Machado 2009

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Applied Microbiology ISSN 1364-5072

ORIGINAL ARTICLE

Removal of heavy metals using a brewer’s yeast strain of


Saccharomyces cerevisiae: advantages of using dead
biomass
M.D. Machado1,2, S. Janssens1,3, H.M.V.M. Soares2 and E.V. Soares1,4
1 Bioengineering Laboratory, Chemical Engineering Department, Superior Institute of Engineering from Porto Polytechnic Institute,
Rua Dr António Bernardino de Almeida, Porto, Portugal
2 REQUIMTE-Department of Chemical Engineering, Faculty of Engineering of Porto University, Rua Dr Roberto Frias, Porto, Portugal
3 KaHo St.-Lieven, Industrial Engineering, Department of Biochemistry-Microbiology, Gebroeders Desmetstraat 1, Gent, Belgium
4 IBB-Institute for Biotechnology and Bioengineering, Centre for Biological Engineering, Universidade do Minho, Campus de Gualtar,
Braga, Portugal

Keywords Abstract
bioremediation, copper, nickel,
Saccharomyces cerevisiae, yeast flocculation, Aim: The capacities of live and heat-killed cells of Saccharomyces cerevisiae at
zinc. 45C for the removal of copper, nickel and zinc from the solution were com-
pared.
Correspondence Methods and Results: Kinetic studies have shown a maximum accumulation of
Eduardo V. Soares, Bioengineering Laboratory,
Ni2+ and Zn2+ after 10 min for both types of cells, while for Cu2+ this was
Chemical Engineering Department, Superior
Institute of Engineering from Porto Polytechnic
attained after 30 and 60 min for dead and live cells, respectively. Equilibrium
Institute, Rua Dr António Bernardino de studies have shown that inactivated biomass displayed a greater Zn2+ and Ni2+
Almeida, 431, 4200-072 Porto, Portugal. accumulation than live yeasts. For Cu2+, live and dead cells showed similar
E-mail: evs@isep.ipp.pt accumulation. Fluorescence, scanning electron microscopy and infrared spec-
troscopy studies have shown that no appreciable structural or molecular
Helena M.V.M. Soares, REQUIMTE-Department changes occurred in the cells during the killing process. The increased metal
of Chemical Engineering, Faculty of
uptake observed in dead cells can be most likely explained by the loss of
Engineering of Porto University, Rua Dr
Roberto Frias, s ⁄ n, 4200-465 Porto, Portugal.
membrane integrity, which allows the exposition of further metal-binding sites
E-mail: hsoares@fe.up.pt present inside the cells.
Conclusions: Heat-killed cells showed a higher degree of heavy metal removal
2008 ⁄ 0292: received 20 February 2008, than live cells, being more suitable for further bioremediation works.
revised 2 June 2008 and accepted 16 June Significance and Impact of the Study: Dead flocculent cells can be used in a
2008 low cost technology for detoxifying metal-bearing effluents as this approach
combines an efficient metal removal with the ease of cell separation.
doi:10.1111/j.1365-2672.2009.04170.x

and tend to circulate; thus, they remain for a long time


Introduction
in nature and are accumulated through the food chain,
Heavy metal pollution is a growing problem mainly which creates a threat to public health.
caused by industrialization, which concerns particularly Metal-bearing effluents require a pretreatment step
developed countries. Electroplating, surface finishing, before being discharged in a water body or directed to a
metallurgy, mining, mineral and electronic processing are municipal sewage treatment plant. Conventional techno-
examples of industries that produce large quantities of logies, namely chemical and physical treatments, which
wastewaters containing high concentrations of heavy include precipitation-filtration, ion exchange and mem-
metals. The problem of heavy metal pollution is basically brane technologies, have been proposed for the removal
associated with: (i) acute toxicity linked with particular of heavy metals from wastewaters. For large volumes of
metals (Cu, Hg or Cd), even in lower concentration; (ii) wastewater containing relative low metal concentrations,
the fact that heavy metals are not degraded or destroyed these technologies are not adequate (the metal removal is

