Download Lead resistance in micro-organisms

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Coordination complex wikipedia , lookup

Lead glass wikipedia , lookup

Metalloprotein wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Transcript
Microbiology (2014), 160, 12–25
Review
DOI 10.1099/mic.0.070284-0
Lead resistance in micro-organisms
Anna Jarosławiecka and Zofia Piotrowska-Seget
Correspondence
Department of Microbiology, University of Silesia, Jagiellońska Street 28, Katowice 40-032, Poland
Anna Jarosławiecka
[email protected]
Lead (Pb) is an element present in the environment that negatively affects all living organisms. To
diminish its high toxicity, micro-organisms have developed several mechanisms that allow them to
survive exposure to Pb(II). The main mechanisms of lead resistance involve adsorption by
extracellular polysaccharides, cell exclusion, sequestration as insoluble phosphates, and ion efflux
to the cell exterior. This review describes the various lead resistance mechanisms, and the
regulation of their expression by lead binding regulatory proteins. Special attention is given to the
Pbr system from Cupriavidus metallidurans CH34, which involves a unique mechanism
combining efflux and lead precipitation.
Introduction
Heavy metal resistance in bacteria results from their
primary contact with metals that appear naturally in the
environment. However, intensive human activity and
exploitation of natural deposits has led to the expansion
of metal-resistant micro-organisms (Bruins et al., 2000).
Apart from 13 other metals and metalloids, lead (Pb) is
considered to be one of the major pollutants (Sparks,
2005). Industrial activities, such as production of batteries
and pigments and metal smelting, as well as manufacture of
products such as lead arsenate insecticides or lead water
pipes are the main sources of Pb in the environment.
Natural processes including soil erosion, volcanic emission
and mobilization of Pb from minerals contribute only to a
minor degree to Pb pollution of the environment (Gadd,
2010; Gerba, 1996; Siegel, 2002; Spiro & Stigliani, 2002;
Yokel & Delistraty, 2003). The total Pb concentration in
industrial areas can reach up to 10 000 mg kg21, while the
average value in soils ranges from 10 to 100 mg kg21
(Akmal & Jianming, 2009; Schwab et al., 2005).
Additionally, the level of Pb(II) in industrial wastewaters
reaches 200–250 mg l21, whereas, according to accepted
quality standards, it should not exceed 0.05–0.10 mg l21
(Sag et al., 1995).
Pb(II) enters bacterial cells through the uptake pathways
for essential divalent metals such as Mn(II) (Laddaga et al.,
1985; Tynecka et al., 1981) and Zn(II) (Bruins et al., 2000;
Grass et al., 2002; Laddaga & Silver, 1985; Makui et al.,
2000). Pb is not known to be of any biological importance
and is toxic at very low concentrations (Bruins et al., 2000).
The toxicity of Pb depends upon its bioavailability. Since
most Pb(II) is bound to clay minerals and complex organic
molecules, the bioavailable fraction of Pb(II) to which
microbes are exposed may be rather low (Kotuby-Amacher
Abbreviations: CDF, cation-diffusion facilitator; EP, extracellular
polymer; MBD, metal binding domain; MFS, major facilitator superfamily;
MT, metallothionein; PPB, polyphosphate bodies.
12
et al., 1992; Singh et al., 1996). Thus, in laboratory
experiments it is necessary to choose the right medium
components to avoid the precipitation and complexation
of Pb(II), since this leads to an apparent decrease of Pb
toxicity (Roane, 1999; Sani et al., 2001).
Pb(II) toxicity occurs as a result of changes in the
conformation of nucleic acids and proteins, inhibition of
enzyme activity, disruption of membrane functions and
oxidative phosphorylation, as well as alterations of the
osmotic balance (Bruins et al., 2000; Vallee & Ulmer,
1972). Pb(II) also shows a stronger affinity for thiol and
oxygen groups than essential metals such as calcium and
zinc (Bruins et al., 2000). Despite the high toxicity of Pb,
many micro-organisms have evolved mechanisms that
enable them to survive Pb exposure (Fig. 1).
Micro-organisms resistant to Pb(II) have been isolated from
metal contaminated soils, industrial wastes, and from plants
growing on metal contaminated soil. Among these isolates the
following examples have been identified: the Gram-positive
bacteria Bacillus cereus, Arthrobacter sp. and Corynebacterium
sp.; the Gram-negative bacteria Pseudomonas marginalis,
Pseudomonas vesicularis and Enterobacter sp.; and fungi
Saccharomyces cerevisiae and Penicillium sp. Psf-2 (Chen &
Wang, 2007; Hasnain et al., 1993; Sun & Shao, 2007; Roane &
Kellogg, 1996; Trajanovska et al., 1997; Zanardini et al., 1997).
Some lead-resistant bacteria have been found to play a specific
role in the growth of lead-exposed plants. For example, the
endophyte Bacillus sp. MN3-4 increases Pb(II) accumulation
in Alnus firma, and Pseudomonas fluorescens G10 and
Mycobacterium sp. G16 promote plant growth and reduce
Pb toxicity in Brassica napus (Sheng et al., 2008; Shin et al.,
2012). Additionally, the activity of the lead-resistant bacteria
Streptomyces sp. and Ps. vesicularis has been noticed on the
marble facade of a cathedral, where it was manifested in the
form of red stains. These bacteria were capable of producing a
red and red–brown pigment, identified as the lead tetroxide,
minium (red lead, Pb3O4) (Zanardini et al., 1997).
Downloaded from www.microbiologyresearch.org by
070284 G 2014 SGM
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Printed in Great Britain
Lead resistance in micro-organisms
Pb2+
Pb2+
Pb2+
OM
IM
Pb2+
Pb2+
Pb2+
PbrA
Pb2+
Pb2+
Pb2+
Pb2+
ZntA
Pb2+
Pb2+
MT
CadA
Fig. 1. Selected mechanisms of cell protection against lead toxicity (based on Dopson et al., 2003, modified). Pb(II) could be
kept away from the micro-organisms through its precipitation as insoluble phosphates outside the cell (circles), adsorption on
extracellular polysaccharides (violet stars) or by polymers naturally occurring in the cell wall. After entering the cell through the
essential metal transporters, Pb(II) can be further inactivated by binding to the metallothioneins (MT), sequestered as insoluble
phosphates or removed from the cell via transporters such as CadA, ZntA or PbrA. OM, outer membrane; IM, inner membrane.
The purpose of this review is to summarize current
knowledge of the various lead resistance mechanisms in
micro-organisms.
Cell wall and exopolysaccharides
One of the mechanisms that micro-organisms use to avoid
the toxicity of metals of no biological function is to limit
their movement across the cell envelope (Bruins et al.,
2000). Early studies initiated on the interactions between
Pb(II) and the cell envelope in Micrococcus luteus and
Azotobacter sp. revealed that these ions are mostly present
in the cell wall and cell membrane, with a minor portion
present in the cytoplasmic fraction (Tornabene & Edwards,
1972; Tornabene & Peterson, 1975). The cell wall is a
natural barrier for Pb(II), since the functional groups of
several macromolecules are involved in binding this metal.
In Gram-negative bacteria, this role is played mainly by
lipopolysaccharide, a significant component of the outer
membrane. In Gram-positive bacteria, peptidoglycan
together with teichoic and teichuronic acids are responsible for Pb binding (Fig. 1) (Beveridge & Fyfe, 1985). This
protection mechanism is an uncontrolled process, and
specific resistance mechanisms become operative when the
cell envelope reaches its saturation point (Beveridge & Fyfe,
1985).
Çabuk et al. (2006) reported that hydroxyl and carboxyl
groups, as well as nitrogen based bio-ligands including
amide and sulfonamide, were involved in the binding of
Pb(II) by Bacillus sp. ATS- 2. In consequence, this strain was
able to bind 91.7 % of the Pb(II) added to growth medium at
50 mg l21 (Çabuk et al., 2006). In S. cerevisiae, amide and
phosphate groups were recognized to participate in the
http://mic.sgmjournals.org
immobilization of Pb(II) in the cell wall, whereas carbonyl,
phosphate, hydroxyl and amino groups had this role in
Pseudomonas aeruginosa ASU6a (Çabuk et al., 2007; Gabr
et al., 2008). In Synechococcus sp., molecules with amide,
amino, hydroxyl or carboxyl groups participate in the
binding of metal ions, including Pb(II) (Shen et al., 2008).
The strains mentioned above differed in their efficiency
when removing Pb(II) from metal solutions. Lead was
accumulated to 30, 64.5 and 123 mg Pb g21 of microbial
biomass for S. cerevisiae, Ps. aeruginosa and Synechococcus
sp., respectively (Çabuk et al., 2007; Gabr et al., 2008; Shen
et al., 2008). In E. coli, almost 97 % of Pb(II) was bound in
the cell membrane, while almost no Pb(II) was detected in
peptidoglycan (Kumar & Upreti, 2000).
Many micro-organisms synthesize extracellular polymers
(EPs) that bind cations of toxic metals, thus protecting
metal-sensitive and essential cellular components (Fig. 1)
(Bruins et al., 2000). The composition of EPs is very
complex, including proteins, humic acids, polysaccharides
and nucleic acids, which chelate metals with different
specificity and affinity (Pal & Paul, 2008; Guibaud et al.,
2003; Roane, 1999). Pb(II) binding by EPs has been
reported for Bacillus firmus, Pseudomonas sp. (Salehizadeh
& Shojaosadati, 2003), Cyanobacteria (Freire-Nordi et al.,
2005; Paperi et al., 2006), Halomonas sp. (Amoozegar et al.,
2012) and Paenibacillus jamilae (Perez et al., 2008). EP
produced by the latter strain is highly Pb(II) specific, and
adsorbs ten times more Pb(II) than other metals such as
Cd(II), Co(II), Ni(II), Zn(II) and Cu(II) (Aguilera et al.
2008; Morillo et al., 2006). This EP exhibits a characteristic
feature, namely a high content of uronic acids (28.29 %),
which are considered to play an important role in its Pb(II)
binding specificity. Immobilization of Pb(II) in the EP
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
13
A. Jarosławiecka and Z. Piotrowska-Seget
from Pa. jamilae was not influenced by the presence of
other metals in solution, and was strictly pH dependent
(optimal absorption was at pH 6), with the maximum
binding capacity being 303 mg (g EP)21 (Perez et al.,
2008). In the case of Ps. marginalis, isolated from metal
contaminated soil, EPs had the capacity to bind Pb(II) up
to 2.5 mM (0.3 mM of soluble lead in minimal medium,
pH 6). EPs were produced by this strain independently of
its exposure to the metals, and the bound Pb(II) was visible
as dark granules outside the cell, when observed through
transmission electron microscopy (Roane, 1999). EP with a
similar capacity to the EPs from Pa. jamilae [316 mg (g
EP)21] and a higher content of uronic acids (38 % of EP)
was found in Alteromonas macleodii subsp. Fijiensis (Loaëc
et al., 1997). Adsorption of Pb(II) to EPs as well as
accumulation of lead inside the cell has been also noted in
the cyanobacterium Gloeothece PC6909. After crossing the
membrane, Pb(II) was found to be bound to intracellular
inclusions, identified as polyphosphate bodies (PPB). The
primary role of PPB is to store energy and phosphorus;
however, they also contribute to the precipitation of metal
ions and to neutralization of their toxic effects. In addition,
accumulation of Pb(II) in the vacuolar structures of this
bacterium was also observed (Pereira et al., 2011).
