi and ectomycorrhizas to a variety of particular importance in any type of remediation of toxic
of copper-, zinc- and lead-containing minerals. X-ray metal polluted soils, whether physico-chemical or
absorption spectroscopy studies revealed that biological. Mycorrhizal fungi solubilize phosphate and
oxygen ligands (phosphate, carboxylate) played a essential metals for the host plant in the course of ‘het-
major role in toxic metal coordination within the erotrophic leaching’ as a result of protonation (acidolysis),
fungal and ectomycorrhizal biomass during the accu- chelation (complexolysis) and metal accumulation by the
mulation of mobilized toxic metals. Coordination of biomass (Burgstaller and Schinner, 1993; Burford et al.,
toxic metals within biomass depended on the fungal 2003). Fungal and plant cell walls can act as a cation
species, initial mineral composition, the nitrogen exchanger due to their negative charge originating from
source, and the physiological state/age of the fungal functional groups, e.g. carboxylic, phosphate, amine or
mycelium. sulfhydryl, in different wall components (hemicelluloses,
pectin, lignin, chitin, etc.) (Gadd, 1993; Sarret et al.,
1998a; 1999). Mechanisms for metal immobilization within
Introduction
plant and fungal biomass also include intracellular uptake
The influence of microbiological processes on mineral with complexation to ligands such as S-containing pep-
transformations and metal speciation is of economic and tides (metallothioneins, phytochelatins), carboxylic acids
environmental significance. Certain processes solubilize (citrate, malate, oxalate) and phenolic acids (Gadd, 1993;
metals from minerals and bound locations thereby Sarret et al., 1998b; 2002; 2003; Fomina et al., 2004;
increasing metal bioavailability, whereas others immobi- 2005c; Bellion et al., 2006). Because of the amorphous
lize them and reduce bioavailability (Gadd, 1993; Burford state or poor crystallinity of metal complexes within
biomass and relatively low metal concentrations, the
Received 21 March, 2006; accepted 8 August, 2006. *For
correspondence. E-mail g.m.gadd@dundee.ac.uk; Tel. determination of metal speciation in biological systems
(+44) 1382 384765; Fax (+44) 1382 388216. remains a challenging problem with only a few studies on
© 2006 The Authors
Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd
Toxic metal speciation in fungi 309

Fig. 1. Diagram illustrating X-ray absorption spectroscopy.


A. The Cu K-edge X-ray absorption spectrum of copper phosphate.
B. The Cu K-edge EXAFS spectrum of copper phosphate. The oscillations are displayed as a function of k, the wave-vector, and the
amplitude (normalized to the size of the edge-step) is multiplied by k3, in order to enhance the features at high k (high energy).
C. The Fourier transform of the EXAFS spectrum of copper phosphate. The major peak is due to the four nearest neighbour oxygen atoms
around the copper, at a distance of 1.95 Å. The smaller peak is due to a combination of two copper atoms at 2.98 Å and three phosphorus
atoms at 3.15 Å.

fungal biomass that mainly clarify the nature of adsorption (scatterers). Interactions between the outgoing and scat-
sites on cell walls (Fourest et al., 1996; Sarret et al., tered electron waves give rise to the oscillations in the
1998a,b; 1999). However, synchrotron-based X-ray spectrum above the edge, which are the EXAFS
absorption spectroscopy (XAS) provides a means for (extended X-ray absorption fine structure) (Fig. 1A). In
studying element complexation in environmental samples certain situations, such as a highly symmetrical octahe-
varying from biological to mineralogical in nature (Sarret dral coordination site, the electron wave can interact with
et al., 1998a,b; 1999; 2002; 2003; Kemner et al., 2005). more than one scatterer, giving rise to multiple scattering
X-ray absorption spectroscopy is a technique increasingly effects.
used to study biological systems. The theory is well- The EXAFS data are isolated from the absorption spec-
documented (Hasnain, 1988; Charnock, 1995) and analy- trum by subtracting the ionization energy, E0, removal of
sis of the data gives information about the oxidation state a smooth background, and normalization of the oscilla-
of the target element and its coordination environment, tions to the edge step (the difference in absorbance just
including the number and identity of neighbouring atoms before and just after the edge) (Fig. 1B). The data are
(Gardea-Torresdey et al., 2004). It allows studies involv- displayed as a function of the wave-vector k (proportional
ing fast data collection, small samples, low concentra- to the square root of the difference between the X-ray
tions, crystalline and amorphous solids, and solutions. energy and E0). In k-space each scatterer contributes a
X-ray absorption spectroscopy is also a non-destructive, damped sine wave to the total EXAFS, with a frequency
non-invasive method that could probe metal transforma- which depends on the absorber–scatterer distance and
tions at the mineral–microbe interface, and directly study an intensity which depends on the atomic number of the
samples in their natural hydrated states. The basic scatterer. The damping is affected by thermal oscillations
process is absorption of an X-ray photon by the atom of along the absorber–scatterer axis. Thus, the observed
interest (the absorber atom) in the sample. The spectrum, EXAFS spectrum is an interference pattern of the contri-
plotting absorbance of the sample as a function of photon butions from all the surrounding scatterers (Fig. 1B).
energy (Fig. 1), shows a sudden increase, the absorption Taking a Fourier transform of the EXAFS spectrum gives
edge, corresponding to the ionization of a 1s (Cu or Zn, peaks corresponding to the different frequencies, yielding
K-edge) or 2p [Pb, L(III)-edge] electron, the energy of an approximate radial distribution function showing
which is sensitive to the oxidation state of the absorber ‘shells’ of scatterers (i.e. groups of scatterers at a similar
(Fig. 1A). When the X-ray is absorbed the photoelectron distance) around the absorber atom (Fig. 1C). This can be
is emitted in the form of a simple spherical wave. If the used to build up a model of distribution of scatterers
absorber atom is isolated the electron wave propagates surrounding the absorber. The EXAFS spectrum is analy-
freely outwards, but if the atom is incorporated in a sed by calculating a spectrum based on the model of the
molecule or dense matter the outgoing spherical wave absorber coordination site using various possible scatter-
is elastically scattered by the surrounding atoms ers and refining parameters in the model (distances,

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
310 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd

damping factors) to minimize a least squares residual, of grain size in the sample or a slight degree of non-
giving the best fit to the experimental data. crystallinity. Copper precipitation on fungal hyphae is not
This XAS study of toxic metal speciation within fungal homogeneous, and related studies have shown that well-
and ectomycorrhizal biomass not only aims to further structured moolooite (copper oxalate) crystals (10–
understand the physico-chemical mechanisms involved in 40 mm) on the mycelium comprised approximately 35% of
toxic metal transformations and accumulation by fungi the total copper accumulated (Fomina et al., 2005c).
and mycorrhizal associations, but also to demonstrate Various electron microscopic techniques (including wet
that XAS is an ideal approach for investigating metal mode Environmental SEM) coupled with elemental analy-
transformations at the mineral–microbe interface in bio- sis EDX have shown that the majority of copper accumu-
geochemical systems (Gardea-Torresdey et al., 2004; lated by B. caledonica was deposited extracellularly,
Kemner et al., 2005). being biomineralized within a well-hydrated mucilaginous
sheath of extracellular polymeric substances (EPS) sur-
rounding hyphae and hyphal cords (Fig. 2A).
Results and discussion The effect of nitrogen source on the coordination of
accumulated copper was examined for strains of the ecto-
Copper
mycorrhizal fungus R. rubescens in both axenic culture
Copper makes a significant contribution to global pollu- and in mycorrhizal association with Scots pine (Table 1,
tion (Nriagu, 1996). Although only a small fraction of Fig. 3G and H). The nitrogen source is a factor that can
copper content in sludge-amended soils is soluble, this is affect organic acid production by fungi with nitrate consid-
found almost exclusively in low molecular weight com- erably increasing oxalate excretion (Lapeyrie et al., 1991;
plexes with amino acids, small peptides or polycarboxylic Gharieb and Gadd, 1999; Whitelaw et al., 1999). Analysis
acids. This fraction is a major contributor to copper of the EXAFS spectrum from biomass of R. rubescens
mobility and bioavailability threatening lower trophic grown on nitrate and copper phosphate gave the best fit to
levels of the food chain (Vulkan et al., 2002). In situ the inner shell with a coordination shell of four oxygens
EXAFS and X-ray absorption near edge structure bound to the copper at a distance of 1.96 Å and two
(XANES) spectroscopy of copper speciation in electroki- oxygens bound at 2.31 Å. This is a typical distortion from
netically remediated metal contaminated soil revealed perfect octahedral symmetry for a six-coordinate copper
that 50% of the copper was chelated by humic sub- (II) centre, and is known as a Jahn-Teller distortion
stances, 28% formed CuCO3 and 22% formed Cu2O and (Greenwood and Earnshaw, 1984). The EXAFS contrib-
CuO (Liu and Wang, 2004). Fungi can play a significant uting to two further peaks in the Fourier transform were
role in copper transformations (Fomina et al., 2005a,b,c) best fitted with four carbons at 2.72 Å and four coppers at
and many soil fungi were able to withstand copper tox- 3.82 Å, showing carboxylate coordination close to oxalate
icity in heavily contaminated soils (500–11 500 mg Cu (Table 1, Figs 3H and 4A). If ammonium was used as a
per kg soil) (Fomina et al., 2005a). Nearly 60–70% of nitrogen source, the best fit to the inner shell was with a
tested ericoid and ectomycorrhizal fungal cultures were Jahn-Teller distorted coordination shell of four oxygens at
able to grow in the presence of copper phosphate and 1.94 Å and two oxygens at 2.43 Å. There was a small
cuprite with more than half of them being able to solubi- outer shell which was best fitted with two coppers at
lize copper phosphate and over 40% of them solubilizing 3.79 Å. However, attempts to fit phosphorus atoms at
cuprite (Fomina et al., 2005b). 3.1 Å (phosphate) or carbon atoms at 2.7 Å (oxalate/
In the present study, all tested samples of fungal and carboxylic acid) did not improve the fit (Table 1, Fig. 3G).
ectomycorrhizal biomass (Beauveria caledonica, Rhizo- For other biomass from organisms grown on ammonium-
pogon rubescens and R. rubescens/Scots pine) grown in containing medium, the fit with carbons was slightly better
the presence of copper phosphate, and Aspergillus niger and more consistent with oxalate coordination than with
grown on CuO, showed Cu coordination with oxygen phosphate, but was not conclusive probably indicating
ligands, fitting carboxylate and/or phosphate coordination mixed carboxylate/phosphate coordination of copper.
(Table 1). B. caledonica, grown on copper phosphate- Phosphate coordination of copper was observed for
containing medium, accumulated approximately A. niger grown on CuO (Table 1). The best fit to the inner
100 mg Cu g-1 dry weight (Fomina et al., 2005c). The coordination sphere was with four oxygens at 1.94 Å and
EXAFS spectrum and parameters, and the Fourier trans- two oxygens at 2.48 Å, typical of six-coordinate octahe-
form were very similar to those of the copper oxalate dral Cu(II) with a Jahn-Teller distortion which was
standard (Table 1) and the spectrum and the three clear improved by the addition of 2 phosphorus scatterers at
peaks in the Fourier transform could be fitted effectively 3.00 Å.
with the same parameters. In the standard, the outer In the microcosm experiments with A. niger,
shells were slightly better defined which may be an effect B. caledonica and Penicillium simplicissimum grown on

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 311
Table 1. Cu K-edge EXAFS parameters.

Sample Scatterers r (Å) 2s2 (Å2) Residual

Cu(acetate) standard 4¥O 1.95 0.012 25.2


1 ¥ Cu 2.60 0.007
4¥C 2.83 0.017
Cu(gluconate) standard 4¥O 1.93 0.009 25.1
2¥O 2.30 0.024
4¥C 2.73 0.037
4 ¥ Cu 3.78 0.027
Cu(malate) standard 4¥O 1.92 0.011 34.8
2¥O 2.38 0.020
4¥C 2.75 0.017
Cu2O standard 2¥O 1.85 0.012 29.3
12 ¥ Cu 3.03 0.024
6 ¥ Cu 4.27 0.029
24 ¥ Cu 5.27 0.035
12 ¥ Cu 6.06 0.025
Cu(oxalate) standard 4¥O 1.96 0.011 27.8
2¥O 2.35 0.024
4¥C 2.66 0.010
4 ¥ Cu 3.73 0.019
4 ¥ Cu 4.48 0.026
4 ¥ Cu 5.52 0.014
Cu3(PO4)2 standard 4¥O 1.95 0.011 25.3
2 ¥ Cu 2.98 0.020
3¥P 3.15 0.041
B. caledonica on copper phosphate-containing MMN agar 4¥O 1.99 0.021 41.0
2¥O 2.39 0.028
4¥C 2.70 0.013
4 ¥ Cu 3.74 0.025
4 ¥ Cu 4.00 0.039
B. caledonica on azurite-containing malt extract agar 3¥S 2.20 0.027 36.4
2 ¥ Cu 2.60 0.030
B. caledonica on malachite-containing malt extract agar 3¥S 2.16 0.025 37.6
2 ¥ Cu 2.58 0.031
A. niger on azurite-containing malt extract agar (margin) 3¥S 2.23 0.033 33.6
2 ¥ Cu 2.60 0.040
A. niger on azurite-containing malt extract agar (centre) 4¥O 1.95 0.022 34.4
2¥O 2.53 0.022
2¥P 2.89 0.018
2 ¥ Cu 3.70 0.018
A. niger on malachite-containing malt extract agar (margin) 3¥S 2.26 0.017 37.5
2 ¥ Cu 2.68 0.028
A. niger on malachite-containing malt extract agar (centre) 4¥O 1.94 0.032 31.9
2¥O 2.48 0.022
2¥P 3.00 0.046
A. niger on CuO-containing AP1 agar 4¥O 1.93 0.028 30.6
2¥O 2.48 0.035
2¥P 2.97 0.039
2 ¥ Cu 3.75 0.039
P. simplicissimum on azurite-containing malt extract agar 3¥S 2.14 0.022 45.4
1 ¥ Cu 2.56 0.018
P. simplicissimum on malachite-containing malt extract agar 3¥S 2.13 0.033 35.2
1 ¥ Cu 2.58 0.040
R. rubescens axenic culture grown on copper phosphate 4¥O 1.96 0.012 29.8
and nitrate MMN agar medium 2¥O 2.31 0.022
4¥C 2.72 0.019
2 ¥ Cu 3.82 0.039
R. rubescens axenic culture grown on copper phosphate 4¥O 1.94 0.013 27.1
and ammonium MMN agar medium 2¥O 2.37 0.037
R. rubescens/P. sylvestris ectomycorrhizas grown in a 4¥O 1.99 0.021 27.2
square Petri dish mesocosm with copper phosphate 2¥O 2.39 0.028
and nitrate Ingestad agar medium 4¥C 2.78 0.006
2 ¥ Cu 3.79 0.028
R. rubescens/P. sylvestris ectomycorrhizas grown in a 4¥O 1.94 0.012 33.7
square Petri dish mesocosm with copper phosphate 2¥O 2.42 0.054
and Ingestad agar medium 2 ¥ Cu 3.79 0.034

r is the copper–scatterer distance in Angstroms, ⫾0.02 Å inner shells, ⫾0.05 Å outer shells; 2s2 is the Debye–Waller type factor, ⫾15% inner shells, ⫾30%
outer shells (a measure of both the thermal motion between the absorber and scatterer and of the static disorder or range of absorber–scatterer distances);
the residual is a least squares residual from fitting the spectrum of the model to the experimental data.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
312 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd

Fig. 2. Electron microscopic observations of toxic metal localization in fungal biomass. (A) High-vacuum mode ESEM image of air-dried
Au/Pd-coated sample of extracellular moolooite (white arrows) precipitation by B. caledonica hyphae and cords when grown on copper
phosphate agar medium. (B) Low-vacuum (auxiliary gas) mode ESEM (uncoated samples) and (C) TEM images of zinc-containing crystals
(white arrows) found within cells of the extraradical mycelium of the R. rubescens/P. sylvestris ectomycorrhizal association grown in a
mesocosm on Ingestad agar with zinc phosphate and nitrate. Scale bars are (A) 20 mm, (B) 10 mm, (C) 1 mm.

