pleurotus s00231

Upload: daniel-costas-imbernon

Post on 01-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/9/2019 Pleurotus s00231

    1/16

  • 8/9/2019 Pleurotus s00231

    2/16

    340   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    Fig. 1. Pollution of the environment by inorganic and organic compounds.

    2. Bioremediation and extra cellular enzymes

    Bioremediation, either as a spontaneous or as a managed

    strategy, is the application of biological processes for the

    clean up of hazardous chemicals present in the environment.

    The main bioremediation agents are plants, microor-

    ganisms and plant–microorganisms associations. All are

    effective agents in the transformation of organic pollutants

    because their enzymatic components are powerful cata-

    lysts, able to extensively modify structure and toxicological

    properties of contaminants or to completely mineralize the

    organic molecule into innocuous inorganic end products.

    Furthermore, enzymes carry out processes for which no

    efficient chemical transformations could have been devised.The degradative efficiency of biological processes, how-

    ever, depends on the biodegradability of the contaminants.

    It indicates the susceptibility of the contaminant to be de-

    graded into less toxic products, and is strongly influenced by

    the chemical structure, concentration and properties of the

    contaminant, and by environmental conditions. As claimed

    by Suthersan “A synthetic chemical that is not a product

    of biosynthesis will be degraded only if an enzyme or an

    enzyme system is able to catalyze the conversion of this

    compound to an intermediate or a substrate able to partici-

    pate in existing metabolic pathways” and also “The greater

    the difference in structure of the xenobiotic form from the

    compounds produced in nature, the less is the likelihood for

    significant biodegradation” [2].

    The most common contaminants can be classified on

    the basis of their biodegradability. Pollutants like simple

    hydrocarbons C1–C15, alcohols, phenols, amines, acids, es-

    ters, and amides are very easily biodegraded. By contrast,

    polychlorinated biphenyls (PCBs), polycyclic aromatic hy-

    drocarbons (PAHs) as well as pesticides are very difficult

    biodegradable. Usually the most complex is the chemical

    structure, the less biodegradable is the compound. Theo-

    retical predictions of compounds biodegradability can be

    usefully obtained by studies of molecular topology, i.e.

    studies of the molecular structure of pollutants, occurrence

    of branching, and types of atom-to-atom connections. The

    molecular connectivity, that can be determined by the struc-

    tural formula of the compound, can result of particular

    significance in molecular topology studies [1].

    In order to be biodegraded, contaminants must interact

    with enzymatic systems in the degrading organisms. If sol-uble, they can easily enter cells, if insoluble, they must be

    transformed into soluble or easily cell-available products.

    The first effective step for cell-transformation of insoluble

    substances, including xenobiotics and even plastic materials,

    is usually the reaction catalyzed by ecto- and extra cellular

    enzymes, which are deliberately released by the cells into

    their nearby environment. The process can be quite rapid

    for some natural compounds like cellulose or very slow for

    many xenobiotics.

    Extra cellular enzymes include a large range of oxidore-

    ductases and hydrolases. Both these enzymes may explicate

    a degradative function and transform polymeric substances

    into partially degraded or oxidized products that can be eas-ily up-taken by cells (Fig. 2). These latter in turn provide to

    their complete mineralization. For instance, partial oxidation

    of recalcitrant pollutants such as PAHs by extra cellular ox-

    idative enzymes give rise to products of increased polarity

    and water solubility and thus with a higher biodegradability

    [4].

    Oxidoreductases, however, may also explicate a protective

    function by oxidizing toxic soluble products into insoluble,

    not yet cell-accessible, products (Fig. 2).

    Tabatabai and Fu [5] demonstrated that several oxidore-

    ductases and hydrolases were extractable in a free form from

    soil, being the two classes of enzymes involved in the trans-formation of both xenobiotic molecules and natural prod-

    ucts. Later, Nannipieri et al.  [6]  reviewed more extensively

    the same topic. The authors underlined that enzyme-like

    activities rather than purified enzymatic proteins could be

    extracted from soil. The majority of these activities were

    exhibited by humic–enzyme complexes, showing properties

    often dissimilar from those of the free enzymes.

    Fig. 2. Roles of extra cellular enzymes in cell metabolism.

  • 8/9/2019 Pleurotus s00231

    3/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   341

    2.1. Plant root extra cellular enzymes

    Root-associated microorganisms often are assumed to be

    the only effectors of the increased xenobiotic biodegrada-

    tion occurring in the rhizosphere. An important contribution

    can instead derive from the involvement of degradative en-

    zymes released by plant roots in their surrounding environ-ment. These enzymes are usually wall-associated enzymes

    and provide to partially transform substances in products

    more easily up-taken by plant roots or rhizosphere microor-

    ganisms.

    Gramms et al.   [7]   followed the exudation of enzymes

    by the roots of 12 plant species in non-sterile soils for 56

    days. They demonstrated that several members of  Fabaceae,

    Gramineae and  Solanaceae efficiently and considerably re-

    leased both oxidases and hydrolases in the root regions of 

    the soil. Table 1 reports the amounts of peroxidase, laccase,

    monophenol monoxygenase, and proteinase–lipase–esterase

    activity, expressed as fluorescein diacetate hydrolase, re-

    leased by some of the investigated plants. The enzymes were

    of plant origin because the results were confirmed in ster-

    ile soils. Furthermore, SDS–PAGE and the isolectrofocus-

    ing of protein extracts from sterile and non-sterile cultures

    of alfalfa and from the unplanted sterile soil confirmed that

    the protein preparations from planted soils contained protein

    fractions with features characteristic of alfalfa peroxidase

    [7].

    In another experiment, Chroma et al. [8]  showed that, in

    vitro, some cultures of plants of various species and mor-

    phology transformed polychlorinated biphenyls as well as

    polycyclic aromatic hydrocarbons, and that their activity was

    ascribable to the production of a constitutive cell-wall asso-ciated peroxidase. The presence of PCBs, however, in other

    cases was toxic to both the plant and its peroxidase produc-

    tion.

    Extra cellular plant peroxidases may exhibit different

    functions depending on the pH values and the nature of the

    electron donors. Recently, Harvey et al.   [9]   have demon-

    strated that cultures of   Ipometa batada   produced a wide

    range of peroxidases with isoelectric points between 3 and

    Table 1

    Extractable enzyme activities in non-sterile soil–root regions of plants grown for 56 days

    Plant species Peroxidase (guaiacol,

    mM min−1)

    Laccase (ABTS,

    M min−1)

    Monophenol monoxygenase

    (DL-DPOA, M min−1)

    FDA hydrolase (fluorescein

    diacetate, M min−1)

    Control soil 0 0.14 0.48 0.034

    Fabaceae

    Alfaalfa 1984.7 48.6 33.1 0.30

    Soybean 11.1 0.033 0.80 0

    Gramineae

    Grass mixture 443.3 196.3 79.3 4.34

    Maize 2.6 0.074 0.61 0.004

    Solanaceae

    Tobacco 15.4 0.14 0.67 0.046

    Tomato 0 0017 0 0

    As modified from  [7].

    9. The enzymes were able to transform several hydroxy

    benzoic acids, and the enzyme activity depended on the

    pH and the nature of the electron donors. At pH 7.0, the

    salicylhydroxamic acid was oxidized in its quinonic form,

    whereas at pH 3.0 and in the presence of a secondary sub-

    strate, such as NADH, it served as redox mediator, and

    allowed the peroxidase-dependent oxidation of NADH.

    2.2. Microbial extra cellular enzymes

    2.2.1. Microbial oxidoreductases

    In vivo microbial oxidoreductases are periplasmic en-

    zymes associated with the cell surface of viable cells.

    Their main sources are fungi such as wood-degrading ba-

    sidiomycetes, terricolous basidiomycetes, ectomycorrizal

    fungi, soil-borne microfungi, and actinomycetes.

    A great interest is growing for the use of fungi as biore-

    mediating agents [10–13]. Most fungi are robust organisms

    and may tolerate higher concentrations of pollutants than

    bacteria. In particular, white-rot fungi appear unique and at-

    tractive organisms for the bioremediation of polluted sites

    for several reasons [10,14]. They can be summarized as fol-

    lows:

    (1) White-rot fungi are ubiquitous organisms in natural

    environments.