Extra- and intracellular precipitation of Pb(II)
Pb(II) is known to react with several anions such as
chlorides, phosphates and hydroxyl ions to form insoluble
precipitates. Additionally, Pb(II) forms strong complexes
with organic compounds such as tryptone, cysteine, yeast
extract or succinic acid (Babich & Stotzky, 1979; Mayer &
Godwin, 2006). The precipitation of Pb(II) is used by
several micro-organisms to lower the concentration of free
Pb(II) by sequestering it in the form of phosphate salts
outside and inside the cell (Fig. 1). The chemical
composition of these precipitates varies between strains.
For example, Citrobacter freundii precipitates Pb(II) as an
extracellular phosphate (PbHPO4). Pb(II) precipitation in
this strain is presumably a consequence of the hydrolysis of
organic phosphorus by an acid phosphatase and the
concomitant precipitation of Pb(II) on the cell surface
(Aickin & Dean, 1977, 1979; Aickin et al.,1979).
Levinson et al. (1996) divided the process of intracellular
lead precipitation into three stages: (1) binding of metal to
the cell surface; (2) metabolism-dependent intracellular
uptake, and (3) final precipitation of Pb(II) as crystals of
lead phosphates. For instance, Staphylococcus aureus
precipitates Pb(II) inside the cell as lead phosphate
Pb3(PO4)2. This kind of protection allows this strain to
withstand a 600-fold higher dose of Pb(II) compared to a
sensitive strain (Levinson et al., 1996; Levinson & Mahler,
1998). The marine bacterium Vibrio harveyi is capable of
precipitating Pb(II) inside the cell in the form of an
unusual phosphate compound Pb9(PO4)6. This process is
regulated partially by quorum sensing. The exact mechanisms of this regulation are not known, but it has been
14
suggested that quorum sensing controls the availability of
inorganic phosphates (Mire et al., 2004). Precipitation of
Pb(II) in the form of Pb9(PO4)6 was also observed in
Providencia alcalifaciens 2EA as a result of phosphatase
activity that releases inorganic phosphates (Naik et al.,
2012). In Klebsiella aerogenes NCTC418 grown in phosphate limited medium, Pb(II) was precipitated inside the
cell as PbS (Aiking et al., 1985). Karnachuk et al. (2002)
showed that sulfate reducing bacteria used anglesite
(PbSO4) as the electron acceptor and reduced this
compound to poorly soluble galena (PbS). No toxic effect
of Pb(II) on these bacteria during the transformation
process was observed (Karnachuk et al., 2002). The
phosphate solubilizing Enterobacter cloacae (Park et al.,
2011) and Burkholderia cepacia immobilize Pb(II) in the
form of pyromorphite Pb5(PO4)3Cl (Templeton et al.,
2003). Lead precipitation has also been found in other
strains, such as Shewanella putrefaciens CN32, Ps. fluorescens and the endophyte Bacillus sp. MN 3-4; however, the
chemical compositions of these precipitates have not
been determined (al-Aoukaty et al., 1991; Shin et al.,
2012; Smeaton et al., 2009). Intracellular sequestration of
Pb(II) by micro-organisms was also demonstrated by in
situ experiments. Precipitates containing lead were present
in different soil bacterial strains and identified as PPB
(Perdrial et al., 2008). In the cyanobacterium Synechocystis
PCC 6803 exposed to Pb(II), an increase of PPB in
number and size was reported (Arunakumara & Xuecheng,
2009).
Binding of Pb(II) by specific proteins
Metallothioneins (MTs) are frequently referred to as
proteins involved in protecting cells against toxic metals;
however, their primary role is in zinc homeostasis (Fig. 1)
(Blindauer, 2011). MTs, first discovered in Synechococcus
PCC 7942, are encoded by the smt locus, which consists of
two divergently transcribed genes smtA and smtB. smtA
encodes a class II MT, and the product of smtB represses the
transcription of smtA. SmtA is involved in zinc homeostasis,
as the deletion of smtA in Synechococcus PCC 7942 led to its
hypersensitivity to Zn(II) (Blindauer, 2011). In addition to
Zn(II), Pb(II) is one of the few metalions that is capable of
switching on the expression of smtA. Apart from
Synechococcus PCC 7942, the induction of smt genes and
biosynthesis of MT in the presence of Pb(II) has been
detected in B. cereus, Streptomyces sp., Salmonella choleraesuis 4A and Proteus penneri GM-10 (Huckle et al., 1993;
Murthy et al., 2011; Naik et al., 2012; Rifaat et al., 2009).
Another gene (bmtA) encoding a Pb(II) binding metallothionein has been identified in Ps. aeruginosa WI-1 (Naik
et al., 2012). Additionally, it has been suggested that a
metallothionein-like protein binds Pb(II) in Bacillus megaterium (Roane, 1999).
Extracellular enzymes, such as superoxide dismutase from
Streptomyces subrutilis, can also be engaged in Pb(II)
binding (So et al., 2001).
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
Biotransformation of lead compounds
A number of micro-organisms inhabiting soil and water can
transform inorganic and organic lead compounds into
volatile forms, which diminishes their toxic effect (Thayer,
2002). This effect has been observed in lake sediments
(Silverberg et al., 1977; Wong et al., 1975). Nevertheless, the
contribution of microbes to this process requires further
investigation, since methyl lead compounds are also present
in the environment as a result of pollution (Gadd, 2010;
Thayer, 2002). The biomethylation of Pb(II) has been
reported to be catalysed by Pseudomonas sp. (Walton et al.,
1988; Wong et al., 1975), Alcaligenes sp. (Wong et al., 1975),
Acinetobacter sp. (Wong et al., 1975), Flavobacterium sp.
(Wong et al., 1975), Aeromonas sp. (Wong et al., 1975) and
Bacillus sp. (Walton et al., 1988). In general, two forms of
volatile methylated lead have been identified as products of
microbial activity: tetramethyl (Me4Pb) and trimethyl
(Me3Pb) lead. For example, Aeromonas sp., isolated from
Lake Ontario, transformed lead acetate into tetramethyl lead
(Me4Pb). This reaction was observed both in the lake water as
well as in laboratory cultures (Silverberg et al., 1976). Arctic
marine bacteria cultured under polar conditions were capable
of converting Pb(II) into trimethyl lead (Me3Pb) (Pongratz &
Heumann, 1999) while Pseudomonas sp., Acinetobacter sp.,
Flavobacterium sp. and Aeromonas sp. transformed lead
nitrate or trimethyl lead acetate into tetramethyl lead (Me4Pb)
(Hughes & Poole, 1989; Reisinger et al., 1981; Thayer &
Brinckman, 1982; Wong et al., 1975).
Binding of Pb(II) by siderophores
Siderophores are small, high-affinity iron chelating compounds secreted by various micro-organisms. Apart from
iron, siderophores may interact with other metals outside
of the cell (Saha et al., 2012; Schalk et al., 2011). They have
been implicated in the complexation of Pb(II) by Ps.
aeruginosa 4EA and by endophytic Pseudomonas putida
KNP9, isolated from mung bean (Phaseolus vulgaris) (Naik
& Dubey, 2011; Tripathi et al., 2005). Siderophores
produced by these two strains have been identified as
yellow–green, fluorescent pyoverdines, which are characteristic for Pseudomonas. The pyoverdines produced by
the plant growth promoting strain Ps. putida KNP9
reduced the concentration of Pb(II) in mung bean roots
by 93 % and in shoots by 56 % (Tripathi et al., 2005).
Studies by Naik & Dubey (2011) revealed that production
of siderophores by Ps. aeruginosa 4EA may double in
response to Pb(II) exposure. In Ps. aeruginosa PAO1,
pyochelin also contributed to Pb(II) binding, apparently
with a higher affinity than pyoverdine (Braud et al., 2010).
Efflux systems
Ions that escape extra- and intracellular binding can be
excluded from the cell through the activity of efflux
systems, the most effective mechanisms providing resistance to toxic metals (Bruins et al., 2000). In this
mechanism, proteins from the superfamily of P-type
http://mic.sgmjournals.org
ATPases transport metal cations across the cytoplasmic
membrane using ATP as the energy source (Apell, 2004).
Efflux of Pb(II) is mediated by P-type ATPases from the
PIB family, which are also involved in the transport of
Zn(II) and Cd(II). However, other PIB ATPases also exist
that are specific for other metals such as Cu(I) (Argüello,
2003). PIB pumps have been widely recognized in most
bacterial and archaeal species (Coombs & Barkay, 2005).
Among them, CadA from S. aureus, ZntA from E. coli,
CadA2 from Ps. putida KT2440 and PbrA from Cupriavidus
metallidurans CH34 are known to transport Pb(II), as well
as Cd(II) and Zn(II), to the periplasm (Fig. 1) (Argüello,
2003; Hynninen et al., 2009; Leedjärv et al., 2008, Rensing
et al., 1998). Other transmembrane metal transporters
belong to the CBA family of transporters, and the cationdiffusion facilitator (CDF) family. The former are present
in Gram-negative bacteria, and are chemiosmotic antiporters transferring metals from the cytoplasm to the
periplasm. CDF transporters are chemiosmotic ion–proton
antiporters and also transport ions from the cytoplasm to
the periplasm (Nies, 2003). So far, only CzcCBA1 from Ps.
putida KT2440, a CBA transporter, has been recognized to
participate in the export of Pb(II) from the cell, while none
of the transporters from the CDF group has yet been
shown to have this activity (Hynninen, 2010).
Metalloregulatory proteins and Pb(II) sensing
MerR family regulators. Bacteria can control the intra-
cellular concentration of Cd(II), Zn(II), Co(II), Cu(I), Ag(I),
Au(I), Hg(II) and Pb(II) through metalloregulatory proteins
belonging to the MerR family (Brown et al., 2003). These
proteins regulate the expression of genes encoding defence
systems active against toxic metals or elevated levels of
microelements. Differences within the C-terminal ligand
binding domains of MerR family regulators allow them to
recognize a specific metal (Chen & He, 2008). One of the
MerR-like proteins is PbrR from C. metallidurans CH34
(CmPbrR). This protein is encoded by pbrR, located on
pMOL30, and it regulates expression of the pbr lead
resistance genes and, consequently, synthesis of the lead
efflux pump (Borremans et al., 2001; Taghavi et al., 2009). It
has been shown that CmPbrR is highly specific towards
Pb(II), with an almost 1000-fold specificity over other metals
such as Hg(II), Cd(II), Zn(II), Co(II), Ni(II), Co(II) and
Ag(I) (Chen et al., 2005). CmPbrR and its two homologues
from the same micro-organism (chromosomally encoded
CmPbrR2, previously PbrR691, and CmPbrR3) are the only
known lead specific regulatory proteins so far identified
(Chen & He, 2008; Taghavi et al., 2009). The molecular
mechanism of selectivity and recognition of Pb(II) by
CmPbrR has been studied for CmPbrR2. CmPbrR2 has
three highly conserved cysteine residues: Cys78, Cys113 and
Cys122 (Fig. 2a), which have been proved to be involved in
Pb(II) binding; however, the existence of a fourth ligand
cannot be excluded (Chen & He, 2008). The coordination of
Pb(II) is specific, as it possesses a stereochemical lone pair of
electrons on one side, not engaged in binding to ligands. As a
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
15
A. Jarosławiecka and Z. Piotrowska-Seget
(a)
DNA binding domain
α1
β1
CmMerR
CmPbrR
CmPbrR2
CmMerR
CmPbrR
CmPbrR2
α2
β2
β3
α3
14
MENNLENLTIGVFAKAAGVNVETIRFYQRKGLLLEPDKPYGSIRRYGEADVTRVRFVKSA
-----MNIQIGELAKRTACPVVTIRFYEQEGLLPPPGRSRGNFRLYGEEHVERLQFIRHC
------MMRIGELGKKADCLVQTVRFYESEGLLPEPARSEGNFRLYDEVHLQRLLFIRRC
: ** : * :
* *: ** *: : *** * : * . : * * * . : * : *: : .