ª 2009 The Authors


1792 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

incomplete) or often economically prohibitive and ⁄ or more Cu2+ than the isogenic (except for the marker genes
impracticable. In the case of precipitation, the technology and the gene FLO1), nonflocculent strain (Soares et al.
most widely used, toxic sludges are generated. Thus, the 2002). Most likely, heavy metals can also occupy lectin
problem of metallic pollution of these hazardous sludges Ca2+-binding sites with the consequent increase of their
remains owing to the difficulties of disposal (Volesky uptake; thus, cell walls of flocculent cells may provide
2003). additional metal-binding sites than the nonflocculent
Biological processes are an alternative to the traditional ones. Consequently, flocculent yeast cells have a higher
treatments. It has been reported that algae, bacteria, fila- capacity of metal accumulation than the nonflocculent
mentous fungi and yeast have the ability to remove heavy cells (Soares et al. 2002). Owing to these reasons, floccu-
metals from solutions (Volesky and Holan 1995; Bakkalo- lent yeast cells are a promising low-cost biosorbent, which
glu et al. 1998; Vieira and Volesky 2000; Wang and Chen combines a better efficiency in the removal of metals with
2006). The application of microbial biomass presents an inexpensive and rapid separation of biomass. This
several advantages such as low operating cost, minimiza- allows the use of different configurations of suspended
tion of the volume of chemical and ⁄ or biological sludges biomass reactors without the risk of washout (Soares
produced and high efficiency in detoxifying dilute et al. 2002; Machado et al. 2008).
effluents (Volesky 2001). Among the different kinds of The use of live or dead cells for heavy metal removal
biomass available, yeast cells of Saccharomyces cerevisiae from wastewaters is a debatable question. Live biomass,
constitute a good alternative wastewater treatment mainly especially yeast cells, may have a higher metal-accumulat-
because of the following reasons: (i) it is generally recog- ing capacity than dead cells owing to the transport of
nized as a safe micro-organism and can be used without metals across the cell membrane (Norris and Kelly 1977;
public concern; (ii) it is available in large quantities at a Avery and Tobin 1992; Volesky et al. 1993). On the other
very low cost as it is a by-product of fermentation indus- hand, some authors have emphasized the advantages of
tries (brewing and wine); (iii) it has the ability to accu- using dead biomass for industrial applications such as
mulate a broad range of heavy metals under a wide range ease of storage, insensitiveness to metal toxicity and the
of external conditions (Blackwell et al. 1995; Volesky and fact that no nutrient supply is required (Gadd 1990, 1993;
May-Phillips 1995; Ferraz and Teixeira 1999; Kim et al. Volesky 1990). In order to enhance metal uptake capaci-
2005; Wang and Chen 2006). ties, different techniques for killing and processing bio-
A common problem of biomass application is usually mass, such as the treatment of S. cerevisiae cells with
associated with the difficulty of separation from the reac- alkaline solutions, detergents, ethanol, heat drying or the
tion mixture. To solve this problem, immobilized micro- modification of surface biomass with cetyl trimethyl
bial biomass has been used (Tsezos 1990). Biomass ammonium, were developed (Gadd 1990; Avery and
immobilization allowed its use in well-known chemical Tobin 1992; Lu and Wilkins 1996; Ashkenazy et al. 1997;
engineering reactor configurations such as packed or flu- Bingol et al. 2004; Göksungur et al. 2005). However, it
idized bed reactors (Tsezos 2001). Nevertheless, immobili- was reported that hot alkali yeast treatment can moder-
zation techniques are very expensive when used in large ately decrease Cu2+ accumulation, when compared with
scale, are unsuitable at high pH, can present mass transfer native (untreated) yeast cells (Brady et al. 1994a); simi-
limitations and result in decreased efficiency of metal larly, other authors reported that dead biomass shows a
adsorption and cell viability (Cassidy et al. 1996). lower Sr2+ (Avery and Tobin 1992) or Pb2+ uptake (Suh
A different approach may involve the use of flocculent and Kim 2000) than the respective live cells.
yeast cells. Yeast flocculation is a well-known phenome- In this work, a detailed comparison (using kinetic and
non usually associated with brewer’s strains (Verstrepen equilibrium uptake studies) of the ability of live and
et al. 2003). Flocculation involves the interaction of spe- dead (killed at 45C) flocculent cells of S. cerevisiae to
cific cell wall proteins, called ‘lectins’ (Miki et al. 1982). accumulate nickel, copper and zinc was carried out. To
The N-terminal part of this lectin-like protein stick out of our knowledge, this is the first work that compares,
the cell walls of flocculent cells and binds mannose chains under identical conditions, the accumulation of different
(receptors) present in the cell walls of neighbouring cells. metals by live and killed cells at mild temperature. Fur-
Calcium ions are required to ensure the correct confor- thermore, possible structural and molecular modifica-
mation of the lectins (Miki et al. 1982; Bony et al. 1998; tions of dead biomass were examined in order to
Kobayashi et al. 1998). The increase of cell surface hydro- elucidate the different accumulation abilities of dead and
phobicity (Jin et al. 2001) or the presence of 3-OH oxyli- live biomass. Finally, the advantages of using flocculent
pins (Strauss et al. 2005) has been positively correlated heat-killed biomass in the development of a clean
with the promotion of flocculation. In recent years, it was technology for heavy metals removal from industrial
shown that a flocculent strain of S. cerevisiae accumulated effluents are discussed.

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1793
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

Cell suspensions were agitated at 150 rev min)1, in an orbi-


Materials and methods
tal shaker (Braun Certomat S, Melsungen, Germany), at
25C. At defined intervals of time, samples (10 ml) were
Strains, media and culture conditions
taken and filtered through a 0Æ45-lm pore size filter. Con-
In this work, the flocculent brewing strain of S. cerevisiae trol experiments in the absence of biomass and in the same
National Collection of Yeast Culture (NCYC) 1364 was experimental conditions described before were performed;
used. The strain was routinely maintained at 4C on YEPD these controls showed that no appreciable amount of
agar slants (10 g yeast extract per litre, 20 g peptone per metals (less than 1%) was adsorbed to the plastic flasks and
litre, 20 g glucose per litre and 20 g agar per litre). to the filter membrane. At the end of kinetic studies, no
Precultures were prepared in 40 ml of YEPD broth in bacterial contamination, checked by phase-contrast micro-
100 ml Erlenmeyer flasks. The cells were incubated at scopy, was observed in the yeast cell suspensions.
25C on an orbital shaker, Sanyo Gallenkamp IOC 400 Heavy metals were determined in the filtrates by
(West Sussex, UK), at 150 rev min)1, for 24 h. atomic absorption spectroscopy (AAS) with flame atom-
Cultures in YEPD with 50 g glucose per litre were pre- ization in a Perkin Elmer Analyst 400 spectrometer (Nor-
pared by inoculating 1 l of the culture medium in 2 l walk, CT, USA), after appropriate dilution of the samples.
Erlenmeyer flasks with 4% (v ⁄ v) from precultures. Cells The amount of metal uptake, q (lmol of metal per gram
were incubated in the same conditions as the preculture of dry weight biomass), at each reaction time was calcu-
for 48 h. lated using the following equation:
q ¼ V  ðCi  Ct Þ  1000=ðm  MÞ; ð1Þ
Preparation of cell suspensions where V is the volume of the reaction mixture (l); Ci and
After growth, cells were harvested by centrifugation (2000 g, Ct are the metal concentrations in the solution (mg l)1)
5 min) and washed twice with 30 mmol l)1 ethylenedi- at the initial and defined periods of time indicated in the
aminetetraacetic acid (EDTA) solution (Merck). Subse- figures, respectively; m is the amount of the added dry
quently, the cells were washed once with deionized water, weight biomass (g) and M is the molar mass of the metal.
once with 10 mmol l)1 of MES [2-(N-morpholino)ethane-
sulfonic acid] pH buffer (Sigma) at pH 6Æ0 and resuspended Equilibrium studies
in MES pH buffer. MES is a suitable pH buffer for heavy Equilibrium metal studies were carried out in 100-ml plas-
metal uptake studies because it does not complex with tic flasks containing 30 ml of the cell suspension (4 g l)1
several metal ions, such as cadmium, copper, lead and zinc dry weight biomass) in MES pH buffer (10 mmol l)1 at pH
(Soares et al. 1999a,b). Additionally, yeast cells maintain 6Æ0) with an appropriate concentration of the metal
viability and no detectable amounts of protein or inorganic (between 5 and 200 mg l)1 for nickel and from 5 to
phosphate are released when incubated in this buffer during 50 mg l)1 for copper and zinc). Cell suspensions were
48 h (Soares et al. 2000). The dead cells were obtained by shaken at 150 rev min)1 at 25C. After 1 h of contact of
drying live biomass at 45C until constant weight. the biomass with metals, the samples were collected and
filtered through a 0Æ45-lm pore size filter. The filtrates were
analysed for metal equilibrium concentration by AAS.
Determination of biomass
Cell concentration was determined spectrophotometrically Equilibrium model
(Unicam, Helios c) at 600 nm after appropriate dilution of The Langmuir adsorption model was used to evaluate the
the samples with EDTA solution (30 mmol l)1) to prevent metal uptake of live and dead biomass at 45C:
cell aggregation. Calibration curves (absorbance vs either q ¼ qmax bCeq =ð1 þ bCeq Þ;
number of cells or dry weight) were previously made.
where q is the metal-specific uptake (lmol of metal per
gram dry weight biomass), qmax (lmol of metal per gram
Metals accumulation experiments dry weight biomass) is the maximum metal uptake biomass
Kinetic studies capacity, b is the Langmuir constant (l lmol)1 of metal)
Kinetic studies were performed in 500-ml plastic flasks and Ceq is the metal equilibrium concentration (lmol l)1).
containing 100 ml of cell suspension (4 g l)1 dry weight
biomass) in MES pH buffer (10 mmol l)1 at pH 6Æ0). Evaluation of lethality induced by heat and Cu2+
Metals (Cu, Ni or Zn) were added in a final concentration
of 5 or 100 mg l)1 from a 2000 mg l)1 [Cu (NO3)2 and Zn Cells heated at 45C until constant weight in sterile
Cl2] or 1000 mg l)1 [NiCl2] metal stock solution (Merck). conditions were suspended in deionized water at