    (2) White-rot fungi are unique among eukaryotic or

    prokaryotic microorganisms, because they posses

    a very powerful extra cellular oxidative enzymatic

    system: the lignin-degrading enzyme system (LDS),

    which has broad substrate specificity and is able to ox-

    idize several environmental pollutants  [10,13]. Othernon-ligninolytic enzymes, like cellobiose dehydroge-

    nases, may participate in the transformation of pol-

    luting substances [11,12]. Cellobiose dehydrogenases

    are usually secreted under non-ligninolytic condi-

    tions when cellulose is the nutrient carbon, and either

    directly or indirectly they may oxide several contami-

    nants [11,12]. As a consequence, a vast range of toxic

    environmental pollutants, including low soluble com-

  • 8/9/2019 Pleurotus s00231

    4/16

  • 8/9/2019 Pleurotus s00231

    5/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   343

    thalene), that were not substrates of LiP action, suggested

    that other enzymes (i.e. monoxygenases) could be involved

    in the transformation of PAHs by the fungus  [33].

    2.2.2. Microbial hydrolases

    The other microbial enzymes involved in the pollutant

    transformation are hydrolases. Several bacteria and fungiproduce a group of extra or ecto cellular enzymes that in-

    clude proteases, carbohydratases (e.g. cellulases, amylases,

    xylanases, etc.), esterases, phosphatases and phytases. These

    enzymes are physiologically necessary to living organisms.

    Some of them (e.g. proteases and carbohydratases) catalyze

    the hydrolysis of large molecules, such as proteins and car-

    bohydrates, to smaller molecules for subsequent absorption

    by cells. Others, like phosphatases and phytases, contribute

    to the nutrition of plants and microbes by hydrolysis of or-

    ganic P compounds into inorganic P, the only form of phos-

    phorous available to plants and microbial cells.

    Due to their intrinsic low substrate specificity, hydrolases

    may play a pivotal role in the bioremediation of severalpollutants including insoluble wastes.

    A list of microbial hydrolases involved in the transfor-

    mation of natural and non-natural insoluble compounds,

    their preferential substrates and main sources is reported

    in   Table 3.   Keratinic wastes deriving from animal breed-

    ing, processing and handling, as well as used paper prod-

    ucts discarded by human population, may contribute largely

    to enrich the environment with solid wastes. Singh  [36]

    showed that the   C. keratinophilum   fungus, isolated by a

    waste site containing organic pollutants, was able to pro-

    duce an extra cellular protease with an optimum proteolytic

    activity against keratin substrates. As previously demon-strated by Kornillowicz-Kowalska [41], no clear correlation

    was, however, observed between the rate of keratin degra-

    dation and the enzyme production, thus indicating that other

    non-enzymatic factors were involved. The author concluded

    that the fungus and its proteolytic activity have a great po-

    tential for the in situ bioremediation of keratinous wastes.

    Table 3

    Biodegradation of natural and non-natural insoluble materials by microbial hydrolases

    Material Enzyme Source References

    Natural materialsCellulose materials Cellulase   Trichoderma resei,  Penicillium funiculosum   [34]

    Chitin Chitinase Actinobacteria   [35]

    Keratin Keratinase   Chysosporium keratinophilum   [36]

    Kraft pulp Xylanase,  -xylosidase   Sreptomyces thermoviolaceus   [37,38]

    Sewage sludge Protease, phosphatase Sulphate reducing bacteria   [39]

    Starch materials Amylase   Bacillus licheniformis   [40]

    Non-natural materials

    Nylon Nylon-degrading enzyme (MnP) Un-named white-rot fungus (+Mn and lactate)   [43,44]

    Poly(l-lactic acid) Depolymerase, alkaline protease   Amycolatopsis,  Bacillus  sp.   [45,46]

    Polyacrylate Cellobiose dehydrogenase White-rot fungi   [11]

    Polyurethane Esterase   Curvularia senegalensis (not wrf),

    Corynebacterium,   Comamonas acidovarans

    [47–49]

    Polyvinyl alcohol 2-4-Pentanedione esterase, laccase   Pseudomonas vesicularis,   Pycnoporus cinnabarinus   [50,51]

    An interesting use of cellulases was reported by van Wyk 

    [34], who demonstrated that cellulases from Penicillum funi-

    colosum and  Tricoderma resei   transformed paper materials

    of different origin (e.g. foolscap, filter, office, and newspa-

    per and microcrystalline cellulose (MCC)), and contributed

    to the treatment of solid municipal wastes. Both cellulases

    degraded all paper wastes though with a different efficiency.When the two fungal enzymes were tested in combination,

    the susceptibility of the cellulose substrates to their hy-

    drolytic degradation depended on the different ratio of the

    two cellulases in the mixture (Fig. 3).

    In a recent paper, Metcalfe et al. [35] investigated an ex-

    tra cellular group of bacterial hydrolases: the chitinases, that

    hydrolyze the    1–4 glycosidic bonds of chitin in soil. The

    authors reported the first molecular ecological assessment

    of chitinase diversity within a terrestrial environment. Their

    results confirmed that actinobacteria are important effective

    agents in chitin degradation in soil and that amendment ap-

    plication such as sludge application may contribute to affect

    the presence, the activity and diversity of these chitinases. Inparticular, the authors observed that an increase of enzyme

    activity but a decrease of chitinase diversity occurred upon

    sludge application.

    That a prominent role in the hydrolysis of phytin by fungi

    was explicated by an extra cellular phytase was demonstrated

    in comparative studies on the relative efficiency of intra and

    extra cellular phosphatases and phytases in six fungi  [42].

    The extra cellular phytases were found 60% more efficient

    than their intracellular counterparts; whereas most of the

    acid and alkaline phosphatase activity was found inside the

    cell.

    Pollutants of a great environmental concern are plastics,i.e. synthetic, man-made materials widely used in modern

    society for several purposes. Polyurethanes, polyacrylates,

    polylactides, nylon, starch polymers, and mixed composites

    of different starting materials are widely applied in the med-

    ical, automotive and industrial fields. These materials are

    highly recalcitrant and when discarded in the environment,

  • 8/9/2019 Pleurotus s00231

    6/16

  • 8/9/2019 Pleurotus s00231

    7/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   345

    waste treatment were previously reviewed by Nannipieri and

    Bollag [55], Karam and Nicell [56] and Nicell [57].

    As pointed out by the authors, cell-free enzymes can

    offer several advantages over the use of microbial cells.

    The most significant features of cell-free enzymes are their

    unique substrate-specificity and catalytic power; their capa-

    bility to act in the presence of many toxic, even recalcitrant,substances, and/or under a wide range of environmental

    conditions, often unfavourable to active microbial cells (i.e.

    relatively wide temperature, pH and salinity ranges, high

    and low concentrations of contaminants); and their low

    sensitivity or susceptibility to the presence of predators,

    inhibitors of microbial metabolism, and drastic changes in

    contaminant concentrations.

    Table 4 illustrates a wide range of cell-free enzymes ap-

    plied to the biodegradation of different pollutants.

    Pesticides of different chemical nature, very recalcitrant

    compounds like asphaltenes and PCBs, polychlorophe-

    nols, PAHs and others toxic pollutants were successfully

    transformed by oxidoreductases and hydrolases isolated byfungal, bacterial and plant cells. The majority of results

    summarized in Table 6, however, have been obtained under

    laboratory conditions. As it will be addressed below, several

    causes concur to hamper and render difficult the use of iso-

    lated enzymes as tools in the detoxifications of polluted sites.