α4
α5
78, 79
113
QRLGFSLDEIAELLRLEDG--THCEEASSLAEHKLKDVREKMADLARMEAVLSELVCACH
RSLDMPLSDVRTLLSYRKRPDQDCGEVNMLLDEHIRQVESRIGALLELKHHLVELREACS
RAKDMTLDEIRQLLNLRDRPELGCGEVNALVDAHIAQVRTKMKELRALERELMDLRRSCD
* * .. * : :: :*. :: * :: * :* :*
** ..
:
: * ::
2-turn α-helix
122
132
CmMerR ARRGNVS------------CPLIASL---------------CmPbrR GARPAQS------------CGILQGLSDCVCDTRGTTAHPSD
CmPbrR2 SARTSRQGGASLAGSAMPECGILNSLAEPA-----------* :: .*
. *
.
(b)
CadC
CmtRMtb
CmtRSc
SmtB
AztR
α1
N
7
11
---MK-----------KKDTCEIFCYDEEKVNRI-QGDLQTVDISGVSQILKAIADENRA
------------------------------------MLTCEMRESALARLGRALADPTRC
------------------------------------MLTLAADIDVLARFGRALADPIRC
-------MTKPVLQDGETVVCQGTHA----AIASELQAIAPEVAQSLAEFFAVLADPNRL
MNKHKKKQDLDLIQSSDTPTCDTHLVHLDNVRSSQAQILPTDKAQQMAEIFGVLADTNRI
. :: .:
. : ** *
DNA binding domain
α3
CadC
CmtRMtb
CmtRSc
SmtB
AztR
α2
αR
β1
β2
58 60
57 61
KITYALCQDEELCVCDIANILG-VTIANASHHLRTLYKQGVVNFRKEGKLALYSLGDEHI
RILVALL-DGVCYPGQLAAHLG-LTRSNVSNHLSCLRGCGLVVATYEGRQVRYALADSHL
RILLALR-QAPAYPADLADSLG-ISRTRLSNHLACLRDCGLVVTVPDGRRTRYELADERL
RLLSLLA-RSELCVGDLAQAIG-VSESAVSHQLRSLRNLRLVSYRKQGRHVYYQLQDHHI
RLLSALA-SSELCVCDLAALTKMNSESAVCHQLRLLKAMRLVSYRREGRNVYYSLADSHV
::
*
::*
: :
. .:* *
:*
:*: . * * * ::
α5
CadC
CmtRMtb
CmtRSc
SmtB
AztR
102
RQIMMIALAHKKEVKVNV-----------------ARALGELVQVVLAVDTDQPCVAERAASGEAVEMTGS
GHALDDLRAAVVAVEADRTCADADEKGCC------VALYQNALD------HLQECR--------------INLYRSLVE------NNTYATGTG-----------α6
C
Fig. 2. Amino acid sequence alignment of metalloregulators involved in Pb(II) sensing (generated using CLUSTAL Omega 1.2.0).
(a) Metalloregulators from the MerR family: CmMerR (Tn501), CmPbrR2 and CmPbrR from C. metallidurans CH34. (b)
Metalloregulators from SmtB/ArsR family: SmtB from Synechococcus sp. strain PCC 7942, CadC from S. aureus pI158, CmtR
from M. tuberculosis, CmtR from St. coelicolor, AztR from Anabaena PCC 7120. The cysteine residues involved in the Pb(II)
sensing in metalloregulators and corresponding residues are marked in yellow. Amino- and carboxyl-terminal domain, secondary
structures and DNA binding domain are marked in the figures. Description in the text (Cavet et al., 2003; Hobman et al., 2012;
Pennella & Giedroc, 2005).
consequence, Pb(II) forms a hemidirected geometry with all
the ligands on one side of the metal and with the pair of
electrons on the other. CmPbrR2 most likely uses
hemidirected geometry to recognize Pb(II) over other ions
16
(Chen et al., 2007; Chen & He, 2008). This unique feature has
been utilized in the construction of lead specific indicators
for monitoring the intracellular presence of Pb(II) in human
cells, and biosensors sensing Pb(II) contamination in the
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
environment (Chakraborty et al., 2008; Chen et al., 2005;
Chiu & Yang, 2012; Corbisier et al., 1999).
Acidithiobacillus ferrooxidans ATCC 23270 (Zheng et al.,
2012).
In recent studies, residues Cys14, Cys79 and Cys134 have
been identified in CmPbrR to be involved in the sensing of
Pb(II), and in promoter (PpbraA) activation (Fig. 2a)
(Hobman et al., 2012). Cys14 is situated in the helix–turn–
helix DNA binding domain of CmPbrR and may play an
essential role in the PbrR–DNA interaction. In addition,
Cys79, which is present in all the divalent metal ion
responsive MerR regulators examined so far, is necessary
for the detection of +2 charged ions over +1 ones. In
contrast, Cys134 is unique and is not found in other
characterized MerR regulators (Hobman et al., 2012).
Upon lead binding, the PbrR transcriptional regulator
induces the expression of genes encoding Pb(II) specific
transporters to pump out metal ions out of the cell (Chen
et al., 2007).
AztA from the cyanobacterium Anabaena PCC 7120 is
a Zn(II)/Pb(II)/Cd(II) translocating PIB-type ATPase,
expression of which is regulated by an SmtB/ArsR family
repressor, AztR (Liu et al., 2005). AztR, which has high
sequence similarity to SmtB, differs slightly in its structure
since it lacks the a5 zinc specific binding site, which is
essential for the allosteric regulation of DNA binding by
SmtB and ZiaR (VanZile et al., 2002; Turner et al., 1996;
Thelwell et al., 1998). AztR possesses only one metal
binding site, a3N, for Zn(II), Cd(II) and Pb(II), which is
similar to the S3(N9O) metal site from CadC in S. aureus
(Fig. 2) (Liu et al., 2005).
Another transcriptional regulator from the MerR family is
ZntR, which regulates the expression of the zntA gene
encoding ZntA. This PIB ATPase transports Pb(II) as well
as Zn(II) and Cd(II) ions out of the cell in E. coli (Binet &
Poole, 2000; Brocklehurst et al., 1999). Binet & Poole
(2000) demonstrated that all these metal ions can be
inducers for zntAR, and the Cd(II)- and Pb(II)-dependent
regulation of zntA is also mediated by ZntR. The
interactions between Zn(II) and ZntR are well known,
but ZntR interactions with Pb(II) and Cd(II) remain
poorly studied (Binet & Poole, 2000).
SmtB/ArsR family regulators. CadC from S. aureus is a
regulatory protein from the SmtB/ArsR family, which
regulates the expression of cadAC, encoding a membrane
bound Cd(II)/Zn(II)/Pb(II) specific PIB-type ATPase,
CadA (Busenlehner et al., 2003; Yoon et al., 1991). CadC
responds to the metal ions in the following order: Pb(II) .
Cd(II) . Zn(II) (Rensing et al., 1998). In the CadC
structure, an a3N metal binding site containing three
cysteine residues, Cys7, Cys58 and Cys60, ensures selection
of the correct ion (Fig. 2b) (Sun et al., 2001). CadC binds
Pb(II) through a trigonal coordination complex, with
cysteine thiolate ligands derived from the N-terminal
domain (Cys7/11) and a pair of cysteines in the a4 helix
(Cys58/60) (Fig. 2) (Busenlehner & Giedroc, 2006).
Another protein that belongs to the SmtB/ArsR family is
CmtR from Mycobacterium tuberculosis (CmtRMtb). CmtR
has also been found to sense Pb(II) in Mycobacterium
smegmatis (Cavet et al., 2003). In contrast to other SmtB/
ArsR Cd(II)/Pb(II) sensing proteins, CmtRMtb possesses a
unique pair of a4C metal binding sites, formed from the a4
helix and a C-terminal tail (Banci et al., 2007). Pb(II) is
predicted to be bound to Cys57 and Cys61 in the helix a4,
and to Cys102 in the C-terminal region (Fig. 2). A CmtR
sensor with an a4N site sensing Pb(II) and Cd(II) has been
found in Streptomyces coelicolor (Wang et al., 2010).
Recently, another CmtR transcriptional regulator responsive to Cd(II) and Pb(II) has also been reported in
http://mic.sgmjournals.org
Other regulators. ZraSR (formerly HydH/G) is a two-
component regulatory system, which regulates the
expression of zraP, encoding ZraP, a periplasmic Zn(II)
binding protein in E. coli involved in zinc homeostasis
(Noll et al., 1998). ZraS activates ZraR (a cytoplasmic
response regulator) by phosphorylation. Although ZraP
was not reported to bind Pb(II), its regulators ZraSR are
induced by high concentrations of Zn(II) and Pb(II) (Lee
et al., 2005; Leonhartsberger et al., 2001).
The C. metallidurans CH34 Pbr efflux system
A Pbr efflux system which removes Pb(II) from the cell was
first reported by Borremans et al. (2001) in C. metallidurans
CH34 (formerly known as Alcaligenes eutrophus, Ralstonia
metallidurans and Wautersia metallidurans). This bacterium
is resistant to 20 heavy metals and is used as a model microorganism in studies of metal resistance and metal tolerance
(Borremans et al., 2001; Janssen et al., 2010). The Pbr efflux
is encoded by pbr, located within a mer-pbr-czc island on
pMOL30 (Monchy et al. 2007). Seven genes of the pbr locus
(pbrUTRABCD) are organized into two divergently oriented
transcription units, pbrUTR and pbrABCD (Hynninen et al.,
2009). The two pbr promoters are both regulated by the
transcriptional activator PbrR (Fig. 3) (Permina et al., 2006;
Hynninen et al., 2009). So far, pbr has only been found on
plasmids in several lead-resistant bacteria, frequently with
pbrU*, pbrT, pbrR
p/o
pbrA, pbrB, pbrC, pbrD
PbrR
Pb2+
Fig. 3. Regulation of pbr by PbrR. In the presence of Pb(II), PbrR binds
to the pbr promoter (p/o) and initiates its transcription in two opposite
directions. pbrU is inactivated by the insertion with transposone
TnCme2 (marked with asterisk). (Based on Hynninen, 2010).