ª 2009 The Authors


1794 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

1 · 108 cells ml)1. Four replicates of 100 ll of the cell Sigma) and resuspended in osmium tetroxide (2%;
suspension were incubated at 25C for 1 week. In the case Fluka). After 1 h of incubation at room temperature with
of lethality induced by Cu2+, cells were suspended in an agitation of 100 rev min)1, the cells were washed with
10 mmol l)1 of MES pH buffer (pH 6Æ0) in a final con- 0Æ1 mol l)1 of sodium cacodylate buffer (pH 6Æ8), centri-
centration of 4 g dry weight biomass per litre; i.e. the fuged and resuspended in deionized water.
same conditions used in the kinetic studies. Before and Cell suspensions were transferred to a microscope slide
after the addition of 5 mg l)1 of metal, samples (two rep- previously coated with 0Æ1% (w ⁄ v) of aqueous solution of
licates of 100 ll) were taken at defined intervals of time, poly-l-lysine (Sigma). Fixed cells were dehydrated by
serially diluted with sterile deionized water and plated on incubation in a series of ethanol : water mixtures [10, 20,
YEPD agar (two replicates of 100–200 ll of convenient 30, 50, 70 and 100% (v ⁄ v)] for 10 min for each sample.
dilutions). After 3–4 days of incubation at 25C, the colo- Next, the cells were dried by critical point using liquid
nies were counted; no further colonies were observed carbon dioxide as transition solvent in a Balzer’s appara-
when incubation was prolonged for 1 week. tus. Then, the samples were coated with gold using a
SEM coating unit (Polaron SC 500) and examined in a
JEOL JSM-35C SEM at 15 kV and ·10 000 magnification.
Assessing of membrane integrity
Dead cells (5 · 107 cells ml)1) were washed and resus-
Infrared spectra
pended in phosphate buffer saline (PBS; 50 mmol l)1 at
pH 7Æ4). Propidium iodide (PI; Sigma, Steinheim, Ger- Infrared spectra of live and dead cells (heated at 45C), as
many) was added to a final concentration of 4Æ5 lmol l)1. well as dead biomass exposed to copper, nickel or zinc,
The cell suspension was incubated in the dark at room were performed. Live or dead biomass (2 g l)1 dry weight
temperature for 20 min. Cells were examined using a biomass) was washed as described for the preparation of
Leica DLMB epifluorescence microscope (Leica Microsy- cell suspension. The dead biomass was suspended in MES
tems, Wetzlar GmbH, Germany) equipped with a HBO- pH buffer (10 mmol l)1 at pH 6Æ0) containing solutions
100 mercury lamp and a filter set I3 (excitation filter BP of metals at concentrations of (in mg l)1): 22Æ2 of Cu,
450-490, dichromatic mirror 510 and suppression filter 32Æ3 of Ni or 26Æ2 of Zn in order to achieve saturation of
LP 515), from Leica. The images were acquired with a biomass by the different metals. Cell suspensions were
Leica DC 300F camera (Leica Microsystems, Heerbrugg, agitated at 150 rev min)1 for 1 h, at 25C, in plastic flasks
Switzerland) using a N plan ·100 objective; the images and then dried at 25C before being analysed in an infra-
were processed using Leica IM 50-Image manager red spectrometer with Fourier transform (FTIR; Bomem
software. MB- 154 S). Disks of 100 mg of KBr containing 3%
(w ⁄ w) of finely ground power of each sample were
prepared and then examined. Spectra were measured
Staining of cell walls
between 4000 and 650 cm)1 and further treated with
Live and dead cells, at 1 · 107 cells ml)1, were washed Win-Bomem Easy ver. 3.04 level software.
twice and resuspended in water. Calcofluor white M2R
(Sigma, Steinheim, Germany) was added at a final con-
Reproducibility of the studies
centration of 25 lmol l)1. The cell suspensions were
incubated in the dark at room temperature for 30 min. All experiments were repeated independently at least
Then, the cells were washed twice with deionized water twice. In kinetic and equilibrium studies, each repetition
and examined using an epifluorescence microscope was carried out in duplicate.
equipped with a HBO-100 mercury lamp and a filter
set A (excitation filter BP 340-380, dichromatic mirror
Results
400 and suppression filter LP 425) from Leica. The
images were acquired and processed as described
Kinetic studies
before.
The kinetics of metal accumulation (Cu2+, Ni2+ and
Zn2+) by live and dead cells of S. cerevisiae was examined
Scanning electron microscopy (SEM)
in buffer solution, at pH 6Æ0, using two different concen-
Cells were fixed by 3% glutaraldehyde (w ⁄ v; Fluka) for trations of metals: 5 and 100 mg l)1. These concentra-
2 h with an agitation of 100 rev min)1 at room tempera- tions were selected because they correspond to a lower
ture. Cells were then centrifuged (500 g, 5 min), washed (5 mg l)1) and a higher (100 mg l)1) metal concentra-
thrice with 0Æ1 mol l)1 sodium cacodylate buffer (pH 6Æ8; tion found in electroplating wastewater streams. Copper