    The three- and four-ring polycyclic aromatic hydrocar-

    bons anthracene, phenanthrene, pyrene and fluoranthene

    were in vitro-oxidized by extra cellular lignin peroxidase,

    manganese peroxidase and laccase, prepared from the white

    rot fungus   Nematoloma frowardii   and by mush-room ty-

    rosinase (Tyr) and horseradish peroxidase (HRP)  [58]. LiP

    transformed 58.6% of anthracene and 34.2% of pyrene,whereas 31.5% of anthracene and 11.2% of pyrene were ox-

    idized by MnP. In the presence of the mediating substances

    veratryl alcohol (for LiP), reduced glutathione (GSH) (for

    MnP), and ABTS (for L, Tyr, HRP), the transformation of 

    Table 4

    Biodegradation of pollutants by cell-free enzymes

    Pollutant Enzymes Source Properties References

    Anthracene, pyrene LiP, MnP, laccase   Nematoloma forwardii   High (veratryl alcohol)   [58]

    Asphaltenes Chloroperoxidase   Cladariomyces fumago   High activity in organic solvents   [59,60]

    Carbofuran, carbaryl Carbamate hydrolase   Achromobacter   sp.,

    Pseudomonas  sp.

    Cytosolic, high specificity   [61,62]

    Cyanides Cyanidase, cyanide hydratase   Alcaligenes denitrificans, several

    fungi

    Usually inducible, very stable in

    immobilized form

    [63–66]

    Estrogenic chemicals MnP, laccase   Trametes versicolor    70–100% transformation   [67]

    Nitrile compounds Nitrilase, nitrile hydratase,

    amidase

     Nocardia sp.,   Rhodococcus   sp.,

    Fusarium solani

    High stability up to 60 ◦C and

    pH 6.0–11.0

    [68]

    PCP, DDT, PCBs, lindane Dehalogenases, laccase Several microorganisms Sensitivity to SH-agents,

    stereospecificity

    [54,69]

    Phenols, PAH MnP, LiP, laccase,

    chloroperoxidases, peroxidase

    White-rot fungi Very versatile under different

    operational conditions

    [57,70–72]

    Pyrethroids, parathion,

    coumaphos, diazinon

     Agrobacterium,   Pseudomonas  sp.,

    Flavobacterium  sp.,  Nocardia  sp.,

     Bacillus cereus

    High stability at 50 ◦C and pH

    5.5–10.0

    [73–75]

    DDT: dichloro diphenyl tricholorethane. For other compounds and enzyme symbols see Table 2.

    Table 5

    Biodegradation of phenols by cell-free laccase from   Cerrena unicolor 

    Substrate –(OH) Substituent Substrate

    decrease (%)

    o-Cl-phenol 1   −1Cl 18

     p-Cl-phenol 1   −1Cl 20

    2,4-DCP 1   −2Cl 66

     p-Tyrosol 1 –CH2CH2OH 73

    o-Tyrosol 1 –CH2CH2OH 28

    m-Tyrosol 1 –CH2CH2OH 11

    Catechol 2 – 100

    Resorcinol 2 – 40

    Methylcatechol 2 –CH3   76

    Hydroxytyrosol 2 –CH2CH2OH 86

    Pyrogallol 3 – 100

    Gallic acid 3 –COOH 98

    From [70,71].

    PAHs was enhanced in most cases. Furthermore, studies

    performed with PAH-derivatives, known as intermediates or

    potential dead-end-products of microbial PAHs metabolism,

    demonstrated that the hydroxylated PAHs metabolites were

    oxidized by all the oxidoreductases, whereas PAH-quinones

    and oxo-metabolites were not transformed.

    The studies by Bollag and co-workers (for a detailed list

    see Table 2 and reported references in  [18]) and Gianfreda

    and co-workers [70,71,76–78]  have demonstrated that lac-

    cases from different fungal origins were capable of removing

    a large variety of phenols and the efficiency was strictly de-

    pendent on the chemical structure of the phenol, the type and

    the number of substituents on the aromatic ring (Table 5).

    An interesting application of fungal oxidases with es-trogenic chemicals was shown by Tsutsumi et al.   [67].

    Estrogenic chemicals are man-made chemicals that mimic

    the effects of hormones (particularly estrogens). Like nat-

    ural estrogens they can bind to the estrogen receptor and

  • 8/9/2019 Pleurotus s00231

    8/16

    346   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    Fig. 4. Bisphenol (BPA) and nonylphenol (NP) disappearance by oxidoreductases from various origins (from  [67]).

    regulate the activity of estrogen responsive genes. Such

    effects have raised concern that prolonged exposure to

    environmentally relevant concentrations of these chem-

    icals may adversely affect reproduction in wildlife and

    humans.

    Bisphenol A (2, 2-bis(4-hydroxyphenyl) propane; BPA)

    and nonylphenol (NP) are widely used in a variety of 

    industrial and residential applications, and are suspected

    of having estrogenic (endocrine-disrupting) activity. As

    shown in   Fig. 4   both BPA and NP disappeared within

    1 h-treatment with a MnP, isolated and partially purified

    from the strain   P. chrysosporium  ME-446. A laccase, par-

    tially purified from   T. versicolor   IFO7043, also removed

    BPA by 70% and NP by 60%, and the addition of the

    mediator 1-hydroxybenzentriazole (HBT) in the reaction

    system greatly improved the laccase potential. Analysis by

    gel permeation chromatography indicated oligomers as themain products of BPA and NP enzymatic transformation

    [67].

    The greatest concern for the biodegradation of estrogenic

    chemicals is aimed at the removal of their estrogenic ac-

    tivity. In vitro screening tests for chemicals with hormonal

    activities using yeasts [79] were utilized to evaluate the es-

    trogenic activity of BPA or NP after the enzymatic action

    [67].  Both Mn peroxidase and laccase removed the estro-

    Fig. 5. Decrease of bisphenol (BPA) and nonylphenol (NP) estrogenic activities by oxidoreductases from various origins (from   [67]).

    genic activities of BPA and NP within 12 h. Similar results

    were obtained within 6 h with laccase in the presence of HBT

    (Fig. 5).  The overall results led the authors to explain the

    BPA and NP transformation mechanisms as due to the poly-

    merization and partial degradation of the chemicals brought

    about by enzymatic oxidation [67].

    Another interesting application of cell-free enzymes is

    the use of the so-called nitrile-degrading enzymes capable

    of degrading nitrile compounds   [68]. Nitrile compounds

    are widespread in the environment. They comprise some

    of plant origins, such as cyanoglycosides, cyanolipids, rici-

    nine, phenylacetonitrile, etc. Nitrile compounds are also

    extensively used in chemical industries to produce a variety

    of polymers and other chemicals. Other different nitrile

    compounds are used as feedstock, solvents, extractants,

    pharmaceuticals, drug intermediates (chiral synthons), pes-

    ticides (dichlobenil, bromoxynil, ioxynil, buctril), or asintermediates in the organic synthesis of a variety of differ-

    ent compounds (e.g. amines, amides, amidines, carboxylic

    acids, esters, aldehydes, ketones and heterocyclics).Most ni-

    triles are however highly toxic, mutagenic and carcinogenic

    in nature (Pollak et al. [80]). If present in the environment at

    high concentrations they may cause severe diseases in hu-

    mans. Consequently, efforts have been made to control their

    release in the environment, and/or to remove them from it.

  • 8/9/2019 Pleurotus s00231

    9/16

  • 8/9/2019 Pleurotus s00231

    10/16

    348   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    Table 6

    Effects of co-substrates on bentazon transformation by oxidoreductases

    Humic monomers Laccaseb

    (pH 4.0)

    Peroxidasec

    (pH 3.0)

    Bentazona 0 6

    +Catechol 100 95

    +Ferulic acid 9 19

    +Guaiacol 10 9

    +Protocatechiuc acid 40 65

    +Pyrogallol 0 0

    +Resorcinol 2 0

    +Syringaldeyde 39 49

    +Vanillic acid 6 94

    +Caffeic acid 59 27

    Modified from  [86].a One mM for all the compounds.b From   Polyporus pinsitus. Incubation with 4 Units ml−1 for 24h at

    25 ◦C.c From horseradish. Incubation with 6 Units ml−1 for 24 h at 25 ◦C.

    references therein). The nature of polymeric products as wellas the possible mechanisms involved in the process was ad-

    dressed. These model system studies were aimed at achiev-

    ing a basic understanding of the possible mechanisms for

    oxidative coupling reactions occurring in natural soils and

    sediments [85].