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
17
A. Jarosławiecka and Z. Piotrowska-Seget
(a) C. metallidurans CH34
pbrT
pbrR
pbrA
pbrBC
pbrD
(b) A. faecalis NCIB8687, Sh. frigidimarina NCIMB400, K. pneumoniae NTUH-K2044, K. pneumoniae CG43
pbrR
pbrA
pbrBC
pbrR
pbrA
pbrB
(c) R. picketti 12D and 12J
pbrT
(d) H. arsenicoxydans
pbrB
pbrR
pbrC
pbrT
pbrA
Fig. 4. The pbr genes from various bacteria (based on Hynninen et al., 2009).
some modifications (Hynninen et al., 2009; Permina et al.,
2006). In comparison to C. metallidurans CH34, in all other
known lead-resistant bacteria the pbrD gene is missing,
whereas some of them lack pbrT (Alcaligenes faecalis
NCIB8687, Shewanella frigidimarina NCIMB400, Klebsiella
pneumoniae NTUH-K2044, K. pneumoniae CG43) or
pbrC (Ralstonia picketti 12D and 12J). Additionally, in
Herminiimonas arsenicoxydans, pbrC and pbrB are separated,
and pbrT is next to pbrC (Fig. 4) (Chen et al., 2004;
Hyninnen, 2009). The pbrU gene, located downstream of
pbrT in C. metallidurans CH34, is inactivated by the
insertion of the TnCme2 transposon (Monchy et al., 2007).
Proteins encoded by pbr and the mechanism of Pb(II)
resistance
The group of proteins encoded by pbr in C. metallidurans
CH34 includes: (1) PbrT, the membrane protein; (2) PbrA, a
PIB-type ATPase, the main efflux transporter; (3) PbrB,
undecaprenyl pyrophosphate phosphatase, an integral
membrane protein; (4) a putative signal peptidase PbrC;
(5) PbrD, a putative intracellular lead binding protein, and
(6) PbrR, a positive regulator (Fig. 5) (Borremans et al.,
2001; Hynninen et al., 2009; Monchy et al., 2007). PbrT is a
lead uptake permease belonging to the iron/lead transporter
superfamily (Debut et al., 2006), which participates (along
with other transporters for essential metals) in the influx of
Pb(II). In the cytoplasm, Pb(II) is most likely bound by the
PbrD protein, reducing its toxic effect, and the expression of
pbrA is activated. PbrD might play the role of a chaperone
for Pb(II) inside the cell. It possesses a potential metal
binding motif, rich in cysteine residues (Cys-7X-Cys-Cys7X-Cys-7X-His-14X-Cys) with a large number of proline
18
and serine residues. Although PbrD is not required for full
lead resistance, cells deficient in this protein show a
decreased accumulation of Pb(II) (Borremans et al., 2001).
The PbrA protein, supplemented by other PIB ATPases,
CadA and ZntA, exports Pb(II), Cd(II) and Zn(II) ions to
the periplasm. Then, inorganic phosphate groups released
by PbrB precipitate lead, preventing the re-entry of Pb(II)
into the cell (Fig. 5). This mechanism is the only system
known so far which is capable of combining efflux and ion
precipitation (Hynninen et al., 2009). Although the Pbr
efflux system was first predicted to be Pb(II) specific,
Hynninen et al. (2009) proved that it also participates in the
protection of the cell against Cd(II) and Zn(II), whereas the
specific mechanism of Pb(II) resistance requires the
cooperation of PbrA and PbrB (Borremans et al., 2001;
Hynninen et al., 2009). The roles of PbrT, PbrC and PbrD
are being challenged, since their absence does not impair the
overall ability of this mechanism to neutralize toxic ions
(Borremans et al., 2001; Hynninen et al. 2009).
The mechanism of Pb(II) removal from the periplasm to
the cell exterior in C. metallidurans CH34 is not known,
but may involve the CzcCBA efflux system (Hynninen et al.,
2009). CzcP (a P1B4-type ATPase, first assigned to the PIB
family of Pb(II)/Cd(II)/Zn(II) efflux ATPases) was also
suggested to be a transporter involved in the further efflux
of Pb(II) from the cell. However, additional research
showed that it does not participate in this activity (Monchy
et al., 2007; Scherer & Nies, 2009).
In C. metallidurans CH34, the pbrR2cadApbrC2 gene
cluster located on chromosome 1, and zntA located on
chromosome 2, cooperate with pbr in providing Pb(II)
resistance (Monsieurs et al., 2011; Taghavi et al., 2009).
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
Pb2+
OM
Pb2+--P
Pb2+--P
Pb2+
Pb2+
PbrA
ZntA
Pb2+
CadA
P
Pb2+--P
Periplasm
PP PP
PbrB
ATP
Pb2+
ADP
+Pi
Pb2+
PbrD
PbrA, a member of the PIB ATPase family, is not Pb(II)
specific, and also transports Zn(II) and Cd(II) to the
periplasm (Hynninen et al., 2009). Efflux through PbrA can
compensate for the transport of Pb(II) into the cell via the
PbrT protein (Borremans et al., 2001). Detailed information about the PbrA protein is included in the next section
concerning PIB ATPases.
PbrB is a membrane protein, which has been identified
as a C55 lipid phosphatase, with six predicted transmembrane domains and a periplasmically located active site.
Presumably, this protein does not participate in the cell
wall synthesis as other C55 lipid phosphatases do, being
only a part of the Pbr efflux system. The other
phosphatases may contribute to the release of phosphate
groups. However, it is rather unlikely that the expression of
genes responsible for their synthesis is induced by Pb(II)
(Hynninen et al., 2009). Inactivation of pbrB and the
absence of PbrB results in lead sensitivity in C. metallidurans CH34 (Borremans et al., 2001; Taghavi et al., 2009).
The mmfB gene, located on the Tn6048 transposon, shows
significant similarity to pbrB and hence might compensate
for the loss of this gene (Monsieurs et al., 2011). PbrB is
fused with the PbrC protein, a putative signal peptidase;
however, PbrC is not essential for the functioning of the Pbr
efflux system. It is assumed that its role can be substituted by
other peptidases, or that peptidase activity is not essential for
the correct functioning of PbrB. PbrC might be important
for the maturation or activity of PbrB in the periplasm
(Hobman et al., 2012; Hynninen et al., 2009).
Although pbrU is not expressed in C. metallidurans CH34, it
has been suggested that this gene encodes a permease,
located in the inner membrane and belonging to the major
facilitator superfamily (MFS) (Monchy et al., 2007; Taghavi
et al., 2009). Apart from pbrU, eight other genes encoding
MFS proteins have been found to be upregulated by Pb(II).
Interestingly, the most overexpressed gene encoding an
MFS is located inside the cluster responsible for the
biosynthesis of a siderophore. This fact suggests that there
is cross-talk between the responses to lead and iron
(Taghavi et al., 2009).
http://mic.sgmjournals.org
IM
Fig. 5. The lead transport system in C.
metallidurans CH34 (based on Hynninen et al.,
2010, modified). Pb(II), entering the cell through
the Zn(II) and Mn(II) transporters, is actively
pumped out by PbrA to the periplasmic space
and then precipitated with phosphates (P)
released by the PbrB protein. PbrA is supported
in this role by CadA and ZntA proteins. PbrD
is a putative intracellular Pb(II) binding protein
(Borremans et al., 2001; Hynninen et al., 2009).
OM, outer membrane; IM, inner membrane.
In C. metallidurans CH34, Pb(II) also induces general stress
mechanisms via alternative sigma factors such as s24/rpoK,
s32/rpoH and s28/fliA (Grosse et al., 2007).
PIB ATPases PbrA, CadA, CadA2 and ZntA
As mentioned above, the PIB ATPase PbrA participates in
Pb(II) efflux in C. metallidurans CH34 (Hynninen et al.,
2009). This transporter possesses two metal binding
domains (MBD): the N-terminal cytoplasmic domain and
a domain located in the transmembrane region (Argüello,
2003; Borremans et al., 2001; Dutta et al., 2006, 2007).
Primarily, the N-terminal domain was suggested to be
involved in Pb(II) binding, as its amino acid sequence differs
from the equivalent one in ZntA from E. coli and CadA from
S. aureus (Borremans et al., 2001; Hynninen et al., 2009).
However, mutations in the N-terminal MBD in ZntA from
E. coli proved that this part of the transporter is not
necessary for metal recognition; hence a similar conclusion
was drawn for PbrA from C. metallidurans CH34 (Mitra &
Sharma, 2001; Hou et al., 2001; Hynninen et al., 2009).
Instead, the transmembrane region has been proposed as the
area responsible for the recognition of a specific metal,
which has been confirmed experimentally (Argüello, 2003;
Dutta et al., 2006, 2007; Liu et al., 2006). The sequence of
conserved amino acids in the transmembrane MBD from
PbrA in C. metallidurans CH34 is the same as in ZntA from
E. coli, CadA from S. aureus, CadA2 from Ps. putida and
ZntA from C. metallidurans CH34. The PbrA transporter is
not Pb(II) specific and is also capable of transporting Cd(II)
and Zn(II) to the periplasm (Argüello, 2003; Borremans
et al., 2001; Hynninen et al., 2009; Leedjärv et al., 2008;
Rensing et al., 1998).
Efflux of Pb(II) has also been confirmed for CadA from S.
aureus (a PIB-type ATPase) and its homologue ZntA from
E. coli (Rensing et al., 1997, 1999). The CadA protein is
encoded by the cadAC genes, located on plasmid pI258.
This transporter, driven by ATP hydrolysis, functions as an
efflux pump, transferring ions from the cytoplasm to the
periplasm. Low level resistance to the metal is provided by
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
19
A. Jarosławiecka and Z. Piotrowska-Seget
the pump itself; however, the full efflux function is
achieved in cooperation with CadC, which negatively
regulates expression of the cadA gene (Nies, 1992; Rensing
et al., 1999; Silver & Phung, 2005).
CadA2 is the main efflux transporter for Cd(II) and Pb(II)
in Ps. putida KT2440. Surprisingly, it might be supported
in this role by CzcCBA1, which so far has only been known
for transporting Zn(II), Cd(II) and Co(II) from the
periplasm to the cell exterior in C. metallidurans CH34
(Leedjärv et al., 2008). It has also been postulated that Ps.
putida KT2440 might possess an additional mechanism for
lead resistance, as mutations in genes encoding lead
resistance transporters only slightly impaired the ability
of this bacterium to resist the toxicity of Pb(II) (Hynninen,
2010). In Ps. putida KT2440, PAP2 phosphatase encoded
by pp4813 is present, and can substitute for the activity of
PbrB in C. metallidurans CH34. It is hypothesized that the
novel family of transporters exemplified by MrdH in this
strain may also participate in lead resistance (Haritha et al.,
2009; Hynninen, 2010).
ZntA is a Zn(II), Pb(II) and Cd(II) specific PIB ATPase that
transports these ions in E. coli. The expression of the zntA
gene is regulated by ZntR, a transcriptional regulator from
the MerR family (Brocklehurst et al., 1999). The N-terminal
domain of the ZntA ATPase possesses a highly conserved
GXXCXXC motif, characteristic for P1-type ATPases, as well
as an additional CCCDGAC sequence (Liu et al., 2005). Liu
et al. (2005) proved that the first 46 residues of the Nterminal domain together with the CCCDGAC sequence
play a significant role in the binding of Pb(II), while having
no influence on either Zn(II) or Cd(II) binding (Liu et al.,
2005). The transport of specific metal ions by this pump is
further determined by structural features in the transmembrane domain (Mitra & Sharma, 2001).
Summing up, it has to be pointed out that although ZntA,
CadA and their close homologues function as transporters
involved in bacterial resistance to Pb(II) and Cd(II), their
primary physiological role is participation in zinc homeostasis (Nies, 1992; Rensing et al., 1998; Raimunda et al.,
2012).