agar medium containing azurite or malachite, sulfur coor- tion (Jaeckel et al., 2005; Kumar et al., 2005). Thus, it is
dination of copper accumulated by biomass was found in likely that the majority of copper sequestration in young
most samples (Table 1). A typical example of the best fit to mycelia occurs intracellularly by this mechanism. Proteins
the inner coordination sphere was with three sulfurs at are not ideal for long-term sequestration of copper and we
2.26 Å which was significantly improved by the addition of propose that there is a shift to a more permanent mode of
two copper scatterers at 2.68 Å clearly indicating the for- storage reflected in the phosphate coordination in the
mation of a copper sulfide phase. mature parts of the colony. The possible activity of acid
The age and therefore the physiological and reproduc- phosphatases, which have been linked to toxic metal
tive state of the fungal mycelium was found to play a resistance in fungi, increases with increased copper con-
crucial role in metal coordination within the biomass. On centration and varies with colony age (Tsekova et al.,
both azurite- and malachite-containing media, samples 2002; Tsekova and Galabova, 2003). These enzymes
from the central part of A. niger colonies with mature could mediate the transfer of polypeptide-bound copper to
ageing mycelium and abundant conidiophores with dark- a more stable form such as polyphosphates. Polyphos-
coloured conidia demonstrated phosphate coordination of phates have long been invoked in fungal metal
copper (Table 1, Figs 3J and 4B, D, E). In contrast, sequestration/tolerance mechanisms (Gadd, 1993) and,
samples from the marginal zone of colonies with young for example, are used by ectomycorrhizal fungi to immo-
vegetative mycelium showed sulfide coordination of bilize metals within their vacuoles (Bucking and Heyser,
copper (Table 1, Figs 3I and 4C and D). 1999) which would provide longer-term storage and pro-
The variation in copper coordination in A. niger could tection from toxic effects.
demonstrate the relative significance and spatial distribu-
tion of copper resistance mechanisms within the colony.
Zinc
We suggest that the sulfur coordination of copper in the
immature outer mycelia originates from the binding of In some metal-contaminated soils, a proportion of phos-
copper with metallothionein. These are small cysteine-rich phorus may exist as toxic metal phosphate minerals with
peptides primarily produced in response to copper, and very low solubility, e.g. hopeite [Zn3(PO4)2·4H2O] with a
play an important role in Cu homeostasis and sequestra- Ksp = 10-35.5. Analysis of some highly contaminated

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 313

sequent re-immobilization of zinc within the plant and


fungal biomass (Fomina et al., 2006b).
The results of bulk XAS-analysis of zinc coordination
in biomass (A. niger and P. simplicissimum grown on
ZnO; ectomycorrhizal R. rubescens grown on zinc phos-
phate supplied by either nitrate or ammonium as nitrogen
source; R. rubescens-, P. involutus- and Thelephora
terrestris-Scots pine ectomycorrhizal roots exposed to
zinc phosphate) were typically fitted with six oxygens at
distances between 1.99 and 2.07 Å with no clear evi-
dence for outer shells in any of the samples, although a
peak at c. 4 Å could be fitted well with multiple scattering,
assuming octahedral geometry (Table 2) (Fomina et al.,
2006a,b). This implies either that the zinc is not present in
a coordination environment with a well-defined long range
order, or that there are several zinc species present in
different coordination sites, most likely with carboxylate
ligands, although some phosphate coordination cannot be
ruled out. These data of mixed carboxylate/phosphate
coordination of zinc within the fungi are consistent with
synchrotron-based X-ray microfluorescence and
microEXAFS studies on zinc speciation in the zinc-
hyperaccumulating plant Arabidopsis halleri which
showed that zinc was distributed in Zn-malate, Zn-citrate
and Zn-phosphate (Sarret et al., 2002). Fungal and myc-
orrhizal biomass exposed to toxic metal minerals may
Fig. 3. Copper coordination within fungal biomass: Fourier
transforms of Cu K-edge XANES, EXAFS (solid lines) and fits
accumulate metals coordinated by different ligands (e.g.
(broken lines) for standard compounds (A) Cu(acetate), (B) carboxylate and phosphate). One drawback to the tech-
Cu(gluconate), (C) Cu(malate), (D) Cu(oxalate), (E) Cu(phosphate) nique used is that it is a bulk measurement, with a sample
and (F) Cu(I) oxide. Fungal biomass samples are the
ectomycorrhizal fungus R. rubescens grown in the presence of
area of c. 2 mm ¥ 2 mm, collecting a signal from all the
copper phosphate on MMN agar medium containing (G) ammonium absorber atoms present, so that if the metal is in several
and (H) nitrate, and A. niger grown on malt agar amended with different coordination sites only an average is seen. This
azurite; (I) analysis of marginal part of A. niger colony with new
vegetative mycelium, and (J) central part of colony with older
may be overcome by using micro-XAS techniques if their
conidiating mycelium. sensitivity and resolution is high enough (at least few
microns) to distinguish different metal speciation within
biosystem.
organic soils at moderate pH showed that most of the zinc The phosphate groups in fungal biomass can be of
was organically bound (~45%) and Zn-sorbed phosphate organic and inorganic origin (e.g. nucleic acids, adenosine
comprised only ~10% of the total (Sarret et al., 2004). phosphates, polyphosphates) and the precipitation of sec-
However, as the formation of metal phosphates in metal- ondary metal phosphates originating from the initial metal
contaminated soils [phosphate-induced metal stabiliza- phosphates can also occur. Carboxyl groups can be
tion (PIMS)] has been proposed as an in situ remediation derived from, e.g. low molecular weight organic acids,
technology, this may lead to an increased proportion in proteins, polysaccharides, (poly)phenols/quinones, and
soils of reaction products such as hopeite and mixed- melanin, and these groups have been previously reported
metal hopeites (Conca, 1997; Chen et al., 1997; Brown to account for up to 55% and 70% of zinc binding by cell
et al., 2004). It is well known that many fungi, including walls of Penicillium chrysogenum and Trichoderma reesei
symbiotic fungi, can solubilize insoluble zinc-bearing min- respectively (Sarret et al., 1998a). However, in other
erals, e.g. zinc phosphate and zinc oxide (Sayer et al., EXAFS studies of Zn binding sites in P. chrysogenum cell
1995; Sayer and Gadd, 1997; Martino et al., 2003; walls, zinc was predominantly complexed to four PO4
Fomina et al., 2004; 2005b,c; 2006a). It has been also groups in a tetrahedral configuration (Fourest et al., 1996).
shown that Paxillus involutus/pine ectomycorrhizas For R. rubescens/Pinus sylvestris ectomycorrhizas
exposed to zinc phosphate under phosphorus-deficient grown in a square Petri dish mesocosms with zinc phos-
conditions demonstrated a greater zinc mobilization from phate and nitrate Ingestad agar medium, we have
zinc phosphate compared with abiotic controls with con- observed micropatches of zinc precipitation in extramatri-

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
314 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd
Fig. 4. Models of copper coordination within
fungal biomass.
A. Model of Cu oxalate coordination within
biomass of R. rubescens in axenic culture and
ectomycorrhizal association with Scots pine
grown in the presence of copper phosphate
on media containing nitrate as a nitrogen
source.
The effect of age/physiological state of
mycelium on copper coordination within
A. niger biomass grown in the presence of
azurite or malachite.
B. Model of phosphate coordination of Cu in
central part of colony.
C. Model of sulfide coordination of Cu in
marginal zone of colony.
D. Colony of A. niger on malt extract agar
medium with white marginal zone of young
vegetative mycelium and black central zone
with profuse conidia production. Scale bar is
1 cm.
E. Abundant conidiophores carrying conidia in
black central zone of A. niger colony. Scale
bar is 200 mm.