    Increases and/or decreases of the pollutant transformation

    measured as both polymerization or mineralization of the

    target pollutant were detected. Kim et al.  [86] demonstrated

    that the transformation of bentazon, a very recalcitrant herbi-

    cide, by laccase or peroxidase may be completely annulled,

    or greatly enhanced, by the co-presence of a variety of sub-

    stances that are precursors of humic materials (Table 6).In our studies we have also demonstrated that for exam-

    ple the 2,4-dichlorophenol (2,4-DCP) decrease by laccases

    from different origins (i.e.  Cerrena unicolor ,  Trametes vil-

    losa) may be differently affected by the co-presence of 

    one or two other chlorophenols such as 4-chlorophenol or

    2,4,6 thriclorophenol (Table 7)  [87]. The 2,4-DCP-laccase

    mediated decrease can also be affected by the presence of 

    its parent precursor, 2,4D, another phenol such as catechol,

    or simazine, that is usually supplied to soil in combination

    with 2,4D (Table 7)   [71].   In another study, enhancing or

    depressing effects of a laccase from   Rhus vernicifera   on

    catechol transformation were measured in a model system

    of four phenols (catechol, methylcatechol, tyrosol and hy-

    droxytytrosol) [78]. Binary, ternary, and quaternary aqueous

    phenolic mixtures were investigated (Table 7).  The model

    system simulates a typical wastewater deriving from an

    olive oil factory. In Mediterranean countries, large quan-

    tities of wastewater with a high content of phenolic sub-

    stances are usually produced as characteristic by-products

    of olive-oil production. The main constituents are usually

    the compounds used in the model system.

    Differentiated effects on the enzyme activity were also ob-

    served by the presence of a polluting substance non-substrate

    of the enzyme. Investigations performed under laboratory

    Table 7

    Substrate removal and residual enzymatic activity after laccase action

    Substrate Substrate

    decrease (%)∗Laccase

    activity (%)

    Single phenols   Cerrena unicolor 

    2,4-DCP 66 34

    Tyrosol 11 88

    Resorcinol 40 76Methylcatechol 76 24

    Hydroxytyrosol 86 18

    Pyrogallol 100 89

    Gallic acid 98 83

    Phenol mixtures   Trametes villosa

    2,4-DCP 66 34

    +4-CP 56 56

    +2,4,6-TCP 58 58

    +2,4-D 82 20

    Cerrena unicolor 

    +Catechol 77 20

    +Simazine 46 30

    +4-CP  + 2,4,6-TCP 50 35

    +2,4-D  + Simazine 0 95+Catechol  + Simazine 39 60

    +2,4-D  + Catechol 100 0

     Rhus vernicifera

    Catechol 58 70

    +Methylcatechol (M) 38 11

    +Tyrosol (T) 100 66

    +Hydroxytyrosol (H) 99 27

    +M  + T 16 9

    +T  + H 95 29

    +M  + H 56 10

    +M  + T  + H 63 11

    The substrate decrease values have been normalized by laccase units.

    From [70,71,78,87].

    conditions with acid phosphatase from sweet potato a typ-

    ical, extra or ecto cellular enzyme, and pesticides such as

    carbaryl, atrazine, and glyphosate demonstrated that the ac-

    tivity of the enzyme was depressed by atrazine and carbaryl

    (40 and 34% of activity reduction, respectively), whereas no

    effects were measured with glyphosate [88].

    A key factor possibly affecting the pollutants transfor-

    mation is the availability of the polluting substance to the

    detoxifying agent, that is its bioavailability  [89].   Bioavail-

    ability is affected by both the inherent properties of the

    pollutants (i.e. concentration, molecular structure, water or

    organic solubility) and the characteristics of the environ-

    ment where they occur.

    In soil, several properties such as the amount of or-

    ganic matter, the thickness of silt and clay beds, the pres-

    ence of dissolved organic matter, the soil aggregation and

    sub-superficial heterogeneity can all influence the bioavail-

    ability of pollutants [90].

    Adsorption and/or entrapment of pollutants on/into or-

    ganic and inorganic soil colloids and in soil matrix further

    complicate the whole system, and can limit the amount of 

    pollutant available to microbial cells and their extra cel-

    lular enzymes. Their limiting effect can be greater toward

  • 8/9/2019 Pleurotus s00231

    11/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   349

    isolated enzymes. Aging processes can also lead to a sig-

    nificant reduction of pollutants bioavailability, because of 

    their sequestration from diffusion into sites within sorbing

    matrices or entry into nanopores, both not easily accessible

    to large molecules as enzymatic proteins [91].

    4.2. Enzyme-derived disadvantages

    Disadvantages to the in situ application of extra cellular,

    cell-associated or cell-free enzymes may also arise from the

    enzymes.

    Enzymes may lose their activity upon pollutant transfor-

    mation. In the experiments on phenols previously shown in

    Tables 5 and 7, the residual activity of laccase was measured

    after phenol transformation under standard conditions  [71].

    The results indicated that the enzyme partially, or totally,

    lost its catalytic activity (Table 7). The reduction of activ-

    ity was quite dependent on the type of phenol, the entity

    of its transformation, and the combination of the phenolsin the mixture. Usually, the higher the phenol concentra-

    tion, the more pronounced was the loss of laccase activity.

    Measurements of residual laccase activities made in experi-

    ments performed with increasing concentrations (from 0 to

    0.4 mM) of 2,4-DCP ruled out a possible inhibitory effect

    on laccase activity by 2,4-DCP. Indeed, activity tests per-

    formed on samples filtered with filters that specifically ad-

    sorb 2,4-DCP gave activity values similar to those obtained

    without sample filtration. The authors explained their results

    as due to the a progressive entrapment and/or adsorption of 

    active enzyme molecules within/on phenol polymeric aggre-

    gates as they formed [70]. These processes may represent

    the main mechanism of inactivation because they hinder the

    interaction between the substrate and the enzyme  [92].

    In the soil, enzymes are present in complex, three-

    dimensional assemblages of mineral and organic particles

    that will restrict their mobility and affect their activity

    (Fig. 6) [6,93–95]. Enzyme molecules can be adsorbed, im-

    mobilized, or entrapped in such matrices giving rise to the

    so-called “naturally immobilized enzymes”   [94]. Inactiva-

    tion or degradation of enzymatic molecules may also take

    Fig. 6. Soil bound enzymes.

    Fig. 7. Removal of 2,4-DCP by laccase in the presence of clays, clay-humic

    complexes, sand and soils. M: Na-montmorillonite; AM: montmorillonite

    covered by different amounts of OH aluminum species (AM3  and AM18,

    loading 3 and 18 mmol of Al per gram of montmorillonite); HA: a humic

    acid.

    place. As a consequence, changes in their kinetics, stabil-

    ity and mobility will occur  [96].  These latter, in turn, will

    determine the operating range of enzymes in soil around

    microorganisms and enzymes.

    Investigations were performed to evaluate the effective-

    ness of laccases from different origins towards the 2,4-DCP

    removal when in the presence of different soil components

    [70,71,76,77,97]. A montmorillonite (M), a montmorillonite

    covered by different amounts of OH aluminum species

    (AM3   and AM18, loading 3 and 18 mmol of Al per gram

    of montmorillonite), a humic acid (HA), a combination of 

    HA and AM18, sand and soils with different organic mattercontents were utilized. As respect to the control, i.e. the

    enzymes alone, soil components differently affected the

    2,4-DCP-laccase activity (Fig. 7). Strong depressing effects

    were measured in the presence of montmorillonite, AM18and the combination AM18   +HA. When soils were used,

    the reduction of laccase activity increased by increasing the

    organic matter content of soils [97].

    The immobilization of enzymes on soil components

    affected the response of the enzyme to the presence of sub-

    stances extraneous to the reaction substrates, as well. When

    the influence of carbaryl, atrazine, and glyphosate was

    tested on acid phosphatase immobilized on organic (tan-

    nic acid), inorganic (montmorillonite) and organo-mineral

    (Al(OH) x-tannic acid-montmorillonite ) complexes a dif-

    ferent behavior as respect to the free enzyme was observed

    [88].   Different causes were invoked by the authors to ex-

    plain their results. They related to: (i) the “state” of the

    enzyme if free or immobilized on soil components, that

    makes the whole system homogenous or heterogeneous; (ii)

    the nature of interactions occurring between the enzyme and

    the immobilizing support; (iii) the conformational changes

    achieved by the enzymatic protein upon immobilization;

    and (iv) the influence of the microenvironment created by

    the support in the surrounding of the protein  [88,95].