Resistance of fungi to Pb(II)
Little is known about lead resistance mechanisms in fungi,
but even low concentrations of this metal can affect their
growth and reduce biomass. The growth of S. cerevisiae is
disturbed at 0.05 mM Pb(II). This concentration of metal
reduces the biomass by 60 % (Donmez & Aksu, 1999).
Pb(II) induces a variety of changes in S. cerevisiae such as
the lengthening the lag phase and slowing growth,
decreasing the DNA/RNA ratio and damaging DNA. It
inhibits the assimilation of ammonium ions, as well as
reducing viability and metabolic activity (Bussche &
Soares, 2011; Chen & Wang, 2007; Van der Heggen et al.,
2010; Sakamoto et al., 2010; Suh et al., 1999a; Soares et al.,
2002, 2003; Yuan & Tang, 1999). Pb(II) leads to the death
20
of yeast cells, caused by reactive oxygen species which
trigger apoptosis (Bussche & Soares, 2011).
Fungi have developed different strategies to ameliorate the
toxicity of lead. In Aureobasidium pullulans, EPs bind more
than 90 % of the Pb(II) that it was exposed to. Removal of
EPs enables Pb(II) to penetrate the cell wall and the cell
membrane and so enter the cytoplasm (Suh et al., 1999b,
2001). In the marine fungus Corollospora lacera, 93 % of all
lead was located extracellularly in the mycelia, while 1.7–
5.5 % was found inside the cell. Penicillium sp. Psf-2,
isolated from the Pacific sediment, was able to withstand
24 mM Pb(NO3)2 in a liquid culture. Lead deposits have
been detected in the cytoplasm and vacuole, as well as on
the outer layer of the cell wall of this fungus. Intriguingly,
these precipitates seem to have been composed of pure lead
only (Sun & Shao, 2007; Taboski et al., 2005).
Interactions between S. cerevisiae and Pb(II) are more
complex and three phases can be distinguished in the
reaction of the cell (Chen & Wang, 2007). In the first
phase, Pb(II) is quickly adsorbed on the surface of the cell
and then passively transported through the cell wall. This
phase is rapid, lasts about 3–5 min and is independent of
metabolism. Accumulation of Pb(II) causes changes on the
cell wall’s surface, which becomes rough (Suh et al., 1999a).
In the next step, Pb(II) crosses the cell membrane into the
cytoplasm, though the transport mechanism has not been
characterized. Pb(II) penetrates into the cells easily. This
suggests that the yeast cell wall is not a significant barrier to
Pb(II) and lead binding by functional groups within the
cell wall is poor. The yeast plasma membrane may control
to a certain degree the transport of Pb(II) to the cytoplasm,
since it takes close to an hour for these ions to cross the cell
membrane. However, it is still unknown how the
membrane regulates the flow of Pb(II). Finally, Pb(II) is
accumulated in the cytoplasm. This phase is independent
of metabolism (Suh et al., 1998, 1999a). In the cytoplasm,
glutathione (GSH) has been reported to be involved in the
neutralization of the toxicity of other heavy metals, such as
Cd(II) and Se(II) (Gharieb & Gadd, 2004). Van der Heggen
et al. (2010) revealed that, in the cytoplasm of S. cerevisiae,
Pb(II) is conjugated with glutathione, and the lead–
glutathione complex is transported to the vacuole, where
it is isolated from the rest of the cell. This complex is
transported through the vacuole membrane by yeast
cadmium factor 1, which is an ATP binding cassette
transporter. CAD2, a putative copper transporting ATPase
from S. cerevisiae, seems to have a role in the efflux of
Pb(II) (Li et al., 1997; Prévéral et al., 2006; Shiraishi et al.,
2000; Song et al., 2003; Szczypka et al., 1994). A protective
role of GSH against lead toxicity has also been observed in
Schizosaccharomyces pombe. Mutants of this strain deficient
in glutathione lost the ability to bind Cd(II) and became
more sensitive to Pb(II), Cu(II) and Zn(II) (Coblenz &
Wolf, 1994).
Fungi can also transform Pb(II) into pyromorphite, the
most stable lead mineral, indicating their significant role in
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
the lead cycle in the environment (Clipson & Gleeson,
2012; Rhee et al., 2012). In turn, Aspergillus niger is capable
of solubilizing this mineral and subsequently immobilizing
Pb(II) as lead oxalates (Sayer et al., 1999).
Potential of lead-resistant micro-organisms in
environmental biotechnology
The increasing concentration of such hazardous heavy
metals as Pb(II) in the environment has stimulated scientists
to search for new possible ways of its removal or
neutralization. As mentioned above, micro-organisms show
a wide range of mechanisms for ameliorating the toxicity of
Pb(II) via its extra- and intracellular precipitation, adsorption on polysaccharides, binding to elements of the cell wall,
or by export through the different transporters. The
capability of micro-organisms to neutralize the toxic effect
of heavy metals or to detect heavy metals could be effectively
used in environmental biotechnology. Traditional physicochemical methods applied to the remediation of heavy metal
contaminated soil often damage its structure and quality, as
well as existing ecosystems. Additionally, these techniques
often require soil excavation and transport to a reclamation
site, and this in turn generates high costs and the need for
specialized equipment (Park et al., 2010). Moreover, other
techniques such as phytoremediation often depend on the
climate, water and soil conditions.
Thus, microbial methods in this context may offer an
attractive alternative. So far, lead-resistant micro-organisms have been successfully used for the immobilization
and biosorption of lead (Park et al., 2010; Wu et al., 2006;
Wilson et al., 2006), the enhancement of phytoremediation
as described in the case of Solanum nigrum L. (Gao et al.,
2012), the construction of lead detecting biosensors
(Chakraborty et al., 2008; Chen et al., 2005; Chiu &
Yang, 2012; Corbisier et al., 1999), bioaugmentation of lead
contaminated mine tailings followed by lead biomineralization (Govarthanan et al., 2013) and bioremediation of
lead contaminated soil (Chatterjee et al., 2012; Chen et al.,
2011; Guo et al., 2010). It seems that the combination of
the traditional and microbial approaches in environmental
biotechnology may yield promising results.
Aickin, R. M. & Dean, A. C. R. (1977). Lead accumulation by micro-
organisms. Microbios Lett 5, 129–133.
Aickin, R. M. & Dean, A. C. R. (1979). Lead accumulation by
Pseudomonas fluorescens and by Citrobacter sp. Microbios Lett 9, 55–
66.
Aickin, R. M., Dean, A. C. R., Cheethama, K. & Skarnulias, A. J.
(1979). Electron microscope studies on the uptake of lead by a
Citrobacter species. Microbios Letters 9, 7–15.
Aiking, H., Govers, H. & van’t Riet, J. (1985). Detoxification of
mercury, cadmium, and lead in Klebsiella aerogenes NCTC 418
growing in continuous culture. Appl Environ Microbiol 50, 1262–1267.
Akmal, M. & Jianming, X. (2009). Microbial biomass and bacterial
community changes by Pb contamination in acidic soil. J Agric Biol
Sci 1, 30–37.
al-Aoukaty, A., Appanna, V. D. & Huang, J. (1991). Exocellular and
intracellular accumulation of lead in Pseudomonas fluorescens ATCC
13525 is mediated by the phosphate content of the growth medium.
FEMS Microbiol Lett 67, 283–290.
Amoozegar, M. A., Ghazanfari, N. & Didari, M. (2012). Lead and
cadmium bioremoval by Halomonas sp., an exopolysaccharideproducing halophilic bacterium. Prog Biol Sci 2, 1–11.
Apell, H. J. (2004). How do P-type ATPases transport ions? Bioelectrochemistry 63, 149–156.
Argüello, J. M. (2003). Identification of ion-selectivity determinants in
heavy-metal transport P1B-type ATPases. J Membr Biol 195, 93–108.
Arunakumara, K. K. I. U. & Xuecheng, Z. (2009). Effects of heavy
metals (Pb2+ and Cd2+) on the ultrastructure, growth and pigment
contents of the unicellular cyanobacterium Synechocystis sp. PCC
6803. Chin J Oceanology Limnol 27, 383–388.
Babich, H. & Stotzky, G. (1979). Abiotic factors affecting the toxicity
of lead to fungi. Appl Environ Microbiol 38, 506–513.
Banci, L., Bertini, I., Cantini, F., Ciofi-Baffoni, S., Cavet, J. S.,
Dennison, C., Graham, A. I., Harvie, D. R. & Robinson, N. J. (2007).
NMR structural analysis of cadmium sensing by winged helix
repressor CmtR. J Biol Chem 282, 30181–30188.
Beveridge, T. J. & Fyfe, W. S. (1985). Metal fixation by bacterial cell
walls. Can J Earth Sci 22, 1892–1898.
Binet, M. R. & Poole, R. K. (2000). Cd(II), Pb(II) and Zn(II) ions
regulate expression of the metal-transporting P-type ATPase ZntA in
Escherichia coli. FEBS Lett 473, 67–70.
Blindauer, C. A. (2011). Bacterial metallothioneins: past, present, and
questions for the future. J Biol Inorg Chem 16, 1011–1024.
Borremans, B., Hobman, J. L., Provoost, A., Brown, N. L. & van Der
Lelie, D. (2001). Cloning and functional analysis of the pbr lead
resistance determinant of Ralstonia metallidurans CH34. J Bacteriol
183, 5651–5658.
Acknowledgements
Braud, A., Geoffroy, V., Hoegy, F., Mislin, G. L. A. & Schalk, I. J.
(2010). Presence of the siderophores pyoverdine and pyochelin in
The review was prepared during research on the project financed by
an individual grant from Ministry of Polish Science and Higher
Education N N305 185937. A. J. is a PhD student supported by a PhD
scholarship from the UPGOW project task 55 at the University of
Silesia within the European Social Fund.
the extracellular medium reduces toxic metal accumulation in
Pseudomonas aeruginosa and increases bacterial metal tolerance.
Environ Microbiol Rep 2, 419–425.
References
MerR-like transcriptional regulator of zntA in Escherichia coli. Mol
Microbiol 31, 893–902.
Aguilera, M., Quesada, M. T., Del Aguila, V. G., Morillo, J. A.,
Rivadeneyra, M. A., Ramos-Cormenzana, A. & Monteoliva-Sánchez,
M. (2008). Characterisation of Paenibacillus jamilae strains that
produce exopolysaccharide during growth on and detoxification of
olive mill wastewaters. Bioresour Technol 99, 5640–5644.
http://mic.sgmjournals.org
Brocklehurst, K. R., Hobman, J. L., Lawley, B., Blank, L., Marshall,
S. J., Brown, N. L. & Morby, A. P. (1999). ZntR is a Zn(II)-responsive
Brown, N. L., Stoyanov, J. V., Kidd, S. P. & Hobman, J. L. (2003). The
MerR family of transcriptional regulators. FEMS Microbiol Rev 27,
145–163.
Bruins, M. R., Kapil, S. & Oehme, F. W. (2000). Microbial resistance to
metals in the environment. Ecotoxicol Environ Saf 45, 198–207.
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
21
A. Jarosławiecka and Z. Piotrowska-Seget
Busenlehner, L. S. & Giedroc, D. P. (2006). Kinetics of metal binding
Debut, A. J., Dumay, Q. C., Barabote, R. D. & Saier, M. H., Jr (2006).
by the toxic metal-sensing transcriptional repressor Staphylococcus
aureus pI258 CadC. J Inorg Biochem 100, 1024–1034.
The iron/lead transporter superfamily of Fe/Pb2+ uptake systems.