cum mycelium. Electron microscopy techniques coupled formed by some saprotrophic and mycorrhizal fungi
with elemental analysis by EDX revealed that zinc was (Sayer et al., 1999; Fomina et al., 2004; 2005b,c). This
deposited intracellularly (Fig. 2B and C). Though full ability of fungi to transform pyromorphite should be taken
XRPD identification of precipitates was not possible (due into account in assessments of the long-term environmen-
to either very small amount or poor crystallinity), the tal consequences of in situ chemical remediation
absence of a phosphorus peak in the EDX spectra com- techniques. Some fungi excrete strong chelators (oxalic
bined with our EXAFS results (Table 2) indicated carboxy- acid) that shift the mechanism of pyromorphite dissolution
late coordination of zinc within intracellular deposits. from acidification to much more efficient ligand-promoted
Some precipitates reached a size more than 1 mm across, lead mobilization (Fomina et al., 2004). Fungi are also
filling more than one-third of the inner space of fungal able to immobilize lead released from pyromorphite by
cells, probably making them not functional and viable. extracellular precipitation as lead oxalate (Fomina et al.,
This could mean that mycelial system sacrificed some 2005c) and by Pb accumulation within the biomass
hyphae to immobilize and detoxify the metal by locking it (Fomina et al., 2004; 2005b). Axenic cultures of ericoid
within cell walls, enabling it to survive and advance across and ectomycorrhizal fungi grown with pyromorphite accu-
the ameliorated environment. mulated from 2 to 5 mg Pb per g dry wt biomass: isolates
of the ectomycorrhizal fungus Suillus bovinus were able to
accumulate up to 12 mg Pb per g dry wt biomass (Fomina
Lead
et al., 2005b). Here, in general, biomass of the ectomyc-
Lead is the most common toxic metal which becomes orrhizal fungi (P. involutus, R. rubescens, S. bovinus and
concentrated through the food chain, posing a toxic risk to Suillus collimitus), the ericoid mycorrhizal fungus
many species, including humans (Nriagu, 1996). Pyro- Hymenoscyphus ericae, and B. caledonica exposed to
morphite, or lead chlorophosphate [Pb5(PO4)3Cl], is the pyromorphite demonstrated phosphate coordination of
most geochemically stable lead mineral which forms in accumulated Pb with the inner coordination sphere best
urban and industrially contaminated soils and is also one fitted with a split shell of one oxygen at 2.14 Å, four
of the reaction products of PIMS remediation of lead- oxygens at 2.40 Å and two oxygens at 3.04 Å and a
contaminated soils (Cotter-Howells and Caporn, 1996; significant improvement in the fit on addition of a shell of
Conca, 1997; Zhang et al., 1997; Brown et al., 2004). two phosphorus scatterers at 3.45 Å (Table 3, Figs 5G
However, pyromorphite can be solubilized and trans- and H and 6A).

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 315
Table 2. Zn K-edge EXAFS parameters.

Sample Scatterers r (Å) 2s2 (Å2) Residual

Zn(acetate) standard 4¥O 1.93 0.019 29.7


4¥C 2.86 0.015
4¥O 3.08 0.016
2 ¥ Zn 3.34 0.025
2 ¥ Zn 3.95 0.026
Zn(citrate) standard 6¥O 2.03 0.017 19.6
4¥C 2.78 0.031
Zn3(PO4)2 standard 4¥O 1.96 0.014 23.6
2¥P 3.13 0.032
2 ¥ Zn 3.33 0.016
Zn(gluconate) standard 6¥O 2.06 0.014 17.4
6¥C 2.96 0.047
2 ¥ Zn 3.96 0.026
Zn(malate) standard 6¥O 2.01 0.022 33.6
6¥C 2.81 0.052
2 ¥ Zn 3.88 0.026
Zn(oxalate) standard 6¥O 2.09 0.012 23.2
4¥C 2.81 0.028
12 ¥ O 3.81 0.043
ZnS standard 4¥S 2.35 0.010 31.1
12 ¥ Zn 3.84 0.017
12 ¥ S 4.49 0.022
6 ¥ Zn 5.42 0.016
A. niger on ZnO-containing AP1 agar 6¥O 2.06 0.018 32.0
P. simplicissimum on ZnO-containing AP1 agar 6¥O 2.03 0.024 37.7 (40.9)
R. rubescens axenic culture grown on copper phosphate and ammonium 6¥O 2.07 0.024 27.2
MMN agar medium
R. rubescens/P. sylvestris ectomycorrhizas grown in a square 6¥O 2.02 0.031 22.7
medium Petri dish mesocosm with zinc phosphate and nitrate Ingestad agar
R. rubescens/P. sylvestris ectomycorrhizas grown in a square Petri 6¥O 2.07 0.016 28.8 (22.6)
dish mesocosm with zinc phosphate and Ingestad agar medium
P. involutus/P. sylvestris ectomycorrhizas grown in a square Petri 6¥O 2.06 0.017 33.8 (27.9)
dish mesocosm with zinc phosphate and Ingestad agar medium
T. terrestris/P. sylvestris ectomycorrhizas grown in a square Petri 6¥O 2.05 0.023 33.1 (25.4)
dish mesocosm with zinc phosphate and Ingestad agar medium

r is the zinc–scatterer distance in Angstroms, ⫾0.02 Å inner shells, ⫾0.05 Å outer shells; 2s2 is the Debye–Waller type factor, ⫾15% inner shells,
⫾30% outer shells (a measure of both the thermal motion between the absorber and scatterer and of the static disorder or range of absorber–
scatterer distances); the residual is a least squares residual from fitting the spectrum of the model to the experimental data. For some samples
inclusion of multiple scattering, assuming an octahedral coordination site, improved the fit but didn’t change the fitting parameters. In these cases
the residuals in brackets in the final column are the residuals without multiple scattering.

Fungal ability to solubilize other lead-bearing com- ammonium-containing medium, where outer shells could
pounds (lead phosphate, tetraoxide, carbonate and be fitted, the indication was that carboxylic groups were
sulfide) has also been shown previously (Sayer and the main ligands rather than phosphates (Table 3, Fig. 5I
Gadd, 1997; Fomina et al., 2005c). For most of the and J). For example, EXAFS data for A. niger were col-
EXAFS spectra recorded for A. niger, B. caledonica and lected to a k-range of 11 Å-1. The main peak could be
P. simplicissimum grown on nitrate- or ammonium- fitted with a shell of six to eight oxygens at 2.55 Å, but the
containing media with lead oxides (PbO or Pb3O4), only fit was improved by splitting this into two shells, at 2.45 Å
the inner coordination sphere could be fitted, in all cases and 2.56 Å. The coordination numbers should not be
with oxygen scatterers (Table 3). The nitrogen source did regarded as definitive, because with the large Pb ion there
not seem to alter lead coordination. For most samples, the is usually significant disorder in the coordinated Pb-O
spectra were noisy with useful data up to a k-range of distances and consequently large errors in coordination
7–8 Å-1 and the best fit with a single shell of oxygen numbers. Outer shells were best fitted with carbons at
scatterers at 2.39 Å. For B. caledonica grown on PbO- 3.31 Å, implying carboxylic acid coordination, and a
and ammonium-containing medium a fit of six oxygens at heavier element, such as iron but not lead, at 3.81 Å
2.44 Å could be improved by splitting into two, at 2.14 Å (Fig. 6B).
and 2.40 Å (Table 3). However, in two cases when It has been reported that coordination of lead within
A. niger was grown on PbO- and nitrate-containing fungal biomass can be affected by lead concentration
medium and B. caledonica was grown on Pb3O4- and (Sarret et al., 1999). Previous EXAFS studies of lead

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
316 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd
Table 3. Pb L(III)-edge EXAFS parameters.