  • 8/9/2019 Pleurotus s00231

    12/16

    350   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    As regards to isolated enzymes, a drawback that greatly

    hampers their practical application is the cost of enzymes

    isolation and purification. Given the best producer of the se-

    lected enzyme, the production of a purified enzyme requires

    long and expensive isolation and purification procedures.

    Furthermore, very low amounts of the purified protein

    are usually obtained thus rendering the whole process toocostly for practical applications. Alternative inexpensive

    technologies, using for instance agricultural wastes as the

    carbon source, may be adopted to increase the growth of 

    the enzyme-producing microorganisms and to significantly

    reduce the cost of enzyme production   [98]. Even if the

    enzymes are available at very low costs, their use in soil

    as free proteins may be hindered by their short life in an

    inhospitable environment such soil [54,55,94].

    An interesting, low-cost alternative to the use of pu-

    rified enzymes can be the utilization of plant material

    loading enzymatic activities and proved to be effective

    in the detoxification of polluting substances. As it is the

    case of the results obtained in the investigations performedwith minced horseradish roots  [99,100]   and olive husk, a

    by-product of olive oil production showing phenoloxidase

    activity  [101,102],   used for the successful transformation

    of phenolic pollutants.

    Improvements in the production of isolated enzymes may

    also derive from molecular technologies. Production of an-

    imal and plant enzymes could be performed by means of 

    genetically modified microorganisms or plants [103,104].

    The range of sophisticated modern molecular technolo-

    gies now available has provided the researcher with an im-

    mensely enhanced choice of potential enzyme sources. For

    example, an increasing knowledge of several aspects of theenzymology and molecular biology of the powerful extra

    cellular oxidative system (LDS) of fungi is now available

    [10,13]. The major genes encoding LiPs, MnPs, and laccases

    were cloned and sequenced. The regulation of expression

    of the key ligninolytic genes were studied, and the X-ray

    crystallographic structures of the LiP and MnP isozymes are

    now known and available. These advancements should lead

    to the successful genetic engineering of white-rot fungi and

    their enzyme systems. Most advantageous bioremediation

    strategies for treating contaminated sites can thus possibly

    be planned and applied.

    The engineering of microbes and enzymes may also help

    to improve bioremediation, particularly when the applica-

    tion on a large-scale of the selected biological catalyst is

    hampered by the low rate of pollutant degradation   [105].

    As claimed by Chen and Mulchandani [106], recent advan-

    tages in genetic and molecular techniques have offered new

    chances to modify microbes or enzymes to function as “de-

    signer biocatalysts” in which certain required qualities from

    different organisms are brought together in a single catalyst

    to perform specific bioremediation.

    An example is that regarding the transformation of 

    organophosphates by the so-called “live biocatalysts” [106].

    Organophosphates, like parathion, paraoxon, diazinon,

    coumaphos, and malathion, are among the most toxic pesti-

    cides used in agriculture. They are easily degraded by two

    bacterial hydrolases, organophosphorus hydrolase (OPH)

    and organophosphorus acid anydrase (OPA), giving rise to

    less toxic compounds and further degradable by biological

    and chemical processes.

    Organophosphate hydrolase is an enzyme produced andisolated by several soil microorganisms, including many

    Pseudomonas   strains. It is very effective in degrading

    these compounds but its practical application is hindered

    by the high cost of its purification. By contrast, the use

    of whole-cells producing the enzyme can be more cost

    effective, but their use is restricted by transport barrier of 

    organophosphates across the cell membrane [107,108].

    Parathion and paraoxon were efficiently detoxified with

    OPH anchored and displayed on the cell surface of   Es-

    cherichia coli. The recombinant whole-cells with surface-

    expressed OPH were obtained using the same system used

    for the expression of  -lactamase on E. coli [108]. The sys-

    tem is made by the fusion of the N-terminal lipoprotein se-quence (Lpp), the domain of the pore-forming protein OmpA

    and the enzyme. The Lpp directs the Lpp–OmpA–OPH fu-

    sion to the outer membrane and the OmpA domain is able

    to pass through the lipid by-layer and to localize the enzyme

    on the cell surface (Fig. 8).   These “live biocatalysts” dis-

    played about seven times higher activity than E. coli whole-

    cells expressing similar amounts of intracellular OPH, thus

    demonstrating the higher efficiency of the enzyme when

    acting outside the cell. Furthermore, the enzyme showed a

    higher stability to both temperature (100% of activity upon

    30 day-incubation at 37 ◦C) and organic solvents.

    Enzymes immobilized on natural and synthetic supportsof different nature and through different immobilization

    mechanisms have been often proposed as efficient catalytic

    tools to overcome several disadvantages linked to the use of 

    free enzymes   [54,55]. Immobilized enzymes have usually

    a long-term and operational stability, being very stable to-

    ward physical, chemical, and biological denaturing agents.

    Furthermore, they may be reused and recovered at the end

    of the process. While the use of immobilized enzymes

    is widely diffused in several applicative fields, including

    environmental applications   [55–57,109],   their large-scale

    application in the bioremediation of polluted soils is not

    reported, at least to our knowledge.

    The potential of enzymes for bioremediation purposes can

    greatly increase by the use of microorganisms and their en-

    zymes from extreme environments. Enzymes from both ther-

    mophilic and psycrophilic microorganisms usually display

    some unusual and particular features that may render them

    potential, powerful catalysts for the degradation of pollut-

    ing chemicals ([110] and references therein,  [111]).   Ther-

    mophiles and psycrophiles have to adapt themselves to live

    and survive under extreme environmental conditions  [112].

    As a consequence, specialized enzymatic proteins, particu-

    larly stable to different denaturing agents, and with elevated

    catalytic activity, are produced.

  • 8/9/2019 Pleurotus s00231

    13/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   351

    Fig. 8. Parathion detoxification by organophosphate hydrolase (OPH) anchored and displayed on the cell surface of  E. coli  (from [106]). Lpp: lipoprotein;OmpA: pore forming protein.

    The molecular, structural, kinetic and genetic properties of 

    such enzymes have been in most cases elucidated. Cloning

    and expression of genomic related information in heterol-

    ogous, fast-growing meshophiles, like  E. coli, has allowed

    an increased, much less cheaper commercial production of 

    thermophilic enzymes [113–115]. An increasing knowledge

    of cold-adapted enzymes properties is now being develop-

    ing, as well [110].

    Many industrial applications have benefited from the useof these enzymes. More challenges and potential advantages

    may also be envisaged for their application in biotechnol-

    ogy including bioremediation ([110] and references therein,

    [111]).

    5. Conclusions

    In conclusion, several extra cellular enzymes, either as

    cell-associated or cell-free enzymes, may behave as powerful

    catalysts in the biodegradation of harmful pollutants. How-

    ever, their large-scale application for remediation of polluted

    soils is still limited. This may derive from several drawbacks

    and disadvantages depending on both the pollutants and the

    enzymes. For instance, the simultaneous presence of several

    polluting substances in a contaminated site with synergistic,

    often negative, effects on the enzyme efficiency, the high

    costs associated with the isolation and purification of free

    enzymes, the low stability of enzymes to the harsh condi-

    tions of soil all concur to restrict the wide use of enzymes as

    remediating agents of polluted soils. Although immobilized

    enzymes may present a high stability under soil conditions,

    they are not widely applied in the remediation of polluted

    soils.

    However, exploration of extreme environments, exploita-

    tion of genome using advanced technologies, and protein

    engineering have opened new frontiers for the production

    and application of enzymes. As mentioned by Burton  [84]

    these achievements has lead to the point where rather than

    develop a purpose for an enzyme, it is possible to design

    and develop an enzyme for a purpose. This fascinating per-

    spective has to resolve a still opened question: what should

    be the features of an enzyme to be suited for remediation of soils?

    Finally, a successful management or remediation of a con-

    taminated site should make the site environmentally agree-

    able and usable for some acceptable purposes. This means to

    adopt appropriate cleanup criteria and determine what con-

    stitutes an environmental acceptable endpoint.