J Mol Microbiol Biotechnol 11, 1–9.
Busenlehner, L. S., Pennella, M. A. & Giedroc, D. P. (2003). The
Donmez, G. & Aksu, Z. (1999). The effect of copper (II) ions on the
SmtB/ArsR family of metalloregulatory transcriptional repressors:
structural insights into prokaryotic metal resistance. FEMS Microbiol
Rev 27, 131–143.
growth and bioaccumulation properties of some yeasts. Process
Biochem 35, 135–142.
Bussche, J. V. & Soares, E. V. (2011). Lead induces oxidative stress
and phenotypic markers of apoptosis in Saccharomyces cerevisiae. Appl
Microbiol Biotechnol 90, 679–687.
Çabuk, A., Akar, T., Tunali, S. & Tabak, O. (2006). Biosorption
characteristics of Bacillus sp. ATS-2 immobilized in silica gel for
removal of Pb(II). J Hazard Mater 136, 317–323.
Çabuk, A., Akar, T., Tunali, S. & Gedikli, S. (2007). Biosorption of
Pb(II) by industrial strain of Saccharomyces cerevisiae immobilized on
the biomatrix of cone biomass of Pinus nigra: equilibrium and
mechanism analysis. Chem Eng J 131, 293–300.
Cavet, J. S., Graham, A. I., Meng, W. & Robinson, N. J. (2003). A
cadmium-lead-sensing ArsR-SmtB repressor with novel sensory sites.
Complementary metal discrimination by NmtR and CmtR in a
common cytosol. J Biol Chem 278, 44560–44566.
Chakraborty, T., Babu, P. G., Alam, A. & Chaudhari, A. (2008). GFP
expressing bacterial biosensor to measure lead contamination in
aquatic environment. Curr Sci 94, 800–805.
Chatterjee, S., Anindita, M., Agniswar, S. & Pranab, R. (2012).
Bioremediation of lead by lead-resistant microorganisms, isolated
from industrial sample. Adv Biosci Biotechnol 3, 290–295.
Chen, P. R. & He, C. (2008). Selective recognition of metal ions by
metalloregulatory proteins. Curr Opin Chem Biol 12, 214–221.
Chen, C. & Wang, J. (2007). Response of Saccharomyces cerevisiae to
Dopson, M., Baker-Austin, C., Koppineedi, P. R. & Bond, P. L. (2003).
Growth in sulfidic mineral environments: metal resistance mechanisms in acidophilic micro-organisms. Microbiology 149, 1959–1970.
Dutta, S. J., Liu, J., Hou, Z. & Mitra, B. (2006). Conserved aspartic acid
714 in transmembrane segment 8 of the ZntA subgroup of P1B-type
ATPases is a metal-binding residue. Biochemistry 45, 5923–5931.
Dutta, S. J., Liu, J., Stemmler, A. J. & Mitra, B. (2007). Conservative and
nonconservative mutations of the transmembrane CPC motif in ZntA:
effect on metal selectivity and activity. Biochemistry 46, 3692–3703.
Freire-Nordi, C. S., Vieira, A. A. H., Nakaie, C. R. & Nascimento, O. R.
(2005). Effect of polysaccharide capsule of the microalgae Staurastrum
iversenii var. americanum on diffusion of charged and uncharged
molecules, using EPR technique. Braz J Phys 36, 75–82.
Gabr, R. M., Hassan, S. H. A. & Shoreit, A. A. M. (2008). Biosorption
of lead and nickel by living and non-living cells of Pseudomonas
aeruginosa ASU 6a. Int Biodeterior Biodegradation 62, 195–203.
Gadd, G. M. (2010). Metals, minerals and microbes: geomicrobiology
and bioremediation. Microbiology 156, 609–643.
Gao, Y., Miao, C., Wang, Y., Xia, J. & Zhou, P. (2012). Metal-resistant
microorganisms and metal chelators synergistically enhance the
phytoremediation efficiency of Solanum nigrum L. in Cd- and Pbcontaminated soil. Environ Technol 33, 1383–1389.
Gerba, C. P. (1996). Principles of toxicology. In Environmental and
lead ion stress. Appl Microbiol Biotechnol 74, 683–687.
Pollution Science. pp. 322–44. Edited by I. L. Pepper & other authors.
San Diego: Academic Press.
Chen, Y. T., Chang, H. Y., Lai, Y. C., Pan, C. C., Tsai, S. F. & Peng, H. L.
(2004). Sequencing and analysis of the large virulence plasmid pLVPK
Gharieb, M. M. & Gadd, G. M. (2004). Role of glutathione in
of Klebsiella pneumoniae CG43. Gene 337, 189–198.
Chen, P., Greenberg, B., Taghavi, S., Romano, C., van der Lelie, D. &
He, C. (2005). An exceptionally selective lead(II)-regulatory protein
from Ralstonia metallidurans: development of a fluorescent lead(II)
probe. Angew Chem Int Ed Engl 44, 2715–2719.
Chen, P. R., Wasinger, E. C., Zhao, J., van der Lelie, D., Chen, L. X. &
He, C. (2007). Spectroscopic insights into lead(II) coordination by the
selective lead(II)-binding protein PbrR691. J Am Chem Soc 129,
12350–12351.
detoxification of metal(loid)s by Saccharomyces cerevisiae. Biometals
17, 183–188.
Govarthanan, M., Lee, K. J., Cho, M., Kim, J. S., Kamala-Kannan, S. &
Oh, B. T. (2013). Significance of autochthonous Bacillus sp. KK1 on
biomineralization of lead in mine tailings. Chemosphere 90, 2267–2272.
Grass, G., Wong, M. D., Rosen, B. P., Smith, R. L. & Rensing, C.
(2002). ZupT is a Zn(II) uptake system in Escherichia coli. J Bacteriol
184, 864–866.
Grosse, C., Friedrich, S. & Nies, D. H. (2007). Contribution of
J Environ Protect 2, 130–141.
extracytoplasmic function sigma factors to transition metal homeostasis in Cupriavidus metallidurans strain CH34. J Mol Microbiol
Biotechnol 12, 227–240.
Chiu, T. Y. & Yang, D. M. (2012). Intracellular Pb2+ content
Guibaud, G., Tixier, N., Bouju, A. & Baudu, M. (2003). Relation
Chen, B., Jia-nan, L., Zheng, W., Lei, D., Jing-hua, F. & Juan-juan,
Q. (2011). Remediation of Pb-resistant bacteria to Pb polluted soil.
monitoring using a protein-based Pb2+ indicator. Toxicol Sci 126,
436–445.
between extracellular polymers’ composition and its ability to
complex Cd, Cu and Pb. Chemosphere 52, 1701–1710.
Clipson, N. & Gleeson, D. B. (2012). Fungal biogeochemistry: a
Guo, H., Luo, S., Chen, L., Xiao, X., Xi, Q., Wei, W., Zeng, G., Liu, C.,
Wan, Y. & other authors (2010). Bioremediation of heavy metals by
central role in the environmental fate of lead. Curr Biol 22, R82–R84.
Coblenz, A. & Wolf, K. (1994). The role of glutathione biosynthesis in
heavy metal resistance in the fission yeast Schizosaccharomyces pombe.
FEMS Microbiol Rev 14, 303–308.
Coombs, J. M. & Barkay, T. (2005). New findings on evolution of
metal homeostasis genes: evidence from comparative genome analysis
of bacteria and archaea. Appl Environ Microbiol 71, 7083–7091.
Corbisier, P., van der Lelie, D., Borremans, B., Provoost, A., de
Lorenzo, V., Brown, N. L., Lloyd, J. R., Hobman, J. L., Csoregi, E. &
other authors (1999). Whole cell- and protein-based biosensors for
the detection of bioavailable heavy metals in environmental samples.
Anal Chim Acta 387, 235–244.
22
growing hyperaccumulator endophytic bacterium Bacillus sp. L14.
Bioresour Technol 101, 8599–8605.
Haritha, A., Sagar, K. P., Tiwari, A., Kiranmayi, P., Rodrigue, A.,
Mohan, P. M. & Singh, S. S. (2009). MrdH, a novel metal resistance
determinant of Pseudomonas putida KT2440, is flanked by metalinducible mobile genetic elements. J Bacteriol 191, 5976–5987.
Hasnain, S., Yasmin, S. & Yasmin, A. (1993). The effects of lead-
resistant pseudomonads on the growth of Triticum aestivum seedlings
under lead stress. Environ Pollut 81, 179–184.
Hobman, J. L., Julian, D. J. & Brown, N. L. (2012). Cysteine
coordination of Pb(II) is involved in the PbrR-dependent activation
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
of the lead-resistance promoter, PpbrA, from Cupriavidus metallidurans CH34. BMC Microbiol 12, 109.
is regulated by AztR, an alpha3N ArsR/SmtB metalloregulator.
Biochemistry 44, 8673–8683.
Hou, Z. J., Narindrasorasak, S., Bhushan, B., Sarkar, B. & Mitra, B.
(2001). Functional analysis of chimeric proteins of the Wilson Cu(I)-
Liu, J., Dutta, S. J., Stemmler, A. J. & Mitra, B. (2006). Metal-binding
ATPase (ATP7B) and ZntA, a Pb(II)/Zn(II)/Cd(II)-ATPase from
Escherichia coli. J Biol Chem 276, 40858–40863.
Huckle, J. W., Morby, A. P., Turner, J. S. & Robinson, N. J. (1993).
Isolation of a prokaryotic metallothionein locus and analysis of
transcriptional control by trace metal ions. Mol Microbiol 7, 177–187.
Hughes, M. N. & Poole, R. K. (1989). Metals and Micro-organisms, p.
277. London, UK: Chapman & Hall.
Hynninen, A. (2010). Zinc, cadmium and lead resistance mechanisms
affinity of the transmembrane site in ZntA: implications for metal
selectivity. Biochemistry 45, 763–772.
Loaëc, M., Olier, R. & Guezennec, J. (1997). Uptake of lead, cadmium
and zinc by a novel bacterial exopolysaccharide. Water Res 31, 1171–
1179.
Makui, H., Roig, E., Cole, S. T., Helmann, J. D., Gros, P. & Cellier, M. F.
(2000). Identification of the Escherichia coli K-12 Nramp orthologue
(MntH) as a selective divalent metal ion transporter. Mol Microbiol
35, 1065–1078.
in bacteria and their contribution to biosensing. Academic Dissertation in Microbiology, Department of Food and Environmental Sciences, Faculty of Agriculture and Forestry, University of
Helsinki.
Mayer, R. A. & Godwin, H. A. (2006). Preparation of media and buffers
with soluble lead. Anal Biochem 356, 142–144.
Hynninen, A., Touzé, T., Pitkänen, L., Mengin-Lecreulx, D. & Virta, M.
(2009). An efflux transporter PbrA and a phosphatase PbrB cooperate
quorum-sensing interactions. Appl Environ Microbiol 70, 855–864.
in a lead-resistance mechanism in bacteria. Mol Microbiol 74, 384–394.
Janssen, P. J., Van Houdt, R., Moors, H., Monsieurs, P., Morin, N.,
Michaux, A., Benotmane, M. A., Leys, N., Vallaeys, T. & other authors
(2010). The complete genome sequence of Cupriavidus metallidurans
strain CH34, a master survivalist in harsh and anthropogenic
environments. PLoS ONE 5, e10433.