Sample Scatterers r (Å) 2s2 (Å2) Residual

Pb5Cl(PO4)3 standard 0.5 ¥ O 2.31 0.009 59.2


5¥O 2.56 0.066
1 ¥ Cl 2.87 0.075
2¥P 3.24 0.035
1 ¥ Pb 3.32 0.018
Pb3O4 standard 4.67 ¥ O 2.19 0.021 37.5
0.67 ¥ Pb 3.30 0.012
2.67 ¥ Pb 3.61 0.015
3.33 ¥ Pb 3.78 0.020
Pb(oxalate) standard 3¥O 2.39 0.026 45.7
4¥O 2.59 0.066
6¥C 3.22 0.026
4 ¥ Pb 3.99 0.040
PbS standard 6¥S 2.95 0.022 37.7
12 ¥ Pb 4.19 0.021
8¥S 5.04 0.047
6 ¥ Pb 5.99 0.019
Pb3(PO4)2 standard 4¥O 2.38 0.044 56.4
4¥O 2.61 0.039
4¥P 3.29 0.056
PbO standard 4¥O 2.26 0.016 55.1
4 ¥ Pb 3.59 0.027
4 ¥ Pb 3.72 0.011
4 ¥ Pb 3.93 0.017
Pb(II) (acetate) standard 4¥O 2.49 0.046 71.2
Pb(IV) (acetate) standard 8¥O 2.24 0.020 42.8
4¥C 2.68 0.007
4¥C 4.15 0.012
Pb(malate) standard 4¥O 2.46 0.070 65.4
B. caledonica grown on pyromorphite-containing MMN agar 4¥O 2.62 0.010 51.0
4¥O 2.79 0.007
4¥P 3.25 0.044
H. ericae DGC3 grown on pyromorphite-containing MMN agar 2¥O 2.02 0.036 45.4
4¥O 2.31 0.027
2¥P 3.49 0.061
H. ericae C2 grown on pyromorphite-containing MMN agar 1¥O 1.89 0.013 54.7
4¥O 2.20 0.024
2¥O 2.42 0.015
2¥P 3.52 0.033
P. involutus 23 grown on pyromorphite-containing MMN agar 1¥O 2.11 0.015 50.6
4¥O 2.38 0.030
2¥O 3.13 0.090
2¥P 3.68 0.050
R. rubescens B20 grown on pyromorphite-containing MMN agar 1¥O 2.13 0.021 49.4
4¥O 2.39 0.031
2¥O 3.02 0.030
2¥P 3.55 0.054
S. bovinus Lst8 grown on pyromorphite-containing MMN agar 1¥O 2.14 0.010 51.5
4¥O 2.40 0.025
2¥O 3.04 0.012
2¥P 3.45 0.021
S. collimitus 22 grown on pyromorphite-containing MMN agar 1¥O 2.18 0.015 40.7
4¥O 2.42 0.017
2¥O 2.62 0.015
2¥P 3.61 0.029
A. niger grown on PbO and nitrate-containing AP1 agar 4¥O 2.45 0.027 31.4
4¥O 2.56 0.010
4¥C 3.31 0.014
4 ¥ Fe 3.81 0.026
A. niger grown on Pb3O4 and ammonium-containing AP1 agar 6¥O 2.55 0.034 81.8
A. niger grown on Pb3O4 and nitrate-containing AP1 agar 6¥O 2.49 0.042 53.2
B. caledonica grown on PbO and nitrate-containing AP1 agar 6¥O 2.42 0.051 59.4
B. caledonica grown on Pb3O4 and nitrate-containing AP1 agar 6¥O 2.39 0.059 48.0
B. caledonica grown on PbO and ammonium-containing AP1 agar 2¥O 2.14 0.043 46.8
4¥O 2.40 0.023
B. caledonica grown on Pb3O4 and ammonium-containing AP1 agar 2¥O 2.17 0.056 29.0
4¥O 2.41 0.032
4¥C 3.30 0.042
P. simplicissimum grown on Pb3O4 and nitrate-containing AP1 agar 2¥O 2.22 0.059 56.4
4¥O 2.48 0.014
P. simplicissimum grown on PbO and ammonium-containing AP1 agar 6¥O 2.40 0.060 49.8
P. simplicissimum grown on Pb3O4 and ammonium-containing AP1 agar 6¥O 2.40 0.072 50.1

r is the lead–scatterer distance in Angstroms, ⫾0.02 Å inner shells, ⫾0.05 Å outer shells; 2s2 is the Debye–Waller type factor, ⫾15% inner shells, ⫾30% outer shells
(a measure of both the thermal motion between the absorber and scatterer and of the static disorder or range of absorber–scatterer distances); the residual is a least
squares residual from fitting the spectrum of the model to the experimental data.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 317

species (Table 3). This means that the fungi which grew
on and solubilized red lead oxide (Pb3O4) with mixed lead
valencies [Pb(II) + Pb(IV)] selectively accumulated Pb(II)
in the biomass. However, it is known that the Pb(II) oxi-
dation state is more stable than Pb(IV), and there is a
strong tendency for lead(IV) compounds to react to give
lead(II) compounds, especially under acidic conditions.

Conclusions

Our goal was to determine the speciation of toxic metals


accumulated by fungi and mycorrhizal associations
exposed to toxic metal minerals. XAS techniques gave us
a means of revealing the speciation of toxic metals that
underwent a cycle of chemical transformations caused by
fungi.
Our data have shown that:
(i) Oxygen ligands (phosphate and carboxylate) played
a major role in metal coordination within fungal and
ectomycorrhizal biomass during the accumulation of
toxic metals mobilized by fungi from metal phos-
phates and oxides.
(ii) A copper sulfide phase was found in fungi grown on
azurite and malachite.
(iii) Metal precipitation by fungal and mycorrhizal
biomass was not homogeneous in many cases dem-
Fig. 5. Lead coordination within fungal biomass: Fourier transforms onstrating mixed metal coordination and indicating
of Pb L(III)-edge XANES, EXAFS (solid lines) and fits (broken some limitations of bulk XAS analysis of biomass.
lines) for standard compounds: (A) Pb(malate), (B) Pb(oxalate), (C)
Pb(II) (acetate), (D) Pb3(PO4)2, (E) PbO, (F) Pb3O4. Fungal biomass (iv) Coordination of toxic metals within biomass
samples are (G) S. bovinus Lst8 and (H) B. caledonica grown in depended on the fungal species, the supplied
the presence of pyromorphite, and (I) A. niger grown on AP1 agar mineral, nitrogen source, and the physiological state/
medium containing nitrate and PbO, and (J) B. caledonica grown
on AP1 agar medium containing ammonium and Pb3O4. age of the fungal mycelium.
Metal-contaminated terrestrial environments are
bound to cell walls of P. chrysogenum showed that car- dynamic systems with metal contaminants changing their
boxyl groups have greater affinity for lead ions and pref- speciation with time depending on physico-chemical con-
erentially bound Pb at low metal concentrations whereas ditions and biogeochemical activities of the biota. The ‘in
at higher concentrations phosphoryl groups accounted for situ’ situation in metal-polluted soils has a much more
up to 95% of Pb binding, with carboxyl accounting for the complex geochemical environment than that used in our
remaining 5% (Sarret et al., 1999). Our data have shown experiments. However, our experiments with ectomycor-
that the Pb-O distances were all consistent with Pb(II) rhizal associations grown in the zinc phosphate spiked

Fig. 6. Models of lead coordination within


fungal biomass.
A. Model of phosphate coordination of lead
accumulated within biomass of a variety of
mycorrhizal and saprotrophic fungi (e.g.
H. ericae, P. involutus, S. bovinus,
S. collimitus) grown in the presence of
pyromorphite.
B. A model for carboxylate coordination of
lead accumulated by A. niger grown on PbO-
and nitrate-containing AP1 agar medium with
outer shells also fitting a metal element (Met)
such as iron or nickel.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
318 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd

mesocosms, where vermiculite/peat matrix represented [PbO and Pb3O4 (Aldrich) (red lead)]). Axenic cultures of fungi
both inorganic and organic soil components, simulated were grown at 23–25°C in sterile Petri-dish microcosms on
metal-polluted ‘field’ conditions. It has been observed that top of cellophane membranes on a variety of media: malt
extract agar, AP1 medium [(NH4)2SO4 (5 g l–l), KH2PO4
ectomycorrhizas immobilized half of total zinc mobilized
(0.5 g l–l), MgSO4.7H2O (0.2 g l–l), CaCl2.6H2O (50 mg l–l),
from zinc phosphate whereas only one-tenth of mobilized NaCl (100 mg l–l), FeCl3.6H2O (2.5 mg l–l), ZnSO4.7H2O,
zinc was immobilized by the vermiculite/peat matrix (see MnSO4.4H2O, and CuSO4.5H2O (each 4 mg l–l), D-glucose
Fomina et al., 2006b). Thus, in phytoremediation and (20 g l-1), agar No. 1 (Laboratory M, Bury UK) (15 g l–l)], and
reforestation/revegetation sites with abundant vegetation, modified Melin-Norkrans medium for mycorrhizal fungi
phyto- and fungal biomass may play a crucial role in toxic (MMN) [(NH4)2HPO4 (0.5 g l–l), KH2PO4 (0.3 g l–l), MgSO4.
metal transformations and immobilizations. 7H2O (0.14 g l–l), CaCl2.6H2O (50 mg l–l), NaCl (25 mg l–l),
D-glucose (10 g l-1), glutamic acid (1 mg l–l), thiamine
The conditions of growth both in vitro and ex vitro may
(100 mg l–l) and agar No. 1 (Laboratory M, Bury UK) (14 g l–l)]
change the nature of ligands provided by fungal and myc- (Fomina et al., 2005b). Previously it was shown that all tested
orrhizal biomass. For example, the change of nitrogen fungi were able to solubilize zinc- and copper-containing min-
source from ammonium to nitrate may increase the pro- erals and pyromorphite with oxalic and citric acid overproduc-
portion of oxalate excretion and subsequently shift ing fungi A. niger and B. caledonica being very efficient
towards oxalate coordination of toxic metal (copper) mineral transformers (Sayer et al., 1999; Fomina et al., 2004;
within biomass (R. rubescens). However, it appears that 2005b,c). Biomass of fast-growing saprotrophic fungi A. niger
and P. simplicissimum was harvested after 7 day growth,
for certain toxic metal minerals fungal/mycorrhizal
other slow growing fungi were harvested after 2 months.
biomass provides the same selection of functional groups
Synthesis of ectomycorrhizas was carried out under sterile
(e.g. phosphate, carboxylate, sulfhydryl) in both agar and conditions using a test tube technique (Peterson and Chakra-
vermiculite/peat and in both axenic culture and ectomyc- varty, 1991) in autoclaved (120°C, 60 min) glass tubes
orrhizal association, with an obvious predominance of (culture tubes 25 mm ¥ 150 mm, Sigma-Aldrich Cat C 5916,
oxygen-containing ligands. It is likely that ‘in situ’ fungal/ Dorset, UK) filled with 30 ml vermiculite (Sinclair, Lincoln, UK)
mycorrhizal biomass, comprising the largest biomass in and peat (Irish Moss Peat, B and Q plc, Chandlers Ford, UK)
soils, will also provide those functional groups for metal mixture 5:1 and 10 ml of Ingestad medium comprising: K2SO4
(12.25 mg l–l); KH2PO4 (8.58 mg l–l); K2HPO4 (10.11 mg l–l);
immobilization.
KNO3 (9.71 mg l–l); NH4NO3 (58.56 mg l–l); Ca(NO3)2
The roles of free-living and symbiotic fungi have previ- (8.61 mg l–l); Mg(NO3)2 15.84 mg l–l; HNO3 (13.7 mg l–l);
ously been a much underestimated component of metal H3BO3 (0.57 mg l–l); Mn(NO3)2 (0.913 mg l–l); Zn(NO3)2
biogeochemistry but these organisms are involved in pro- (0.07 mg l–l); CuCl2 (0.04 mg l–l); Na2MoO4 (0.009 mg l–l);
cesses of both metal mobilization and immobilization. FeNaEDTA (1.149 mg l–l) (Ingestad, 1979). D-glucose
Understanding the molecular nature of toxic metal immo- (0.5 g l–l) was also added to the medium to initiate fungal
bilization by fungi is of relevance to the development and growth in the matrix. Seeds of Pinus sylvestris L. (Chiltern
seeds, Cat no. 1014, Ulverston, Cumbria, UK) were surface-
maintenance of efficient remediation techniques as well
sterilized in 30% hydrogen peroxide with one drop of Tween
as a fuller understanding of the chemical and biological 60 for 30 min, then thoroughly washed with sterile water and
interactions of organisms to environmental pollutants. transferred into Petri dishes containing MMN agar medium to
germinate for at least 14 days. Germinating pine seeds were
planted into vermiculite/peat matrix together with four discs of
Experimental procedures
7 mm diameter discs of mycelium cut from the leading edge
Fungi and mycorrhizas of colonies which had been maintained on MMN at 25°C for
at least 14 days. Seedlings were cultivated for 3–4 months in
Fungal cultures (saprotrophic Aspergillus niger, Penicillium the growth chamber (Fisons Fi-totron 600 Growth Cabinet,
simplicissimum, entomopathogenic Beauveria caledonica, Suffolk, UK) with 200 mmol m-2 s-1 PAR, at least 60% relative
ericoid mycorrhizal Hymenoscyphus ericae and Oidioden- air humidity and at day/night regime of 18/6 h and a tempera-
dron maius and ectomycorrhizal Lactarius turpis, Paxillus ture of 23/15°C. Mycorrhizal colonization of the roots was
involutus, Rhizopogon rubescens, Suillus bovinus, Suillus nearly 85–90%. After successful colonization, uniform pine
collimitus) and ectomycorrhizal roots of associations of Scots seedlings were carefully taken out from tubes, washed with
pine (Pinus sylvestris) with P. involutus, R. rubescens and sterile water and transferred into sterile square Petri dishes
T. terrestris were used in this study (see Fomina et al., mesocosms. Square Petri dishes were filled with Ingestad
2005b). We exposed fungi and ectomycorrhizas to a variety agar amended with zinc or copper phosphate to a final metal
of copper-, zinc- and lead-containing minerals at a final metal concentration equivalent to 15 mM as for axenic cultures of
concentration equivalent to 15 mM (copper and zinc phos- fungi. The roots of mycorrhizal seedlings were placed over
phates and oxides [Cu3(PO4)2.2H2O (Fluka), Zn3(PO4)2.2H2O sterile cellophane membrane on top of the agar medium. The
(Alfa) (naturally occurring as hopeite), CuO (Aldrich), Cu2O upper cellophane membrane and an autoclaved piece of
(BDH) (naturally occurring as cuprite), ZnO], azurite foam, soaked in sterile water, were placed on top of the roots
[Cu3(CO3)2(OH)2], malachite [Cu2(CO3)(OH)2], pyromorphite to prevent them from desiccation and to fix the position of the
[Pb5(PO4)3Cl] (Fomina et al., 2004), lead oxide and tetraoxide root system. The shoot was kept outside of the Petri dish by