    Acknowledgments

    This research was partly supported by European Commu-

    nity, Project INCO-MED Contract no. ICA3-2002-10021.The authors are grateful to Dr. Luigi Pagano for his help

    in the preparation of figures and to Dr. Anna Maria Woods

    for her technical help in editing the paper in respect to the

    English language and style. DiSSPA Contribution no. 0062.

    References

    [1] Alexander M. Biodegradation and bioremediation. San Diego:

    Academic Press; 1999.

    [2] Suthersan SS. In situ bioremediation. Remediation engineering:

    design concepts. Boca Raton: CRC Press; 1999.

  • 8/9/2019 Pleurotus s00231

    14/16

    352   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    [3] Iwamoto T, Nasu M. Current bioremediation practice and

    perspective. J Biosci Bioeng 2001;92:1–8.

    [4] Meulenberg R, Rijnaarts HHM, Doddema HJ, Field JA.

    Partially oxidized polycyclic aromatic hydrocarbons show an

    increased bioavailability and biodegradability. FEMS Microbiol Lett

    1997;152:45–9.

    [5] Tabatabai MA, Fu M. Extraction of enzymes in soil. In: Stotzky G,

    Bollag J-M, editors. Soil biochemistry, vol. 7. New York: MarcelDekker; 1992. p. 197–227.

    [6] Nannipieri P, Sequi P, Fusi P. Humus and enzyme activity. In:

    Piccolo A, editor. Humic substances in terrestrial ecosystems.

    London: Elsevier Science; 1996. p. 3–328.

    [7] Gramms G, Voigt KD, Kirsche B. Oxidoreductase enzymes liberated

    by plant roots and their effects on soil humic material. Chemosphere

    1999;38:1481–94.

    [8] Chroma L, Mackova M, Kucerova P, Der Wiesche C, Burkhard

    J, Macek T. Enzymes in plant metabolism of PCBs and Ps. Acta

    Biotechnol 2002;22:35–41.

    [9] Harvey PJ, Xiang M, Palmer JM. Extra cellular enzymes in

    the rhizosphere. In: Proceedings of the Inter-Cost Workshop on

    Soil—Microbe—Root Interactions: Maximising Phytoremediation/ 

    Bioremediation, Grainau, Germany, 23—25 May 2002.

    [10] Reddy CA. The potential for white-rot fungi in the treatment of pollutants. Curr Opin Biotechnol 1995;6:320–8.

    [11] Cameron MD, Aust SD. Degradation of chemicals by reactive

    radicals produced by cellobiose dehydrogenase from  Phanerochaete

    chrysosporium. Arch Biochem Biophys 1999;367:115–21.

    [12] Cameron MD, Timofeevski S, Aust SD. Enzymology of 

    Phanerochaete chrysosporium   with respect to the degradation of 

    recalcitrant compounds and xenobiotics. Appl Microbiol Biotechnol

    2000;54:751–8.

    [13] Pointing SB. Feasibility of bioremediation by white-rot fungi. Appl

    Microbiol Biotechnol 2001;57:20–32.

    [14] Bumpus JA. White-rot fungi and their potential use in soil

    bioremediation processes. In: Bollag JM, Stotzky G, editors. Soil

    biochemistry. New York: Marcel Dekker; 1993. p. 65—100.

    [15] Reddy CA, D’Souza TM. Physiology and molecular biology of 

    the lignin peroxidases of   Phanerochaete chrysosporium. FEMS

    Microbiol Rev 1994;13:137–52.

    [16] Higuchi T. Biodegradation mechanism of lignin by white-rot

    basidiomycetes. Biotechnology 1993;30:1–8.

    [17] Wariishi H, Valli K, Gold MH. Maganese(II) oxidation by

    manganese peroxidase from the basidiomycete   Phanerochaete

    chrysosporium. J Biol Chem 1992;267:23688–95.

    [18] Gianfreda L, Xu F, Bollag J-M. Laccases: a useful group of 

    oxidoreductive enzymes. Bioremediat J 1999;3:1–25.

    [19] Bourbonnais R, Paice MG, Freiermuth B, Bodie E, Borneman S.

    Reactivities of various mediators and laccases with kraft pulp and

    lignin model compounds. Appl Environ Microbiol 1997;63:4627–

    32.

    [20] Pointing SB, Jones EBG, Vrijmoed LLP. Optimization of laccase

    production by   Pycnoporus sanguineus   in submerged liquid culture.Mycologia 2000;92:139–44.

    [21] Pointing SB, Vrijmoed LLP. Decolorization of azo and tri-

    phenylmethane dyes by   Pycnoporus sanguineus  producing laccase

    as the sole phenoloxidase. World J Microbiol Biotechnol

    2000;16:317–8.

    [22] Archibald FS, Paice MG, Jurasek L. Decolorization of kraft

    bleachery effluent chromophores by  Coriolus (Trametes) versicolor .

    Enzyme Microb Technol 1990;12:846–53.

    [23] Limura Y, Hartikainen P, Tatsumi K. Dechlorination of 

    tetrachloroguaiacol by laccase of white-rot basidiomycete  Coriolus

    versicolor . Appl Microbiol Biotechnol 1996;45:434–9.

    [24] Bumpus JA. Biodegradation of polycyclic aromatic hydrocarbons

    by Phanerochaete chrysosporium. Appl Environ Microbiol 1989;55:

    154–8.

    [25] Bogan BW, Lamar RT. Polycyclic aromatic hydrocarbon-degrading

    capabilities of  Phanerochaete laevis HHB-1625 and its extra cellular

    ligninolytic enzymes. Appl Environ Microbiol 1996;62:1597–603.[26] Zeddel A, Majcherczyk A, Hutterman A. Degradation of 

    polychlorinated biphenyls by white-rot fungi   Pleurotus ostreatus

    and Trametes versicolor   in a solid state system. Toxicol Environ

    Chem 1993;40:225–66.[27] Novotny C, Vyas BRM, Erbanova P, Kubatova A, Sasek V. Removal

    of various PCBs by various white-rot fungi in liquid cultures. Folia

    Microbiol 1997;42:136–40.[28] Alleman BC, Loan BE, Gilbertson RL. Degradation of 

    pentachlorophenol by fixed films of white-rot-fungi in rotating tube

    bioreactors. Water Res 1995;29:61–7.[29] Lin JE, Wang HY, Hickey RF. Degradation kinetics of 

    pentachlorophenol by   Phanerochaete chrysosporium. Biotechnol

    Bioeng 1990;35:1125–34.[30] Bumpus JA, Tatarko M. Biodegradation of 2,4,6-trinitrotoluene by

    Phanerochaete chrysosporium: identification of initial degradation

    products and the discovery of a TNT-metabolite that inhibits lignin

    peroxidase. Curr Microbiol 1994;28:185–90.[31] Donnelly KC, Chen JC, Huebner HJ, Brown KW, Autenrieth

    RL, Bonner JS. Utility of four strains of white-rot fungi for the

    detoxification of 2,4,6-trinitrotoluene in liquid culture. Environ

    Toxicol Chem 1997;16:1105–10.[32] Scheibner K, Hofrichter M. Conversion of aminonitrotoluenes by

    fungal manganese peroxidase. J Basic Microbiol 1998;38:51–9.[33] Sutherland JB, Selby AL, Freeman JP, Evans FE, Cerniglia CE.

    Metabolism of phenanthrene by   Phanerochaete chrysosporium.

    Appl Environ Microbiol 1991;57:3310–6.[34] van Wyk JPH. Saccharification of paper products by cellulase from

    Penicillium funiculosum   and   Trichoderma reesei. Biomass Bioeng

    1999;16:239–42.[35] Metcalfe AC, Krsek M, Gooday GW, Prosser JI, Wellington EMH.

    Molecular analysis of a bacterial chitinolytic community in an

    upland pasture. Appl Environ Microbiol 2002;68:5042–50.[36] Singh CJ. Optimization of an extra cellular protease of 

    Chrysosporium keratinophilum and   its potential in bioremediation

    of keratinic wastes. Mycopathologia 2002;156:151–6.