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1795
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

kinetics studies were only performed for the lowest con-


centration owing to early precipitation at pH 6 (Hogfeldt
1982). Kinetic studies are important both from an
academic and applied point of views, as the information
is useful for determining the contact time required for
equilibrium uptake studies and for removal of the metals
from the effluents.
For an initial copper concentration of 5 mg l)1, after
15 min of contact time of live or dead biomass, 38% and
44% of the metal was removed. This represents an accu-
mulation of c. 7Æ0 and 8Æ3 lmol g)1 dry weight biomass,
respectively (Fig. 1a). Increasing the contact time to
30 min resulted in a slight increase of metal removal:
41% and 50% of copper was removed by live and dead
cells, which represents an accumulation of c. 7Æ6 and
9Æ6 lmol g)1 dry weight biomass of Cu2+, respectively.
For dead cells, equilibrium was attained after 30 min
while with live cells equilibrium was only observed after
60 min where an accumulation of c. 9Æ1 lmol g)1 dry
weight biomass was obtained, which represents 49% of
Cu2+ removal (Fig. 1a). Extending the incubation time
between biomass (live or dead) and metals up to 24 h, no
increase of metal uptake was observed (data not shown).
For live and dead biomass, nickel and zinc kinetics
accumulation were faster than with copper. For both
nickel concentrations tested, the maximum nickel uptake
occurred after 10 min of contact time (Figs 1b and 2a).
For 5 mg l)1 of nickel concentration, live and killed cells
accumulated 3Æ7 and 6Æ4 lmol g)1 dry weight biomass of
Ni2+, respectively (Fig. 1b); for 100 mg l)1 of Ni2+ con-
centration, live and heat-killed cells accumulated 24 and
84 lmol g)1 dry weight biomass Ni2+, respectively
(Fig. 2a). For the lower zinc concentration (5 mg l)1), the
same behaviour was observed: live or heat-killed biomass
accumulated 3Æ8 and 11 lmol g)1 dry weight biomass,
respectively, after 10 min of contact time (Fig. 1c). For
the highest zinc concentration (100 mg l)1), dead cells
presented a slight increase in metal removal after 1 and
3 h of contact time [44% and 48%, which corresponds to
c. 142 and 160 lmol g)1 dry weight biomass of Zn2+,
respectively (Fig. 2b)] when compared with the value
obtained after 10 min of contact time [40%, which corre-
sponds to c. 128 lmol g)1 dry weight biomass of Zn2+
(Fig. 2b)]. Similar to what happened with copper, the

Figure 1 Accumulation of Cu2+ (a), Ni2+ (b) and Zn2+ (c) by live (h)
or dead ( ) cells of Saccharomyces cerevisiae NCYC 1364. Cells were
suspended in 10 mmol l)1 of 2-(N-morpholino)ethanesulfonic acid
buffer at pH 6Æ0 in a final concentration of 4 g l)1 dry weight
biomass; the initial metal concentration was 5 mg l)1. Each point
represents the mean of at least two independent experiments
performed in duplicate; SD are presented (n ‡ 4).

ª 2009 The Authors


1796 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

1364, equilibrium experiments were carried out at fixed


pH (6Æ0) and temperature (25C). As was previously
established in kinetic studies reported before (Figs 1 and
2), a contact time of 60 min was used to ensure that
equilibrium between live or dead cells and all three metals
was attained.
Metals uptake (q) by the two types of biomass was
plotted against the equilibrium metal concentration (Ceq)
present in solution (adsorption isotherms; Fig. 3). For live
and dead biomass, Ni2+ or Zn2+ uptake increased with
metal concentrations until biomass reached saturation
(Fig. 3b,c). Heat-inactivated biomass displayed a greater
degree of Ni2+ and Zn2+ accumulation than live yeasts
(Fig. 3b,c). In the case of Cu2+, a biphasic behaviour
seems to occur for both types of biomass studied;
however, no apparent saturation of biomass-binding sites
seems to be achieved (Fig. 3a).
In order to characterize metal uptake by the two types
of yeast biomass, Langmuir model was applied to the
experimental data. Unless for live biomass in the presence
of Zn2+, for which a poor correlation was found
(r = 0Æ854), the Langmuir model fitted quite well into the
experimental data (Table 1). The comparative analysis of
qmax values (this parameter provides information about
the maximum metal accumulation capacity) showed that
heat-inactivated biomass accumulated more Ni2+ and
Zn2+ (about 17 and 14 more times, respectively) than live
biomass (Table 1). For copper, live and dead biomass
displayed similar accumulation capacities: 122 and
156 lmol g)1 dry weight biomass of Cu2+, respectively
(Table 1). For live biomass, metal uptake ability
decreased in the following order: Cu2+ >> Zn2+ and Ni2+;
for dead biomass, a similar accumulation pattern was
found.

Accessing of structural and molecular modifications of


Figure 2 Accumulation of Ni2+ (a) and Zn2+ (b) by live (h) or dead dead biomass
( ) cells of Saccharomyces cerevisiae NCYC 1364. Cells were
suspended in 10 mmol l)1 of 2-(N-morpholino)ethanesulfonic acid For the purpose of explaining modification of metals
buffer at pH 6Æ0 in a final concentration of 4 g l)1 of dry weight uptake by dead biomass when compared with live cells,
biomass; the initial metal concentration was 100 mg l)1. Each point different studies for assessing structural and molecular
represents the mean of at least two independent experiments
modifications of heat-treated biomass were conducted.
performed in duplicate; SD are presented (n ‡ 4).
Membrane integrity was assessed using PI. PI is a mem-
brane-impermeant stain and is generally excluded from
extension of incubation time up to 24 h resulted in no viable cells. Cells with damaged plasma membrane incor-
increase in nickel or zinc removal by live or dead biomass porate PI; this stain binds to DNA by intercalating
for either metal concentration tested (data not shown). between the bases with little or no sequence and exhibits
an orange–red fluorescence (Haugland 2005). As it can be
seen in Fig. 4, all cell population heat-treated at 45C
Equilibrium studies
incorporates PI as a consequence of membrane integrity
For the objective of comparing metal (Cu2+, Ni2+ and lost during the thermal treatment. This killing effect was
Zn2+) uptake performance by live and heat-treated cells additionally confirmed by a cultural procedure: cell
of ale brewing flocculent strain of S. cerevisiae NCYC suspension dried at 45C was incubated for 1 week at

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1797
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