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 319
means of a hole in the side of the plastic dish and sealed ating at 2 GeV with an average current of 150 mA, using a
around the stem with sterile lanolin. Mesocosms were incu- vertically focusing mirror and a sagittally bent focusing
bated for at least 3 months under the same conditions as Si(220) double crystal monochromator detuned to 70% trans-
described above. In some experiments with axenic fungal mission to minimize harmonic contamination. Data were col-
cultures and the R. rubescens/P. sylvestris ectomycorrhizal lected with the station operating in fluorescence mode using
association we also used a nitrate-only version of MMN/AP1 an Ortec 30 element solid state Ge detector. The monochro-
and Ingestad media substituting KNO3 (0.8 g l–l and mator calibration was checked using a Pb3(PO4)2 standard.
17.34 mg l–l) for (NH4)2HPO4 and NH4NO3 respectively. Experiments were performed in a liquid nitrogen cooled
Biomass of fungi and ectomycorrhizas grown over cello- cryostat. Spectra of model compounds were collected in
phane membrane in Petri dishes was harvested by peeling transmission mode. Single scans were collected for the
from the membrane. Ectomycorrhizal roots of seedlings model compounds, and three to four scans were collected
grown in vermiculite/peat mixture were removed from the and summed for each sample. Model compounds were
matrix, washed with distilled water, dried with paper towels, Pb-oxalate, Pb3(PO4)2, Pb5Cl(PO4)3, Pb3O4, PbO, Pb-malate,
and then cut with a razor blade. Freshly harvested biomass of Pb(II)-acetate, Pb(IV)-acetate and PbS.
fungi and ectomycorrhizas, enclosed in cellotape and
quenched in liquid nitrogen, was used for X-ray absorption EXAFS spectra analysis. Background-subtracted EXAFS
spectrometry. spectra were analysed in EXCURV98 using full curved wave
theory (Gurman et al., 1984; Binsted, 1998). Phaseshifts
were derived in the program from ab initio calculations using
Electron microscopy Hedin–Lundqvist potentials and von Barth ground states
Accumulation of toxic metals by fungal mycelia was studied (Hedin and Lundqvist, 1969). Fourier transforms of the
using scanning and transmission electron microscopy. For EXAFS spectra were used to obtain an approximate radial
scanning electron microscopy we used air-dried samples, distribution function around the central metal (Cu, Zn or Pb)
uncoated or coated with 30 nm Au/Pd using a Cressington (the absorber atom): the peaks of the Fourier transform can
208 HR sputter coater. Samples were examined using a be related to ‘shells’ of surrounding back scattering atoms
Philips XL30 environmental scanning electron microscope characterized by atom type, number of atoms in the shell, the
(ESEM) field emission gun (FEG) operating at an accelerat- absorber–scatterer distance, and the Debye–Waller factor,
ing voltage of 15 or 25 kV. For transmission electron micro- 2s2 (a measure of both the thermal motion between the
scopy we used samples of mycelium fixed in 2.5% absorber and scatterer and of the static disorder or range of
glutaraldehyde in 5 mM 1,4-piperazinediethanesulfonic acid absorber–scatterer distances). The data were fitted for each
(PIPES) buffer, pH 6.5, washed in PIPES buffer, then dehy- sample by defining a theoretical model and comparing the
drated with an ascending series of ethanol in distilled water calculated EXAFS spectrum with the experimental data.
and embedded into L.R. White resin. Unstained ultra-thin Shells of back scatterers were added around the absorber
sections (cut on a Reichert OMU-3 microtome) mounted on and by refining an energy correction Ef (the Fermi energy),
pioloform-copper grids were examined using a Jeol-1200 EX the absorber–scatterer distance, and Debye–Waller factor for
transmission electron microscope. each shell, a least squares residual [the R-factor (Hedin and
Lundqvist, 1969; Binsted et al., 1992) was minimized]. Where
appropriate, multiple scattering effects were included in the
X-ray absorption spectrometry fits (Gurman et al., 1986).
Cu and Zn. X-ray absorption spectra at the Cu and Zn
K-edges were collected on Station 7.1 at the CCLRC Dares- Acknowledgements
bury SRS operating at 2 GeV with an average current of G.M.G. gratefully acknowledges funding from the BBSRC/
140 mA, using a vertically collimating plane mirror and a BIRE programme (94/BRE13640) and CCLRC Daresbury
sagittally bent focusing Si(111) double crystal monochroma- SRS (SRS user grants 40107 and 43079). The authors grate-
tor detuned to 80% transmission to minimize harmonic fully acknowledge the help of the 7.1 and 16.5 stations sci-
contamination. Sample data were collected with the station entist Drs Fred Mosselmans, Lorrie Murphy, Bob Bilsborrow
operating in fluorescence mode using a nine-element solid and Steve Fiddy (CCLRC Daresbury SRS) and Dundee Uni-
state Ge diode detector with high count-rate XPRESS pro- versity team members Mr Ewan Starke, Miss Karrie Melville,
cessing electronics: spectra of model compounds were col- Mr Martin Kierans, Miss Louise McGregor, Dr Euan Burford,
lected in the transmission mode. The monochromator was Dr Simon Hockin and Dr Ademola Adeyemi. A.D.B. gratefully
calibrated using a 5 mm Cu or Zn foil. Experiments were acknowledges receipt of a BBSRC postgraduate research
performed using a liquid nitrogen cooled cryostat. Single studentship. We thank Professor Andy Meharg and Dr David
scans were collected for the model compounds, and three to Genney (University of Aberdeen, Scotland), Dr Elena Martino
four scans were collected and summed for each sample. (University of Torino, Italy), Dr Derek Mitchell (University
Model compounds used were Cu-acetate, Cu-gluconate, College Dublin, Ireland), Dr Jan Colpaert (Limburgs Univer-
Cu-malate, Cu-oxalate, Cu3(PO4)2 and Cu2O for Cu and sity Centre, Belgium), Dr Claude Plassard (INRA, France),
Zn-acetate, Zn-citrate, Zn-gluconate, Zn-malate, Zn-oxalate, and Dr Håkan Wallander (Lund University, Sweden) for the
Zn3(PO4)2 and ZnS for Zn. provision of fungal strains, and Mr Martin Kierans [Centre for
High Resolution Imaging and Processing (CHIPs), School of
Pb. X-ray absorption spectra at the Pb L(III)-edge were col- Life Sciences, University of Dundee, Scotland] for assistance
lected on Station 16.5 at the CCLRC Daresbury SRS oper- with electron microscopy.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
320 M. Fomina, J. Charnock, A. D. Bowen and G. M. Gadd