Mire, C. E., Tourjee, J. A., O’Brien, W. F., Ramanujachary, K. V. & Hecht,
G. B. (2004). Lead precipitation by Vibrio harveyi: evidence for novel
Mitra, B. & Sharma, R. (2001). The cysteine-rich amino-terminal
domain of ZntA, a Pb(II)/Zn(II)/Cd(II)-translocating ATPase from
Escherichia coli, is not essential for its function. Biochemistry 40, 7694–
7699.
Monchy, S., Benotmane, M. A., Janssen, P., Vallaeys, T., Taghavi, S.,
van der Lelie, D. & Mergeay, M. (2007). Plasmids pMOL28 and
Karnachuk, O. V., Kurochkina, S. Y. & Tuovinen, O. H. (2002).
pMOL30 of Cupriavidus metallidurans are specialized in the maximal
viable response to heavy metals. J Bacteriol 189, 7417–7425.
Growth of sulfate-reducing bacteria with solid-phase electron
acceptors. Appl Microbiol Biotechnol 58, 482–486.
Monsieurs, P., Moors, H., Van Houdt, R., Janssen, P. J., Janssen, A.,
Coninx, I., Mergeay, M. & Leys, N. (2011). Heavy metal resistance
Kotuby-Amacher, J., Gambrell, R. P. & Amacher, M. C. (1992). The
in Cupriavidus metallidurans CH34 is governed by an intricate
transcriptional network. Biometals 24, 1133–1151.
distribution and environmental chemistry of lead in soil at an
abandoned battery reclamation site. In Engineering Aspects of Metal–
Waste Management, pp. 1–24. Edited by I. K. Iskander & H. M. Selim.
Boca Raton, FL: Lewis Publishers.
Kumar, M. & Upreti, R. K. (2000). Impact of lead stress and adaptation
in Escherichia coli. Ecotoxicol Environ Saf 47, 246–252.
Laddaga, R. A. & Silver, S. (1985). Cadmium uptake in Escherichia
coli K-12. J Bacteriol 162, 1100–1105.
Laddaga, R. A., Bessen, R. & Silver, S. (1985). Cadmium resistant
mutant of Bacillus subtilis 168 with reduced cadmium transport.
J Bacteriol 262, 1106–1110.
Lee, L. J., Barrett, J. A. & Poole, R. K. (2005). Genome-wide
Morillo, J. A., Aguilera, M., Ramos-Cormenzana, A. & MonteolivaSánchez, M. (2006). Production of a metal-binding exopolysacchar-
ide by Paenibacillus jamilae using two-phase olive-mill waste as
fermentation substrate. Curr Microbiol 53, 189–193.
Murthy, S., Geetha, B. & Sarangi, S. K. (2011). Effect of lead on
metallothionein concentration in lead-resistant bacteria Bacillus cereus
isolated from industrial effluent. Afr J Biotechnol 10, 15966–15972.
Naik, M. M. & Dubey, S. K. (2011). Lead-enhanced siderophore
production and alteration in cell morphology in a Pb-resistant
Pseudomonas aeruginosa strain 4EA. Curr Microbiol 62, 409–414.
Naik, M. M., Pandey, A. & Dubey, S. K. (2012). Pseudomonas
transcriptional response of chemostat-cultured Escherichia coli to zinc.
J Bacteriol 187, 1124–1134.
aeruginosa strain WI-1 from Mandovi estuary possesses metallothionein to alleviate lead toxicity and promotes plant growth. Ecotoxicol
Environ Saf 79, 129–133.
Leedjärv, A., Ivask, A. & Virta, M. (2008). Interplay of different
Nies, D. H. (1992). Resistance to cadmium, cobalt, zinc, and nickel in
transporters in the mediation of divalent heavy metal resistance in
Pseudomonas putida KT2440. J Bacteriol 190, 2680–2689.
Leonhartsberger, S., Huber, A., Lottspeich, F. & Böck, A. (2001). The
hydH/G genes from Escherichia coli code for a zinc and lead
responsive two-component regulatory system. J Mol Biol 307, 93–105.
Levinson, H. S. & Mahler, I. (1998). Phosphatase activity and lead
resistance in Citrobacter freundii and Staphylococcus aureus. FEMS
Microbiol Lett 161, 135–138.
Levinson, H. S., Mahler, I., Blackwelder, P. & Hood, T. (1996). Lead
resistance and sensitivity in Staphylococcus aureus. FEMS Microbiol
Lett 145, 421–425.
Li, Z. S., Lu, Y. P., Zhen, R. G., Szczypka, M., Thiele, D. J. & Rea, P. A.
(1997). A new pathway for vacuolar cadmium sequestration in
Saccharomyces cerevisiae: YCF1-catalyzed transport of bis(glutathionato)cadmium. Proc Natl Acad Sci U S A 94, 42–47.
Liu, T., Golden, J. W. & Giedroc, D. P. (2005). A zinc(II)/lead(II)/
cadmium(II)-inducible operon from the Cyanobacterium anabaena
http://mic.sgmjournals.org
microbes. Plasmid 27, 17–28.
Nies, D. H. (2003). Efflux-mediated heavy metal resistance in
prokaryotes. FEMS Microbiol Rev 27, 313–339.
Noll, M., Petrukhin, K. & Lutsenko, S. (1998). Identification of a novel
transcription regulator from Proteus mirabilis, PMTR, revealed a
possible role of YJAI protein in balancing zinc in Escherichia coli.
J Biol Chem 273, 21393–21401.
Pal, A. & Paul, A. K. (2008). Microbial extracellular polymeric
substances: central elements in heavy metal bioremediation. Indian J
Microbiol 48, 49–64.
Paperi, R., Micheletti, E. & De Philippis, R. (2006). Optimization of
copper sorbing-desorbing cycles with confined cultures of the
exopolysaccharide-producing cyanobacterium Cyanospira capsulata.
J Appl Microbiol 101, 1351–1356.
Park, J., Bolan, N., Megharaj, M. & Naidu, R. (2010). Isolation of
phosphate-solubilizing bacteria and characterization of their effects
on lead immobilization. Pedologist 53, 67–75.
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
23
A. Jarosławiecka and Z. Piotrowska-Seget
Park, J. H., Bolan, N., Meghraj, M., Naidu, R. & Chung, J. W. (2011).
Salehizadeh, H. & Shojaosadati, S. A. (2003). Removal of metal ions
Bacterial-assisted immobilization of lead in soils: implications for
remediation. Pedologist 54, 162–174.
from aqueous solution by polysaccharide produced from Bacillus
firmus. Water Res 37, 4231–4235.
Pennella, M. A. & Giedroc, D. P. (2005). Structural determinants of
Sani, R. K., Geesey, G. & Peyton, B. (2001). Assessment of lead
toxicity to Desulfovibrio desulfuricans G20: influence of components
of Lactate C medium. Adv Environ Res 5, 269–276.
metal selectivity in prokaryotic metal-responsive transcriptional
regulators. Biometals 8, 413–428.
Perdrial, N., Liewig, N., Delphin, J. E. & Elsass, F. (2008). TEM
evidence for intracellular accumulation of lead by bacteria in
subsurface environments. Chem Geol 253, 196–204.
Pereira, S., Micheletti, E., Zille, A., Santos, A., Moradas-Ferreira, P.,
Tamagnini, P. & De Philippis, R. (2011). Using extracellular polymeric
substances (EPS)-producing cyanobacteria for the bioremediation of
heavy metals: do cations compete for the EPS functional groups and
also accumulate inside the cell? Microbiology 157, 451–458.
Perez, M. P. J. A., Garcı́a-Ribera, R., Quesada, T., Aguilera,
M., Ramos-Cormenzana, A. & Monteoliva-Sánchez, M. (2008).
Biosorption of heavy metals by the exopolysaccharide produced by
Paenibacillus jamilae. World J Microbiol Biotechnol 24, 2699–2704.
Permina, E. A., Kazakov, A. E., Kalinina, O. V. & Gelfand, M. S. (2006).
Comparative genomics of regulation of heavy metal resistance in
Eubacteria. BMC Microbiol 6, 49–60.
Pongratz, R. & Heumann, K. G. (1999). Production of methylated
mercury, lead and cadmium by marine bacteria as a significant
natural source for atmospheric heavy metals in polar regions.
Chemosphere 39, 89–102.
Prévéral, S., Ansoborlo, E., Mari, S., Vavasseur, A. & Forestier, C.
(2006). Metal(loid)s and radionuclides cytotoxicity in Saccharomyces
cerevisiae. Role of YCF1, glutathione and effect of buthionine
sulfoximine. Biochimie 88, 1651–1663.
Raimunda, D., Subramanian, P., Stemmler, T. & Argüello, J. M.
(2012). A tetrahedral coordination of zinc during transmembrane
transport by P-type Zn2+-ATPases. Biochim Biophys Acta 1818, 1374–
1377.
Reisinger, K., Stoeppler, M. & Nürnberg, H. W. (1981). Evidence for
the absence of biological methylation of lead in the environment.
Nature 291, 228–230.
Rensing, C., Mitra, B. & Rosen, B. P. (1997). The zntA gene of
Escherichia coli encodes a Zn(II)-translocating P-type ATPase. Proc
Natl Acad Sci U S A 94, 14326–14331.
Rensing, C., Sun, Y., Mitra, B. & Rosen, B. P. (1998). Pb(II)-
translocating P-type ATPases. J Biol Chem 273, 32614–32617.
Rensing, C., Ghosh, M. & Rosen, B. P. (1999). Families of soft-metal-
ion-transporting ATPases. J Bacteriol 181, 5891–5897.
Rhee, Y. J., Hillier, S. & Gadd, G. M. (2012). Lead transformation to
pyromorphite by fungi. Curr Biol 22, 237–241.
Sayer, J. A., Cotter-Howells, J. D., Watson, C., Hillier, S. & Gadd,
G. M. (1999). Lead mineral transformation by fungi. Curr Biol 9, 691–
694.
Schalk, I. J., Hannauer, M. & Braud, A. (2011). New roles for bacterial
siderophores in metal transport and tolerance. Environ Microbiol 13,
2844–2854.
Scherer, J. & Nies, D. H. (2009). CzcP is a novel efflux system
contributing to transition metal resistance in Cupriavidus metallidurans CH34. Mol Microbiol 73, 601–621.
Schwab, A. P., He, Y. H. & Banks, M. K. (2005). The influence of
organic ligands on the retention of lead in soil. Chemosphere 61, 856–
866.
Shen, L., Xia, J. L., He, H. & Nie, Z. Y. (2008). Comparative study on
biosorption of Pb(II) and Cr(VI) by Synechococcus sp. Trans Nonferr
Metal Soc 18, 1336–1342.
Sheng, X.-F., Xia, J.-J., Jiang, C. Y., He, L. Y. & Qian, M. (2008).
Characterization of heavy metal-resistant endophytic bacteria from
rape (Brassica napus) roots and their potential in promoting the
growth and lead accumulation of rape. Environ Pollut 156, 1164–
1170.
Shin, M. N., Shim, J., You, Y., Myung, H., Bang, K. S., Cho, M.,
Kamala-Kannan, S. & Oh, B. T. (2012). Characterization of lead
resistant endophytic Bacillus sp. MN3-4 and its potential for
promoting lead accumulation in metal hyperaccumulator Alnus
firma. J Hazard Mater 199–200, 314–320.