Table 1 Langmuir parameters for Cu2+, Ni2+ and Zn2+ removal

qmax
(lmol g)1 dry b
weight (l lmol)1
biomass) metal) r

Metal Live Dead Live Dead Live Dead

Cu2+ 122 156 2Æ1 1Æ3 0Æ992 0Æ945


Ni2+ 8 134 11Æ0 1Æ7 0Æ957 0Æ999
Zn2+ 12 163 5Æ0 2Æ8 0Æ854 0Æ990

25C in YEPD agar plates; no colonies were found, which


demonstrated that all population was dead.
Cell wall modifications owing to thermal treatment
were assessed using calcofluor white. Calcofluor white
binds to chitin with high specificity (Pringle 1991). Chitin
is mostly located in the scar region although some of it is
dispersed throughout the yeast cell wall (Cabib et al.
2001). Dead cells, after being stained with calcofluor
white, exhibited a typical uniform faint fluorescence on
their surface with well-visible bud scars (Fig. 5). This sug-
gests that no redistribution of chitin occurred during the
thermal treatment. Maintenance of morphological charac-
teristics of dead cells was further confirmed by SEM. No
significant differences between live and dead cells were
observed; characteristic bud scars were well visible in dead
cells (Fig. 6).
Molecular modifications in heat-treated biomass were
investigated by FTIR spectroscopy. The spectra contain
information of yeast components, which are represented
by specific peaks in the fingerprinting regions: region
between 790 and 1180 cm)1 corresponding to the sugar
component of the cells; between 1200 and 1290 cm)1 cor-
responding to the nucleic acids; between 1400 and
1700 cm)1 representing the protein component; between
2500 and 3800 cm)1 corresponding to –OH and –NH
groups and hydrogen bonding; finally, peaks at
2900 cm)1 may be because of chitin (Brugnerotto et al.
2001; Galichet et al. 2001; Padmavathy et al. 2003). The
spectra of live and dead biomass were very similar and all
the characteristic peaks owing to the main yeast macro-
molecules were present in both (Fig. 7). More specifically,
it is possible to observe a peak with a maximum at
1074 cm)1 (dead cells), which can mainly be attributed to
the b(1 fi 3) glucan, the major structural component of
the cell wall (30–45% of wall mass). In addition, two
Figure 3 Cu2+ (a), Ni2+ (b) and Zn2+ (c) isotherms for live (h) or dead important peaks were found in dead cells: a peak at
( ) cells of Saccharomyces cerevisiae NCYC 1364. Cells were sus-
1541 cm)1 owing to the presence of protein amide II
pended in 10 mmol l)1 of 2-(N-morpholino)ethanesulfonic acid buffer
at pH 6Æ0 in a final concentration of 4 g l)1 dry weight biomass. The
band (mainly C–N stretching plus N–H bonding) and a
results were obtained after 1 h of contact time between biomass and peak at 1634 cm)1 owing to the protein amide I band
metals. Each point represents the mean of at least two independent (mainly C=O stretching and the contribution of N–H
experiments performed in duplicate; SD are presented (n ‡ 4). bonding). The peak with a maximum at 3416 cm)1,

ª 2009 The Authors


1798 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

Interaction of Cu2+, Ni2+ and Zn2+ with dead biomass


The interaction of Cu2+, Ni2+ and Zn2+ with dead bio-
mass was also investigated by FTIR technique. As can be
seen in Fig. 8, biomass loaded with the different metals
maintained the characteristic peaks of yeast. After metal
accumulation, dead biomass showed a remarkable
decrease of the band at 2500–3800 cm)1 (Fig. 8) and a
shift of the maximum of the peak at 3416 cm)1, recorded
in the absence of metals, to 3364, 3294 and 3285 cm)1, in
the case of biomass loaded with Cu2+, Zn2+ and Ni2+,
respectively. This was probably because of the interaction
of –NH and –OH groups with metals. A small decrease
of the peak at 2928 cm)1 was also observed with all the
biomass loaded with the different metals (Fig. 8). This
was probably because of the interaction of the chitin,
present in the cell wall, with metals. In addition, a
decrease of the peaks in the region of the proteins, 1634
and 1541 cm)1 (in the absence of metals), associated with
a shift of their maxima to 1649 (biomass loaded with
Ni2+ and Zn2+), 1651 (biomass loaded with Cu2+) and to
1539 cm)1 (biomass loaded with Ni2+ and Zn2+) was
observed. Finally, a decrease of the peaks in the region of
the carbohydrates, namely at 1074 cm)1 (in the absence
of the metals), associated with a shift of the maximum to
1080 cm)1 (biomass loaded with Cu2+) was observed.
This was probably because of the interaction of glucans
from the cell wall with the metals (Fig. 8).
Figure 4 Fluorescence microscopy of cells of Saccharomyces cerevisiae
NCYC 1364 heat-killed at 45C and stained with propidium iodide
Discussion
(a) and phase-contrast observation of the same cells (b) are shown.
The knowledge of metal accumulation characteristics by
recorded in dead cells, could correspond to hydroxyl yeast biomass is of fundamental importance for further
groups bounded with NH groups of the secondary amide use in bioremediation processes. Kinetic and equilibrium
group of chitin (Padmavathy et al. 2003).

Figure 5 Fluorescence microscopy of Saccha-


romyces cerevisiae NCYC 1364 heat-killed at
45C and stained with calcofluor white M2R
(a); arrow, bud scar. Phase-contrast observa-
tion of the same cells (b) are shown.

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1799
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

Figure 7 Infrared spectra of live and dead cells of Saccharomyces ce-


revisiae NCYC 1364. Cells heated at 45C (dead cells) (a) and live cells
(b).

Figure 6 Scanning electron microscopy of Saccharomyces cerevisiae


NCYC 1364. Live (a) or heat-killed cells at 45C (b); arrow, bud scar.

studies are usually carried out for quantitative assessment


of biomass performance as well as for process design.
Yeast metal uptake usually occurs in two steps: the first
step (biosorption) is fast (it occurs in the first few min-
utes of contact with the metal), is independent of
metabolism and happens in live and dead cells; the sec-
ond step (bioaccumulation) is generally considered
metabolism-dependent (it occurs only in live cells) and is
attributed to intracellular metal uptake across the cell Figure 8 Infrared spectra of Saccharomyces cerevisiae NCYC 1364
membrane (Blackwell et al. 1995). Nickel and zinc accu- after being exposed to different heavy metals. Dead cells (control)
(a), dead cells after Zn2+ (b), Ni2+ (c) and Cu2+ exposition (d). Dead
mulation by live or dead biomass of S. cerevisiae NCYC
biomass was suspended in 10 mmol l)1 of 2-(N-morpholino)ethane-
1364 was very fast and occurred within 10 min; although sulfonic acid buffer at pH 6Æ0 in a final concentration of 2 g l)1 of dry
both metals are actively transported in S. cerevisiae (Fuhr- weight biomass. The initial metal concentration was: 22Æ2 Cu2+
mann and Rothstein 1968), no further accumulation was mg l)1, 32Æ3 Ni2+ mg l)1 and 26Æ2 Zn2+ mg l)1. The cells were
observed in live cells (Fig. 1) and no appreciable toxic incubated in the metal solutions for 60 min at 25C.