References Fourest, E., Serre, A., and Roux, J.C. (1996) Contribution of
carboxylic groups to heavy metal binding sites in fungal cell
Bellion, M., Courbot, M., Jacob, C., Blaudez, D., and Chalot, walls. Toxicol Environ Chem 54: 1–10.
M. (2006) Extracellular and cellular mechanisms sustaining Gadd, G.M. (1993) Interactions of fungi with toxic metals.
metal tolerance in ectomycorrhizal fungi. FEMS Microbiol New Phytol 124: 25–60.
Lett 254: 173–181. Gadd, G.M. (2004) Mycotransformation of organic and inor-
Binsted, N. (1998) Daresbury Laboratory EXCURV98 ganic substrates. Mycologist 18: 60–70.
Program. Gadd, G.M., Fomina, M., and Burford, E.P. (2005) Fungal
Binsted, N., Strange, R.W., and Hasnain, S.S. (1992) roles and functions in rock, mineral and soil
Constrained and restrained refinement in EXAFS data transformations. In Micro-Organisms and Earth Systems –
analysis with curved wave theory. Biochemistry 31: 12117– Advances in Geomicrobiology. Gadd, G.M., Semple, K.T.,
12125. and Lappin-Scott, H.M. (eds). Cambridge, UK: Cambridge
Brown, S., Chaney, R., Hallfrisch, J., Ryan, J.A., and Berti, University Press, pp. 201–232.
W.R. (2004) In situ soil treatments to reduce the phyto- and Gardea-Torresdey, J.L., de la Rosa, G., and Peralta-Videa,
bioavailability of lead, zinc, and cadmium. J Environ Qual J.R. (2004) Use of phytofiltration technologies in the
33: 522–531. removal of heavy metals: a review. Pure Appl Chem 76:
Bucking, H., and Heyser, W. (1999) Elemental composition 801–813.
and function of polyphosphates in ectomycorrhizal fungi- Gharieb, M.M., and Gadd, G.M. (1999) Influence of nitrogen
and X-ray microanalytical study. Mycol Res 103: 31–39. source on the solubilization of natural gypsum. Mycol Res
Burford, E.P., Fomina, M., and Gadd, G.M. (2003) Fungal 103: 473–481.
involvement in bioweathering and biotransformation of Gorbushina, A.A., Krumbein, W.E., Hamann, R., Panina, L.,
rocks and minerals. Mineral Mag 67: 1127–1155. Soucharjevsky, S., and Wollenzien, U. (1993) On the role
Burgstaller, W., and Schinner, F. (1993) Leaching of metals of black fungi in colour change and biodeterioration of
with fungi. J Biotechnol 27: 91–116. antique marbles. Geomicrobiol J 11: 205–221.
Charnock, J.M. (1995) Biological applications of EXAFS Greenwood, N.N., and Earnshaw, A. (1984) Chemistry of the
spectroscopy. Radiat Phys Chem 45: 385–391. Elements. Oxford, UK: Pergamon Press, p. 1087.
Chen, X.-B., Wright, J.V., Conca, J.L., and Peurrung, L.M. Gurman, S.J., Binsted, N., and Ross, I. (1984) A rapid, exact,
(1997) Evaluation of heavy metal remediation using curved-wave theory for EXAFS calculations. J Phys Chem
mineral apatite. Water, Air, Soil Poll 98: 57–78. 17: 143–151.
Conca, J.L. (1997) Phosphate-Induced Metal Stabilization Gurman, S.J., Binsted, N., and Ross, I. (1986) A rapid, exact,
(PIMS). Final report to the U.S. Environmental Protection curved-wave theory for EXAFS calculations. 2. The
Agency 68D60023, Res. Triangle Park, NC. multiple-scattering contributions. J Phys Chem 19: 1845–
Cotter-Howells, J.D., and Caporn, S. (1996) Remediation of 1861.
contaminated land by formation of heavy metal Hasnain, S.S. (1988) Application of EXAFS to biochemical
phosphates. Appl Geochem 11: 335–342. systems. Top Curr Chem 147: 73–93.
Fomina, M.A., Alexander, I.J., Hillier, S., and Gadd, G.M. Hedin, L., and Lundqvist, S. (1969) Effects of electron-
(2004) Zinc phosphate and pyromorphite solubilization by electron and electron–phonon interactions on the one-
soil plant-symbiotic fungi. Geomicrobiol J 21: 351–366. electron states of solids. Solid State Phys 23: 1–181.
Fomina, M., Burford, E.P., and Gadd, G.M. (2005a) Toxic Ingestad, T. (1979) Mineral nutrient requirements of Pinus
metals and fungal communities. In The Fungal Community. silvestris and Picea abies seedlings. Physiol Plant Pathol
Its Organization and Role in the Ecosystem. Dighton, J., 45: 373–380.
White, J.F., and Oudemans, P. (eds). Boca Raton, FL, Jaeckel, P., Krauss, G., Menge, S., Schierhorn, A., Rückna-
USA: CRC Taylor & Francis, pp. 733–758. gel, P., and Krauss, G.-J. (2005) Cadmium induces a novel
Fomina, M.A., Alexander, I.J., Colpaert, J.V., and Gadd, G.M. metallothionein and phytochelatin 2 in an aquatic fungus.
(2005b) Solubilization of toxic metal minerals and metal Biochem Biophys Res Comm 333: 150–155.
tolerance of mycorrhizal fungi. Soil Biol Biochem 37: 851– Kemner, K.M., O’Loughlin, E.J., Kelly, S.D., and Boyanov,
866. M.I. (2005) Synchrotron X-ray investigations of mineral–
Fomina, M., Hillier, S., Charnock, J.M., Melville, K., Alex- microbe–metal interactions. Elements 1: 217–221.
ander, I.J., and Gadd, G.M. (2005c) Role of oxalic acid Knox, A.S., Seaman, J.C., Mench, M.J., and Vangronsveld, J.
over-excretion in toxic metal mineral transformations by (2000) Remediation of metal- and radionuclides-
Beauveria caledonica. Appl Environ Microbiol 71: 371– contaminated soils by in situ stabilization techniques. In
381. Environmental Restoration of Metals-Contaminated Soil.
Fomina, M., Burford, E.P., and Gadd, G.M. (2006a) Fungal Iskandar, I.K. (ed.). Boca Raton, FL, USA: Lewis Publish-
dissolution and transformation of minerals: significance for ers, pp. 21–61.
nutrient and metal mobility. In Fungi in Biogeochemical Kumar, K.S., Dayananda, S., and Subramanyam, C. (2005)
Cycles. Gadd, G.M. (ed.). Cambridge, UK: Cambridge Uni- Copper alone, but not oxidative stress, induces copper-
versity Press, pp. 236–266. metallothionein gene in Neurospora crassa. FEMS Micro-
Fomina, M., Charnock, J.M., Hillier, S., Alexander, I.J., and biol Lett 242: 45–50.
Gadd, G.M. (2006b) Zinc phosphate transformations by the Lapeyrie, F., Ranger, J., and Vairelles, D. (1991) Phosphate-
Paxillus involutus/pine ectomycorrhizal association. Microb solubilizing activity of ectomycorrhizal fungi in vitro. Can J
Ecol (in press). Bot 69: 342–346.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321
Toxic metal speciation in fungi 321
Liu, S.-H., and Wang, H.P. (2004) In situ speciation studies Sarret, G., Schroeder, W.H., Marcus, M.A., Geoffroy, N., and
of copper-humic substances in a contaminated soil during Manceau, A. (2003) Localization and speciation of Zn in
electrokinetic remediation. J Environ Qual 33: 1280– mycorrhized roots by mSXRF and mEXAFS. J Phys IV
1287. France 107: 1193–1196.
Martino, E., Perotto, S., Parsons, R., and Gadd, G.M. (2003) Sarret, G., Balesdent, J., Bouziri, L., Garnier, J.-M., Marcus,
Solubilization of insoluble inorganic zinc compounds by M.A., Geoffroy, N., et al. (2004) Zn speciation in the
ericoid mycorrhizal fungi derived from heavy metal polluted organic horizon of a contaminated soil by micro-X-ray fluo-
sites. Soil Biol Biochem 35: 133–141. rescence, micro- and powder-EXAFS spectroscopy, and
Meharg, A.A., and Cairney, J.W.G. (2000) Co-evolution of isotopic dilution. Environ Sci Technol 38: 2792–2801.
mycorrhizal symbionts and their hosts to metal- Sayer, J.A., and Gadd, G.M. (1997) Solubilisation and
contaminated environments. Adv Ecol Res 30: 69–112. transformation of insoluble metal compounds to insoluble
Nriagu, J.O. (1996) A history of global metal pollution. metal oxalates by Aspergillus niger. Mycol Res 101: 653–
Science 272: 223. 661.
Peterson, R.L., and Chakravarty, P. (1991) Techniques in Sayer, J.A., Raggett, S.L., and Gadd, G.M. (1995) Solubili-
synthesizing ectomycorrhiza. In Techniques for Mycor- zation of insoluble compounds by soil fungi: development
rhizal Research. Norris, J.R., Read, D.J., and Varma, A.K. of a screening method for solubilizing ability and metal
(eds). London, UK: Academic Press, pp. 75–106. tolerance. Mycol Res 99: 987–993.
Prasad, M.N.V., and De Oliveira Freitas, H.M. (2003) Metal Sayer, J.A., Cotter-Howells, J.D., Watson, C., Hillier, S., and
hyperaccumulation in plants – biodiversity prospecting for Gadd, G.M. (1999) Lead mineral transformation by fungi.
phytoremediation technology. Electron J Biotechnol 6: Curr Biol 9: 691–694.
285–321. Smith, S.E., and Read, D.J. (1997) Mycorrhizal Symbiosis.
Sarret, G., Manceau, A., Spadini, L., Roux, J.R., Hazemann, London, UK: Academic Press.
J.L., Soldo, Y., et al. (1998a) Structural determination of Zn Sterflinger, K. (2000) Fungi as geologic agents. Geomicrobiol
and Pb binding sites in Penicillium chrysogenum cell walls J 17: 97–124.
by EXAFS spectroscopy. Environ Sci Technol 32: 1648– Tsekova, K., and Galabova, D. (2003) Phosphatase produc-
1655. tion and activity in copper (II) accumulating Rhizopus
Sarret, G., Manceau, A., Cuny, D., Van Haluwyn, C., delemar. Enzyme Microb Technol 33: 926–931.
Deruelle, S., Hazemann, J.-L., et al. (1998b) Mechanisms Tsekova, K., Galabova, D., Todorova, K., and Ilieva, S. (2002)
of lichen resistance to metallic pollution. Environ Sci Phosphatase activity and copper uptake during the growth
Technol 32: 3325–3330. of Aspergillus niger. Process Biochem 37: 753–758.
Sarret, G., Manceau, A., Spadini, L., Roux, J.R., Hazemann, Vulkan, R., Mingelgrin, U., Ben-Asher, J., and Frenkel, H.
J.L., Soldo, Y., et al. (1999) Structural determination of Pb (2002) Copper and zinc speciation in the solution of a
binding sites in Penicillium chrysogenum cell walls by soil-sludge mixture. J Environ Qual 13: 193–203.
EXAFS spectroscopy and solution chemistry. J Synch Whitelaw, M.A., Harden, T.J., and Helyar, K.R. (1999) Phos-
Radiation 6: 414–416. phate solubilization in solution culture by the soil fungus
Sarret, G., Saumitou-Laprade, P., Bert, V., Proux, O., Haze- Penicillium radicum. Soil Biol Biochem 31: 655–665.
mann, J.L., Traverse, A., et al. (2002) Forms of zinc accu- Zhang, P.C., Ryan, J.A., and Bryndzia, L.T. (1997) Pyromor-
mulated in the hyperaccumulator Arabidopsis halleri. Plant phite formation from goethite adsorbed lead. Environ Sci
Physiol 130: 1815–1826. Technol 31: 2673–2678.

© 2006 The Authors


Journal compilation © 2006 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology, 9, 308–321

You might also like