    [37] Garg AP, McGarthy AJ, Roberts JC. Biobleaching effect forStreptomyces thermoviolaceus  xylanase preparations on birchwood

    kraft pulp. Enzyme Microbial Biotechnol 1996;18:261–7.[38] Christov LP, Prior BA. Repeated treatments with   Aureobasidium

     pullulans hemicellulases and alkali enhance biobleaching of sulphite

    pulps. Enzyme Microb Biotechnol 1996;18:244–50.[39] Whiteley CG, Heron P, Pletschke B, Rose PD, Tshivhunge S,

    Van Jaarsveld FP, et al. The enzymology of sludge solubilisation

    utilising sulphate reducing systems. Properties of proteases and

    phosphatases. Enzyme Microbial Technol 2002;31:419–24.[40] Copinet A, Bliard C, Onteniente JP, Couturier Y. Enzymatic

    degradation and deacetylation of native and acetylated strach-based

    extruded blends. Polym Degrad Stabil 2001;71:203–12.[41] Kornillowicz-Kowalska T. Studies on the decomposition of keratin

    wastes by saprotrophic microfungi. I. Criteria for evaluating

    keratinolytic activity. Acta Mycol 1997;32:51–79.[42] Tarafdar C, Yadav RS, Niwas R. Relative efficiency of fungal intra-

    and extracellular phosphatases and phytase. Plant Nutr Soil Sci

    2002;165:17–9.[43] Deguchi T, Kakezawa M, Nishida T. Nylon degradation by lignin-

    degrading fungi. Appl Environ Microbiol 1997;63:329–31.[44] Deguchi T, Kitaoka Y, Kakezawa M, Nishida T. Purification

    and characterization of a nylon-degrading enzyme. Appl Environ

    Microbiol 1998;64:1366–71.[45] Nakamura K, Tomita T, Abe N, Kamio Y. Purification and

    characterization of an extra cellular poly(l-lactic acid) depolymerase

    from a soil isolate. Appl Environ Microbiol 2001;67:345–53.[46] Oda Y, Yonetsu A, Utakami T, Tonomura K. Degradation of 

    polylactide by commercial proteases. J Polym Environ 2000;8:29–

    32.

  • 8/9/2019 Pleurotus s00231

    15/16

     L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354   353

    [47] Crabbe JR, Campbell JR, Thompson L, Walz SL, Schultz WW.

    Biodegradation of a colloidal ester-based polyurethane by soil fungi.

    Int Biodeter Biodegr 1994;33:103–13.

    [48] Kay MJ, McCabe RW, Morton LHJ. Chemical and physical changes

    occurring in polyester polyurethane during biodegradation. Inter

    Biodet Biodegrad 1993;31:209–25.

    [49] Nakajima-Kambe T, Shigeno-Akutsu Y, Nomura N, Onuma F,

    Nakahara T. Microbial degradation of polyurethane, polyester

    polyurethanes and polyether polyurethanes. Appl Microbiol

    Biotechnol 1999;51:134–40.

    [50] Kawagoshi Y, Fujita M. Purification and properties of the polyvinyl

    alcohol-degrading enzyme 24-pentanedione hydrolase obtained from

    Pseudomonas vesicularis   var.   povalolyticus. World J Microbiol

    Biotechnol 1998;14:95–100.

    [51] Larking DM, Crawford RL, Christie GBY, Lonergan GT. Enhanced

    degradation of polyvinyl alcohol by   Pycnoporus cinnabarinus

    after pretreament with Fenton’s reagent. Appl Environ Microbiol

    1999;65:1798–800.

    [52] Howard GT. Biodegradation of polyurethane: a review. Int Biodeter

    Biodegrad 2002;49:245–52.

    [53] Langsford ML, Gilkes NR, Sing S, Moser B, Miller Jr RC, Warren

    RAJ, et al. Glycosylation of bacterial cellulases prevents proteolytic

    cleavage between functional domains. FEBS Lett 1987;225:163–7.[54] Gianfreda L, Bollag J-M. Isolated enzymes for the transformation

    and detoxification of organic pollutants. In: Burns RG, Dick 

    R, editors. Enzymes in the environment: activity, ecology and

    applications. New York: Marcel Dekker; 2002. p. 491—538.

    [55] Nannipieri P, Bollag J-M. Use of enzymes to detoxify pesticide-

    contaminated soils and waters. J Environ Qual 1991;20:510–7.

    [56] Karam J, Nicell JA. Potential application of enzymes in waste

    treatment. J Chem Technol Biotechnol 1997;69:141–53.

    [57] Nicell JA. Environmental applications of enzymes. Interdisci

    Environ Rev 2001;3:14–41.

    [58] Guenther T, Sack U, Hofrichter M, Laetz M. Oxidation of PAH

    and PAH-derivatives by fungal and plant oxidoreductases. J Basic

    Microbiol 1998;38:113–22.

    [59] Fedorak PM, Semple KM, Vazquez-Duhalt R, Westlake DWS.

    Chloroperoxidase-mediated modifications of petroporphyrins andasphaltenes. Enzyme Microb Technol 1993;15:429–37.

    [60] Mogollon L, Rodriguez R, Larrota W, Ortiz C, Torres R.

    Biocatalytic removal of nickel and vanadium from petroporphyrins

    and asphaltenes. Appl Biochem Biotechnol 1998;70-72:765–77.

    [61] Derbyshire MK, Karns JS, Kearney PC, Nelson JO. Purification

    and characterization of an  N -methylcarbamate pesticide hydrolyzing

    enzyme. J Agric Food Chem 1987;35:871–7.

    [62] Mulbry WW, Eaton RW. Purification and characterization of the

     N -methylcarbamate hydrolase from   Pseudomonas   strain CRL-OK.

    Appl Environ Microbiol 1991;57:3679–82.

    [63] Basheer S, Kut OM, Prenosil JE, Bourne JR. Kinetics of enzymatic

    degradation of cyanide. Biotechnol Bioeng 1992;41:465–73.

    [64] Basheer S, Kut OM, Prenosil JE, Bourne JR. Development

    of an enzyme membrane reactor for treatment of cyanide-

    containing wastewaters from the food industry. Biotechnol Bioeng1993;39:629–35.

    [65] Barclay M, Hart A, Knowles CJ, Meenssen JCL, Tett VA.

    Biodegradation of metal cyanides by mixed and pure cultures of 

    fungi. Enzyme Microb Tehcnol 1998;22:22.

    [66] Lynch JM. Resilience of the rhizosphere to anthropogenic

    disturbance. Biodegradation 2002;13:21–7.

    [67] Tsutsumi Y, Haneda T, Nishida T. Removal of estrogenic activities

    of bisphenol A and nonylphenol by oxidative enzymes from lignin-

    degrading basidiomycetes. Chemosphere 2001;42:271–6.

    [68] Banerjee A, Sharma R, Banerjee UC. The nitrile-degrading

    enzymes: current status and future prospects. Appl Microbiol

    Biotechnol 2002;60:33–44.

    [69] Beaudette LA, Ward OP, Pickard MA, Fedorak PM. Low surfactant

    concentration increases fungal mineralization of a polychlorinated

    biphenyl congener but has no effect on overall metabolism. Lett

    Appl Microbiol 2000;30:155–60.[70] Gianfreda L, Sannino F, Filazzola MT, Leonowicz A. Catalytic

    behavior and detoxifying ability of a laccase from the fungal strain

    Cerrena unicolor . J Mol Catal B: Enzymatic 1998;14:13–23.[71] Filazzola MT, Sannino F, Rao MA, Gianfreda L. Effect of various

    pollutants and soil-like constituents on laccase from   Cerrena

    unicolor . J Environ Qual 1999;28:1929–38.[72] Bollag J-M, Shuttleworth KL, Anderson DH. Laccase-mediated

    detoxification of phenolic compounds. Appl Environ Microbiol

    1988;54:3086–91.[73] Coppella SJ, Cruz ND, Payne GF, Pogell BM, Speedie MK, Karns

    JS, et al. Genetic engineering approach to toxic waste management

    case study for organophosphate waste treatment. Biotechnol Progr

    1990;6:76–81.[74] Mulbry WW. Selective deletions involving the organophosphorus

    hydrolase gene adpB from   Nocardia   strain B-1. Microbiol Res

    1998;153:213–7.[75] Sutherland T, Russel R, Selleck M. Using enzymes to clean pesticide

    residues. Pest Outlook 2002;13:149–51.[76] Filazzola MT, Sannino F, Gianfreda L. The influence of biological

    agents on the behaviour of pollutants in soil. Fresinius Environ Bull

    1998;7:484–9.