ª 2009 The Authors


1800 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

effects could be found for this concentration (5 mg l)1); maintain their characteristics: chitin remained in the yeast
(Mowll and Gadd 1983; Soares et al. 2003). The absence scars after thermal treatment (Fig. 5). The maintenance of
of a bioaccumulation step in live cells was probably owing morphological characteristics of dead cells was further
to the fact that yeast cells were not incubated in the pres- confirmed by SEM, where no significant differences
ence of an external energy source, such as glucose. Most between live and dead cells were observed (Fig. 6). In
likely, yeast internal energetic reserves were not sufficient addition, the infrared spectra of live and dead biomass
to provoke a supplementary and detectable accumulation were very similar: all characteristic peaks attributed to the
of nickel and zinc. Consistent with this hypothesis, an main yeast macromolecules of cell wall were present in
enhancement of Ni2+ and Zn2+ uptake by yeast cells pre- both spectra (Fig. 7). As heat-killed cells have shown a
treated with glucose comparative with those washed loss of plasma membrane integrity, when assessed by PI
untreated (starved) cells was described (Fuhrmann and (Fig. 4), the increase of nickel and zinc removal by dead
Rothstein 1968; Stoll and Duncan 1996). cells could most likely be attributed to the exposition of
In the previous work, it was found that the killing in further metal-binding sites present inside the cells.
mild conditions (45C) is the most appropriate tempera- According to this hypothesis, an increase in the Cu2+
ture to inactivate yeast cells (Machado et al. 2008). Under accumulation by Penicillium spinulosum and of U accu-
these conditions, yeast cells maintained their flocculation mulation by S. cerevisiae cells permeabilized by the action
properties, when compared with live cells, most likely of detergents (Gadd 1990) or by the action of HCHO or
because no denaturation of the lectin-like proteins HgCl2 (Strandberg et al. 1981), respectively, was also
occurred at this temperature (Machado et al. 2008). In described.
the present work, heavy metal removal by live and dead In the case of copper, live and dead cells showed simi-
cells (inactivated at 45C) was quantitatively assessed lar removal. This can be explained by the high copper
using equilibrium metal uptake studies. Taking into toxicity, which provokes lesions of yeast cell membrane
account the maximum metal uptakes (Qmax.) determined (Soares et al. 2003). In the case of the strain of S. cerevisi-
at 25C and pH 6Æ0, dead biomass was able to accumulate ae used in the present work, 5 mg l)1 of Cu2+ induced a
higher quantities of Ni2+ and Zn2+ (17 and 14 times lethality of 50% after 20 min of contact time and about
higher, respectively) than live biomass (Table 1); for 70% after 60 min (data not shown). Thus, live cells were
Cu2+, live and dead cells showed a similar metal uptake progressively converted into dead cells and copper accu-
capacity. This illustrates that biomass killed at 45C is mulation by the initial live cells became similar to dead
more suitable for bioremediation processes. cells (Fig. 1).
In order to understand why inactivated cells showed a Infrared analyses of dead biomass after being exposed
higher metal accumulation capacity, different studies, to different metals showed a decrease in the peaks of the
namely at the yeast cell wall and membrane levels, were fingerprinting regions (Fig. 8). This suggests the involve-
performed. The aim of these studies was to assess whether ment of carboxyl, amino, hydroxyl and amide groups of
structural and molecular modifications of heat-treated protein and carbohydrate fractions (most likely of
biomass occurred. The cell wall of S. cerevisiae is the first mannoproteins, glucans and chitin) of the cell wall in the
cellular structure to be in contact with metal ions and yeast metal uptake. These results are in agreement with
presents a complex macromolecular structure with a lay- those obtained by Padmavathy et al. (2003), who sug-
ered organization constituted by an amorphous inner and gested that mannoproteins and glucans present on the cell
a fibrilar outer layer. The inner layer is mainly composed wall were responsible for the sorption of Ni2+. Brady and
of b-glucan and chitin. b-glucans constitute by fractions Duncan (1994) had shown that blocking amino, carboxyl
of b(1 fi 3) and b(1 fi 6)-linked glucose residues; chitin or hydroxyl groups of the cell walls reduces the accumu-
is composed of linear chains of b(1 fi 4)-linked to lation capacity of Cu2+. These facts clearly indicated the
N-acetylglucosamine. Chitin is present as small amounts participation of these functional groups in Cu2+ binding.
in the wall (1Æ5–6%) predominantly located in the bud The comparison of metal uptake by S. cerevisiae NCYC
scars. The outer layer consists predominantly of a-mann- 1364 with other S. cerevisiae strains is very difficult partic-
ans highly glycosylated and associated with proteins ularly owing to the different operational conditions such
(mannoproteins), which corresponds to 30–50% of the as biomass pretreatment, biomass concentration, pH and
cell wall mass (Cabib et al. 2001; Lesage and Bussey contact time between biomass and metals. Table 2 shows
2006); these structures appear to be very important in metals accumulation by different types of S. cerevisiae
heavy metal accumulation (Brady et al. 1994b). Calcoflu- biomass previously reported in the literature. The extent
or white has been very useful for assessing abnormal pat- of copper uptakes by live or dead cells of S. cerevisiae
terns of cell wall deposition (Pringle 1991). In the present NCYC 1364 as determined in this work is similar to the
study, assays with calcofluor white show that dead cells values presented in Table 2. For the other two metals,

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1801
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

Table 2 Comparison of metal uptake capacities by Saccharomyces cerevisiae

Metal uptake capacity


(lmol g)1 dry weight Time of
Metal Type of S. cerevisiae biomass biomass) pH contact* (h) Reference

Cu Live baker’s yeast 200 4 16 Volesky and May-Phillips (1995)


Live brewer’s yeast 160 4 16 Volesky and May-Phillips (1995)
Dead brewer’s yeast 78 5Æ5 1 Bakkaloglu et al. (1998)
Dead biomass derived from 90 ** 2 Bustard and McHale (1998)
whiskey distillery
Live baker’s yeast 100 6 24 Al-Saraj et al. (1999)
Ni Dead brewer’s yeast 25 5Æ5 1 Bakkaloglu et al. (1998)
Live baker’s yeast 136 6 24 Al-Saraj et al. (1999)
Dead biomass 789 5 24 Özer and Özer (2003)
Dead and deactivated protonated 194 6Æ75 24 Padmavathy et al. (2003)
baker’s yeast
Zn Dead brewer’s yeast 53 6 1 Bakkaloglu et al. (1998)
Dead biomass derived from 258 ** 2 Bustard and McHale (1998)
whiskey distillery
Live baker’s yeast 290 5 24 Al-Saraj et al. (1999)

*Time of contact of the biomass with metals.