Shiraishi, E., Inouhe, M., Joho, M. & Tohoyama, H. (2000). The
cadmium-resistant gene, CAD2, which is a mutated putative coppertransporter gene (PCA1), controls the intracellular cadmium-level in
the yeast S. cerevisiae. Curr Genet 37, 79–86.
Siegel, F. R. (2002). Environmental Geochemistry of Potentially Toxic
Metals, p. 218. Berlin: Springer-Verlag.
Silver, S. & Phung, T. (2005). A bacterial view of the periodic table:
genes and proteins for toxic inorganic ions. J Ind Microbiol Biotechnol
32, 587–605.
Silverberg, B. A., Wong, P. T. & Chau, Y. K. (1976). Ultrastructural
examination of Aeromonas cultured in the presence of organic lead.
Appl Environ Microbiol 32, 723–725.
Silverberg, B. A., Wong, P. T. S. & Chau, Y. K. (1977). Effect of
Rifaat, H. M., Mahrous, K. F. & Khalil, W. K. B. (2009). Effect of heavy
tetramethyl lead on freshwater green algae. Arch Environ Contam
Toxicol 5, 305–313.
metals upon metallothioneins in some Streptomyces species isolated
from Egyptian soil. J Appl Sci Environ Sanit 4, 197–206.
Singh, S. P., Tack, F. M. G. & Verioo, M. G. (1996). Extractability and
Roane, T. M. (1999). Lead resistance in two bacterial isolates from
heavy metal contaminated soils. Microb Ecol 37, 218–224.
Roane, T. M. & Kellogg, S. T. (1996). Characterization of bacterial
communities in heavy metal contaminated soils. Can J Microbiol 42,
593–603.
Sag, Y., Özer, D. & Kutsal, T. (1995). A comparative study of the
biosorption of lead(II) ions to Z. ramigera and R. arrhizus. Process
Biochem 30, 169–174.
bioavailability of heavy metals in surface soils derived from dredged
sediments. Chem Spec Bioavail 8, 105–110.
Smeaton, C. M., Fryer, B. J. & Weisener, C. G. (2009). Intracellular
precipitation of Pb by Shewanella putrefaciens CN32 during the
reductive dissolution of Pb-jarosite. Environ Sci Technol 43, 8086–
8091.
So, N. W., Rho, J. Y., Lee, S. Y., Hancock, I. C. & Kim, J. H.
(2001). A lead-absorbing protein with superoxide dismutase
Saha, R., Saha, N., Donofrio, R. S. & Bestervelt, L. L. (2012).
activity from Streptomyces subrutilus. FEMS Microbiol Lett 194,
93–98.
Microbial siderophores: a mini review. J Basic Microbiol 52, 1–15.
Soares, E. V., Duarte, A. P., Boaventura, R. A. & Soares, H. M. (2002).
Sakamoto, F., Ohnuki, T., Fujii, T. & Iefuji, H. (2010). Response of
Viability and release of complexing compounds during accumulation
of heavy metals by a brewer’s yeast. Appl Microbiol Biotechnol 58, 836–
841.
Saccharomyces cerevisiae to heavy element stress: lead vs. uranium.
Geomicrobiol J 27, 240–244.
24
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
Microbiology 160
Lead resistance in micro-organisms
Soares, E. V., Hebbelinck, K. & Soares, H. M. (2003). Toxic effects
Trajanovska, S., Britz, M. L. & Bhave, M. (1997). Detection of heavy
caused by heavy metals in the yeast Saccharomyces cerevisiae: a
comparative study. Can J Microbiol 49, 336–343.
metal ion resistance genes in gram-positive and gram-negative
bacteria isolated from a lead-contaminated site. Biodegradation 8,
113–124.
Song, W. Y., Sohn, E. J., Martinoia, E., Lee, Y. J., Yang, Y. Y., Jasinski,
M., Forestier, C., Hwang, I. & Lee, Y. (2003). Engineering tolerance
and accumulation of lead and cadmium in transgenic plants. Nat
Biotechnol 21, 914–919.
Van der Heggen, M., Martins, S., Flores, G. & Soares, E. V. (2010).
Lead toxicity in Saccharomyces cerevisiae. Appl Microbiol Biotechnol
88, 1355–1361.
Tripathi, M., Munot, H. P., Shouche, Y., Meyer, J. M. & Goel, R. (2005).
Isolation and functional characterization of siderophore-producing
lead- and cadmium-resistant Pseudomonas putida KNP9. Curr
Microbiol 50, 233–237.
Turner, J. S., Glands, P. D., Samson, A. C. R. & Robinson, N. J. (1996).
surfaces. Elements 1, 193–197.
Zn2+-sensing by the cyanobacterial metallothionein repressor SmtB:
different motifs mediate metal-induced protein-DNA dissociation.
Nucleic Acids Res 24, 3714–3721.
Spiro, T. G. & Stigliani, W. M. (2002). Chemistry of the Environment, p.
Tynecka, Z., Gos, Z. & Zajac, J. (1981). Reduced cadmium transport
356. Upper Saddle River, NJ: Prentice-Hall.
2+
Suh, J. H., Kim, D. S., Yun, J. W. & Song, S. K. (1998). Process of Pb
determined by a resistance plasmid in Staphylococcus aureus.
J Bacteriol 147, 305–312.
accumulation in Saccharomyces cerevisiae. Biotechnol Lett 20, 153–156.
Vallee, B. L. & Ulmer, D. D. (1972). Biochemical effects of mercury,
Sparks, D. L. (2005). Toxic metals in the environment: the role of
Suh, J. H., Yun, J. W. & Kim, D. S. (1999a). Cation (K+, Mg2+, Ca2+)
cadmium, and lead. Annu Rev Biochem 41, 91–128.
exchange in Pb2+ accumulation by Saccharomyces cerevisiae.
Bioprocess Eng 21, 383–387.
VanZile, M. L., Chen, X. & Giedroc, D. P. (2002). Allosteric negative
Suh, J. H., Yun, J. W. & Kim, D. S. (1999b). Effect of extracellular
polymeric substances (EPS) on Pb2+ accumulation by Aureobasidium
pullulans. Bioprocess Biosyst Eng 21, 1–4.
Suh, J. H., Kim, D. S. & Song, S. K. (2001). Inhibition effect of initial
Pb2+ concentration on Pb2+ accumulation by Saccharomyces
cerevisiae and Aureobasidium pullulans. Bioresour Technol 79, 99–102.
Sun, F. & Shao, Z. (2007). Biosorption and bioaccumulation of lead
by Penicillium sp. Psf-2 isolated from the deep sea sediment of the
Pacific Ocean. Extremophiles 11, 853–858.
Sun, Y., Wong, M. D. & Rosen, B. P. (2001). Role of cysteinyl residues
in sensing Pb(II), Cd(II), and Zn(II) by the plasmid pI258 CadC
repressor. J Biol Chem 276, 14955–14960.
regulation of smt O/P binding of the zinc sensor, SmtB, by metal ions:
a coupled equilibrium analysis. Biochemistry 41, 9776–9786.
Walton, A., Ebdon, L. & Millward, G. (1988). Methylation of inorganic
lead by Tamyar Estuary (UK) sediments. Appl Organomet Chem 2,
87–93.
Wang, Y., Kendall, J., Cavet, J. S. & Giedroc, D. P. (2010).
Elucidation of the functional metal binding profile of a Cd/Pb
sensor CmtR from Streptomyces coelicolor. Biochemistry 49, 6617–
6626.
Wilson, C., Brigmon, R. L., Knox, A., Seaman, J. & Smith, G. (2006).
Effects of microbial and phosphate amendments on the bioavailability
of lead (Pb) in shooting range soil. Bull Environ Contam Toxicol 76,
392–399.
Szczypka, M. S., Wemmie, J. A., Moye-Rowley, W. S. & Thiele, D. J.
(1994). A yeast metal resistance protein similar to human cystic
Wong, P. T. S., Chau, Y. K. & Luxon, P. L. (1975). Methylation of lead
fibrosis transmembrane conductance regulator (CFTR) and multidrug resistance-associated protein. J Biol Chem 269, 22853–22857.
Wu, S. C., Luo, Y. M., Cheung, K. C. & Wong, M. H. (2006). Influence
Taboski, M. A. S., Rand, T. G. & Piórko, A. (2005). Lead and cadmium
uptake in the marine fungi Corollospora lacera and Monodictys
pelagica. FEMS Microbiol Ecol 53, 445–453.
Taghavi, S., Lesaulnier, C., Monchy, S., Wattiez, R., Mergeay, M. &
van der Lelie, D. (2009). Lead(II) resistance in Cupriavidus
metallidurans CH34: interplay between plasmid and chromosomally-located functions. Antonie van Leeuwenhoek 96, 171–182.
Templeton, A. S., Trainor, T. P., Spormann, A. M., Newville, M.,
Sutton, S. R., Dohnalkova, A., Gorby, Y. & Brown, G. E., Jr (2003).
Sorption versus biomineralization of Pb(II) within Burkholderia
cepacia biofilms. Environ Sci Technol 37, 300–307.
Thayer, J. S. (2002). Biological methylation of less-studied elements.
Appl Organomet Chem 16, 677–691.
Thayer, J. S. & Brinckman, F. (1982). The biological methylation of
metals and metalloids. Adv Organomet Chem 20, 313–356.
Thelwell, C., Robinson, N. J. & Turner-Cavet, J. S. (1998). An SmtB-
like repressor from Synechocystis PCC 6803 regulates a zinc exporter.
Proc Natl Acad Sci U S A 95, 10728–10733.
Tornabene, T. G. & Edwards, H. W. (1972). Microbial uptake of lead.
Science 176, 1334–1335.
in the environment. Nature 253, 263–264.
of bacteria on Pb and Zn speciation, mobility and bioavailability in
soil: a laboratory study. Environ Pollut 144, 765–773.
Yokel, J. & Delistraty, D. A. (2003). Arsenic, lead, and other trace
elements in soils contaminated with pesticide residues at the Hanford
site (USA). Environ Toxicol 18, 104–114.
Yoon, K. P., Misra, T. K. & Silver, S. (1991). Regulation of the cadA
cadmium resistance determinant of Staphylococcus aureus plasmid
pI258. J Bacteriol 173, 7643–7649.
Yuan, X. F. & Tang, C. C. (1999). DNA damage and repair in yeast
(Saccharomyces cerevisiae) cells exposed to lead. J Environ Sci Health A
34, 1117–1128.
Zanardini, E., Andreoni, V., Borina, S., Cappitellia, F., Daffonchio,
D., Talottaa, P., Sorlinia, C., Ranallib, G., Brunic, S. & Cariatic,
F. (1997). Lead-resistant microorganisms from red stains of
marble of the Certosa of Pavia, Italy and use of nucleic acidbased techniques for their detection. Int Biodeter Biodegr 40, 171–
182.
Zheng, C., Li, Y., Nie, L., Qian, L., Cai, L. & Liu, J. (2012).
Transcriptional and functional studies of a Cd(II)/Pb(II)-responsive
transcriptional regulator(CmtR) from Acidithiobacillus ferrooxidans
ATCC 23270. Curr Microbiol 65, 117–121.
Tornabene, T. G. & Peterson, S. L. (1975). Interaction of lead and
bacterial lipids. Appl Microbiol 29, 680–684.
http://mic.sgmjournals.org
Edited by: S. Spiro
Downloaded from www.microbiologyresearch.org by
IP: 88.99.165.207
On: Sat, 17 Jun 2017 04:30:35
25