    [77] Sannino F, Filazzola MT, Violante A, Gianfreda L. Fate of herbicides as influenced by abiotic and biotic mechanisms.

    Chemosphere 1999;39:333–41.[78] Rao MA, Sannino F, Gianfreda L, Bollag J-M. Oxidative

    transformation of phenols in aqueous mixtures. Water Res

    2003;37:3205–15.[79] Nishikawa J, Saito K, Goto J, Dakeyama F, Matsuo M, Nishihara

    T. New screening methods for chemicals with hormonal activities

    using interaction of nuclear hormone receptor with coactivator.

    Toxicol Appl Pharmacol 1999;154:76–83.[80] Pollak P, Romender G, Hagedorn F, Gelbke H-P. In: Elvers B,

    Hawkins S, Schulz G, editors. Ullman’s encyclopedia of industrial

    chemistry, vol. A17. 5th ed. Weinheim: Wiley/VCH; 1991. p. 363—

    376.[81] Wyatt J, Knowles C. Microbial degradation of acrylonitrile waste

    effluents: the degradation of effluents and condensates from themanufacture of acrylonitrile. Int Biodeterior Biodegr 1995;35:227–

    48.[82] Battistel E, Bernardi A, Maestri P. Enzymatic decontamination of 

    aqueous polymer emulsions containing acrylonitrile. Biotechnol Lett

    1997;19:131–4.[83] Freyssinet G, Peleissier B, Freyssinet M, Delon R. Crops resistant

    to oxynils: from the laboratory to the market. Field Crops Res

    1996;45:125–33.[84] Burton SG. Development of bioreactors for application of 

    biocatalysis in biotransformation and bioremediation. Pure Appl

    Chem 2001;73:77–83.[85] Bollag J-M, Dec J. Characterization of the interaction between

    xenobiotic residues and humic substances. In: Clapp CE, et al.,

    editors. Humic substances and chemical contaminants. Madison,

    WI: SSSA; 2001. p. 155—64.[86] Kim J-E, Wang C, Bollag J-M. Interaction of reactive and inert

    chemicals in the presence of oxidoreductases: reaction of the

    herbicide bentazon and its metabolites with humic monomers.

    Biodegradation 1998;8:387–92.[87] Bollag J-M, Chu R, Rao MA, Gianfreda L. Enzymatic

    oxidative transformation of chlorophenol mixtures. J Environ Qual

    2003;32:62–71.[88] Sannino F, Gianfreda L. Pesticide influence on soil enzymatic

    activities. Chemosphere 2001;22:1–9.[89] Boopathy R. Factors limiting bioremediation technologies. Biores

    Technol 2000;74:63–7.[90] Gianfreda L, Nannipieri P. Basic principles, agents and feasibility

    of bioremediation of soil polluted by organic compounds. Minerva

    Biotecnol 2001;13:5–12.

  • 8/9/2019 Pleurotus s00231

    16/16

    354   L. Gianfreda, M.A. Rao / Enzyme and Microbial Technology 35 (2004) 339–354

    [91] Alexander M. Aging, bioavailability and overestimation of risk from

    environmental pollutants. Environ Sci Technol 2000;34:4259–65.

    [92] Nakamoto S, Machida N. Phenol removal from aqueous solutions by

    peroxidase-catalyzed reaction using additives. Water Res 1992;1:49–

    54.

    [93] Burns RG. Enzyme activity in soil: location and a possible role in

    microbial ecology. Soil Biol Biochem 1982;14:423–7.

    [94] Gianfreda L, Bollag J-M. Influence of natural and anthropogenic

    factors on enzyme activity in soil. In: Stotzky G, Bollag J-M,

    editors. Soil Biochemistry, vol. 9. New York: Marcel Dekker; 1996.

    p. 123—94.

    [95] Gianfreda L, Rao MA, Sannino F, Saccomandi F, Violante A.

    Enzymes in soil: properties, behaviour and potential applications.

    In: Violante A, Huang PM, Bollag J-M, Gianfreda L, editors. Soil

    mineral-organic matter-microorganism interactions and ecosystem

    health development in soil science, vol. 28B. Amsterdam: Elsevier;

    2002. p. 301—28.

    [96] Nannipieri P, Gianfreda L. Kinetics of enzyme reactions in

    soil environments. In: Huang PM, Senesi N, Buffle J, editors.

    Environmental particles: structure and surface reactions of soil

    particles. New York: Wiley; 1998. p. 449—79.

    [97] Gianfreda L, Bollag J-M. Effect of soils on the behaviour of 

    immobilized enzymes. Soil Sci Soc Am J 1994;58:1672–81.[98] van Beilen JB, Li Z. Enzyme technology: an overview. Curr Opin

    Biotechnol 2002;13:338–44.

    [99] Dec J, Bollag J-M. Use of plant material for the decontamination of 

    water polluted with phenols. Biotechnol Bioeng 1994;44:1132–9.

    [100] Roper JC, Dec J, Bollag J-M. Using minced horseradish roots for

    the treatment of polluted waters. J Environ Qual 1996;25:1242–7.

    [101] Greco Jr G, Toscano G, Ciuffi M, Gianfreda L, Sannino F.

    Dephenolisation of olive mill waste-waters by olive husk. Water

    Res 1999;33:3046–50.

    [102] Toscano G, Colarieti ML, Greco Jr G. Oxidative polymerisation

    of phenols by a phenol oxidase from green olives. Enzyme Microb

    Technol 2003;33:47–54.

    [103] OECD. Biotechnology for clean industrial products and processes.

    Paris, France: OECD; 1998.

    [104] Hood EE. From green plants to industrial enzymes. Enzyme Microb

    Technol 2002;30:279–83.

    [105] Chen W, Brhlmann F, Richins RD, Mulchandani A. Engineering

    of improved microbes and enzymes for bioremediation. Curr Opin

    Biotechnol 1999;10:137–41.

    [106] Chen W, Mulchandani A. The use of live biocatalysts for pesticide

    detoxification. Trend Biotechnol 1998;161:71–6.

    [107] Hung S-H, Liao JC. Effects of ultraviolet light irradiation

    in biotreatment of organophosphates. Appl Biochem Biotechnol

    1996;56:37–47.

    [108] Richins RD, Kaneva I, Mulchandani A, Chen W. Biodegradation

    of organophosphorus pesticides by surface-expressed

    organophosphorus hydrolase. Nat Biotechnol 1997;15:984–7.

    [109] Durán N, Rosa MA, D’Annibale A, Gianfreda L. Applications

    of laccases and tyrosinases (phenoloxidases) immobilized on

    different supports: a review. Enzyme Microb Technol 2002;31:907–

    31.

    [110] Gerday C, Aittaleb M, Bentahir M, Chessa J-P, Claverie P, Collins T,

    et al. Cold-adapted enzymes: from fundamentals to biotechnology.

    Trends Biotechnol 2000;18:103–7.

    [111] Haki GD, Rakshit SK. Developments in industrially importantthermostable enzymes: a review. Biores Technol 2003;89:17–

    34.

    [112] Rothschild LJ, Mancinelli RL. Life in extreme environments. Nature

    2001;409:1092–101.

    [113] Hough D, Danson M. Extremozymes. Curr Opin Chem Biol

    1999;3:39–46.

    [114] Ikeda M, Clark D. Molecular cloning of extremely thermostable

    esterase gene from hyperthermophilic archaeon  Pyrococcus furiosus

    in   Escherichia coli. Biotechnol Bioeng 1998;57:624–9.

    [115] Bruins ME, Janssen AE, Boom RM. Thermozymes and their

    applications: a review of recent literature and patents. Appl Biochem

    Biotechnol 2001;90:155–86.