**Not reported in the publication.

accumulation capacities reported in Table 2 vary widely; Infrared studies suggested the participation of carboxyl,
the values of metal uptake by dead cells of S. cerevisiae amino, hydroxyl and amide groups in metal yeast uptake.
NCYC 1364 determined in this work presents intermedi-
ary values of accumulation when compared with these
Acknowledgements
studies.
A widespread difficulty associated with the industrial The authors thank Fundação para a Ciência e a Tecnologia
use of micro-organisms in different remediation technol- (FCT) of the Portuguese Government for the financial
ogies is linked with their small size and low density; support to this work with FEDER founds, by the Project
these characteristics can limit the choice of suitable POCTI ⁄ CTA ⁄ 47875 ⁄ 2002. Manuela D. Machado is also
reactors and difficult cell separation from the reaction gratefully acknowledged for a grant scholarship financed
mixture after the effluent has been treated. The use of under the same project and another grant from the
flocculent brewing yeast biomass is a natural, easy and FCT (SFRH ⁄ BD ⁄ 31755 ⁄ 2006). Steven Janssens (KaHo
cheap method of cell separation from the treated efflu- St. Lieven, Gent, Belgium) wishes to thank Dr Luc
ents, which overcomes the need for cell immobilization De Cooman for the opportunity to participate in the
(like gel immobilization). Yeast flocculation also facili- ERASMUS bilateral agreement programme between his
tates further recycling of the biomass and recovery of school and ISEP (Portugal).
the metals.
In conclusion, dead yeast cells (at 45C) evidenced
References
higher metals uptake capacities than live cells. This fact
points out that heat-killed biomass at 45C seems to be a Al-Saraj, M., Abdel-Latif, M.S., El-Nahal, I. and Baraka, R.
promising biomass for further application in metal bio- (1999) Bioaccumulation of some hazardous metals by
remediation processes. This increase of metal uptake sol-gel entrapped microorganisms. J Non-Cryst Solids 248,
capacity is most likely explained by the loss of cell mem- 137–140.
brane integrity; this occurs during the killing process and Ashkenazy, R., Gottlib, L. and Yannai, S. (1997) Characteriza-
allows the exposition of further metal-binding sites pres- tion of acetone-washed yeast biomass functional groups
ent inside the cells. Kinetic studies have shown that a involved in lead biosorption. Biotechnol Bioeng 55, 1–10.
contact time of 60 min is enough to achieve the equilib- Avery, S.V. and Tobin, J.M. (1992) Mechanism of strontium
uptake by laboratory and brewing strains of Saccharomyces
rium between metals and dead biomass. Fluorescent
cerevisiae. Appl Environ Microbiol 58, 3883–3889.
microscopy and SEM studies, as well as infrared spectros-
Bakkaloglu, I., Butter, T.J., Evison, L.M., Holland, F.S. and
copy, have shown that no appreciable structural or
Hancock, I.C. (1998) Screening of various types biomass
molecular changes occurred during the killing process.

ª 2009 The Authors


1802 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804
M.D. Machado et al. Advantages of using dead biomass in heavy metals removal

for removal and recovery of heavy metals (Zn, Cu, Ni) by Haugland, R.P. (2005) Assays for cell viability, proliferation
biosorption, sedimentation and desorption. Water Sci and function. In The Handbook – A Guide to Fluorescent
Technol 38, 269–277. Probes and Labeling Technologies, 10th edn ed. Spence,
Bingol, A., Ucun, H., Bayhan, Y.K., Karagunduz, A., Cakici, A. M.T.Z. pp. 699–776. USA: Invitrogen Corp.
and Keskinler, B. (2004) Removal of chromate anions Hogfeldt, E. (1982) Stability Constants of Metal–Ion Complexes
from aqueous stream by a cationic surfactant-modified – Part A: Inorganic Ligands. pp. 49. Oxford: Pergamon
yeast. Biores Technol 94, 245–249. Press.
Blackwell, K.J., Singleton, I. and Tobin, J.M. (1995) Metal cat- Jin, Y., Ritcey, L.L. and Speers, R.A. (2001) Effect of cell sur-
ion uptake by yeast: a review. Appl Microbiol Biotechnol 43, face hydrophobicity, charge, and zymolectin density on the
579–584. flocculation of Saccharomyces cerevisiae. J Am Soc Brew
Bony, M., Barre, P. and Blondin, B. (1998) Distribution of the Chem 59, 1–9.
flocculation protein, flop, at the cell surface during yeast Kim, T.Y., Park, S.K., Cho, S.Y., Kim, H.B., Kang, Y., Kim,
growth: the availability of flop determines the flocculation S.D. and Kim, S.J. (2005) Adsorption of heavy metals by
level. Yeast 14, 25–35. brewery biomass. Korean J Chem Eng 22, 91–98.
Brady, D. and Duncan, J.R. (1994) Binding of heavy metals by Kobayashi, O., Hayashi, N., Kuroki, R. and Sone, H. (1998)
the cell walls of Saccharomyces cerevisiae. Enzyme Microb Region of Flo1 proteins responsible for sugar recognition.
Technol 16, 633–638. J Bacteriol 180, 6503–6510.
Brady, D., Stoll, A. and Duncan, J.R. (1994a) Biosorption of Lesage, G. and Bussey, H. (2006) Cell wall assembly in Saccha-
heavy metal cations by non-viable yeast biomass. Environ romyces cerevisiae. Microbiol Mol Biol Rev 70, 317–343.
Technol 15, 429–438. Lu, Y. and Wilkins, E. (1996) Heavy metal by caustic-treated
Brady, D., Stoll, A. and Duncan, J.R. (1994b) Chemical and yeast immobilized in alginate. J Hazard Mat 49, 165–179.
enzymatic extraction of heavy metal binding polymers Machado, M.D., Santos, M.S.F., Gouveia, C., Soares,
from isolated cell walls of Saccharomyces cerevisiae. Biotech- H.M.V.M. and Soares, E.V. (2008) Removal of heavy met-
nol Bioeng 44, 297–302. als using a brewer’s yeast strain of Saccharomyces cerevisiae:
Brugnerotto, J., Lizardi, J., Goycoolea, F.M., Argüelles-Monal, the flocculation as a separation process. Biores Technol 99,
W., Desbrières, J. and Rinaudo, M. (2001) An infrared 2107–2115.
investigation in relation with chitin and chitosan charac- Miki, B.L.A., Poon, N.H., James, A.P. and Seligy, V.L. (1982)
terization. Polymer 42, 3569–3580. Possible mechanism for flocculation interactions governed
Bustard, M. and McHale, A.P. (1998) Biosorption of heavy met- by gene FLO1 in Saccharomyces cerevisiae. J Bacteriol 150,
als by distillery-derived biomass. Bioprocess Eng 19, 351–353. 878–889.
Cabib, E., Roh, D., Schmidt, M., Crotti, L.B. and Varma, A. Mowll, J.L. and Gadd, G.M. (1983) Zinc uptake and toxicity
(2001) The yeast cell wall and septum as paradigms of cell in the yeasts Sporobolomyces roseus and Saccharomyces cere-
growth and morphogenesis. J Biol Chem 276, 19679–19682. visiae. J Gen Microbiol 129, 3421–3425.
Cassidy, M.B., Lee, H. and Trevors, J.T. (1996) Environmental Norris, P.R. and Kelly, D.P. (1977) Accumulation of cadmium
applications of immobilized microbial cells: a review. J Ind and cobalt by Saccharomyces cerevisiae. J Gen Microbiol 99,
Microbiol 16, 79–101. 317–324.
Ferraz, A.I. and Teixeira, J.A. (1999) The use of flocculating Özer, A. and Özer, D. (2003) Comparative study of the bio-
brewer’s yeast for Cr(III) and Pb(II) removal from residual sorption of Pb(II), Ni(II), and Cr(VI) ions onto S. cerevisi-
wastewaters. Bioprocess Eng 21, 431–437. ae: determination of biosorption heats. J Hazard Mat B
Fuhrmann, G.F. and Rothstein, A. (1968) The transport of 100, 213–229.
Zn2+, Co2+ and Ni2+ into yeast cells. Biochim Biophys Acta Padmavathy, V., Vasudevan, P. and Dhingra, S.C. (2003)
163, 325–330. Biosorption of nickel (II) ions on Baker’s yeast. Process
Gadd, G.M. (1990) Fungi and Yeast for Metal Accumulation. In Biochem 38, 1389–1395.
Microbial Mineral Recovery ed. Ehrlich, H.L. and Brierly, Pringle, J.R. (1991) Staining of bud scars and other cell wall
C.L. pp. 249–275. New York: McGraw-Hill. chitin with Calcofluor. Methods Enzymol 194, 732–735.
Gadd, G.M. (1993) Interaction of fungi with toxic metals. New Soares, H.M.V.M., Conde, P.C.F.L., Almeida, A.A.N. and
Phytol 124, 25–60. Vasconcelos, M.T.S.D. (1999a) Evaluation of N-substituted
Galichet, A., Sockalingum, G.D., Belarbi, A. and Manfait, M. aminosulfonic acid pH buffers with a morpholinic ring for
(2001) FTIR spectroscopic analysis of Saccharomyces cerevi- cadmium and lead speciation studies by electroanalytical
siae cell walls: study of an anomalous strain exhibiting a techniques. Anal Chim Acta 394, 325–335.
pink-colored cell phenotype. FEMS Microbiol Lett 197, Soares, H.M.V.M., Pinho, S.C. and Barros, M.G.R.T.M.
179–186. (1999b) Influence of N-substituted aminosulfonic acids
Göksungur, Y., Üren, S. and Güvenç, U. (2005) Biosorption of with a morpholinic ring pH buffers on the redox processes
cadmium and lead ions by ethanol treated waste baker’s of copper or zinc ions: a contribution to speciation
yeast biomass. Biores Technol 96, 103–109. studies. Electroanalysis 11, 1312–1317.

ª 2009 The Authors


Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804 1803
Advantages of using dead biomass in heavy metals removal M.D. Machado et al.

Soares, E.V., Duarte, A.P.R.S. and Soares, H.M.V.M. (2000) Tsezos, M. (1990) Engineering Aspects of Metal Binding by Bio-
Study of the suitability of 2-(N-morpholino)ethanesulfonic mass. In Microbial Mineral Recovery ed. Ehrlich, H.L. and
acid pH buffer for heavy metals accumulation studies Brierly, C.L. pp. 325–339. USA: McGraw-Hill.
using Saccharomyces cerevisiae. Chem Spec Bioavailability Tsezos, M. (2001) Biosorption of metals. The experience accu-
12, 59–65. mulated and the outlook for technology development.
Soares, E.V., DeConick, G., Duarte, F. and Soares, H.M.V.M. Hydrometallurgy 59, 241–243.
(2002) Use of Saccharomyces cerevisiae for Cu2+ removal Verstrepen, K.J., Derdelinckx, G., Verachtert, H. and Delvaux,
from solution: the advantages of using a flocculent strain. F.R. (2003) Yeast flocculation: what brewers should know.
Biotechnol Lett 24, 663–666. Appl Microbiol Biotechnol 61, 197–205.
Soares, E.V., Hebbelinck, K. and Soares, H.M.V.M. (2003) Vieira, R.H.S.F. and Volesky, B. (2000) Biosorption: a solution
Toxic effects caused by heavy metals in the yeast Saccharo- to pollution? Int Microbiol 3, 17–24.
myces cerevisiae: a comparative study. Can J Microbiol 49, Volesky, B. (1990) Biosorption and biosorbents. In Biosorption
336–343. of Heavy Metals ed. Volesky, B. pp. 3–5. Boca Raton, FL:
Stoll, A. and Duncan, J.R. (1996) Enhanced heavy metal CRC Press.
removal from waste water by viable, glucose pretreated Volesky, B. (2001) Detoxification of metal-bearing effluents: bio-
Saccharomyces cerevisiae cells. Biotechnol Lett 18, 1209– sorption for the next century. Hydrometallurgy 59, 203–216.
1212. Volesky, B. (2003) Potential of biosorption. In Sorption and
Strandberg, G.W., Shumate, S.E. II and Parrot, J.R. Jr Biosorption, pp. 5–12. Montreal: BV Sorbex, Inc.
(1981) Microbial cells as biosorbents for heavy metals: Volesky, B. and Holan, Z.R. (1995) Biosorption of heavy met-
accumulation of uranium by Saccharomyces cerevisiae als. Biotechnol Prog 11, 235–250.
and Pseudomonas aeruginosa. Appl Environ Microbiol 41, Volesky, B. and May-Phillips, H.A. (1995) Biosorption of
237–245. heavy metals by Saccharomyces cerevisiae. Appl Microbiol
Strauss, C.J., Kock, J.L.F., Van Wyk, P.W.J., Lodolo, Biotechnol 42, 797–806.
E.J., Pohl, C.H. and Botes, P.J. (2005) Bioactive Volesky, B., May, H. and Holan, Z.R. (1993) Cadmium bio-
oxylipins in Saccharomyces cerevisiae. J Inst Brew 111, sorption by Saccharomyces cerevisiae. Biotechnol Bioeng 41,
304–308. 826–829.
Suh, J.H. and Kim, D.S. (2000) Effects of Hg2+ and cell condi- Wang, J. and Chen, C. (2006) Biosorption of heavy metals by
tions on Pb2+ accumulation by Saccharomyces cerevisiae. Saccharomyces cerevisiae: a review. Biotechnol Adv 24,
Bioprocess Eng 23, 327–329. 427–451.

ª 2009 The Authors


1804 Journal compilation ª 2009 The Society for Applied Microbiology, Journal of Applied Microbiology 106 (2009) 1792–1804

You might also like