Fact-checked by Grok 2 weeks ago

Microbial biodegradation

Microbial biodegradation is the process by which microorganisms, primarily and fungi, enzymatically decompose complex organic compounds into simpler, non-toxic substances, such as , water, and inorganic ions, allowing their integration into natural biogeochemical cycles. This natural phenomenon relies on the metabolic activities of diverse microbial communities, including genera like , , and , which produce specialized enzymes such as oxidases, hydrolases, and oxygenases to initiate and complete the degradation pathway. Under aerobic conditions, the process culminates in mineralization, while environments yield products like , enabling the breakdown of recalcitrant pollutants in oxygen-limited settings such as sediments or landfills. Key mechanisms of microbial biodegradation involve biodeterioration (initial surface colonization), biofragmentation (cleavage into oligomers), and bioassimilation (incorporation into microbial biomass), often enhanced by microbial consortia that cooperate through horizontal gene transfer and metabolic complementarity. Common substrates include petroleum hydrocarbons, synthetic plastics like polyethylene and polystyrene, and pharmaceuticals such as ibuprofen, with degradation rates influenced by factors like pH, temperature, nutrient availability, and pollutant concentration. For instance, certain Pseudomonas species can degrade up to 90% of polycyclic aromatic hydrocarbons (PAHs) in contaminated soils within weeks under optimal conditions. In environmental applications, microbial biodegradation forms the basis of strategies, including (adding nutrients to boost native microbes) and (introducing specialized strains), which have successfully treated oil spills, industrial effluents, and residues at sites worldwide. Its eco-friendly and cost-effective nature makes it preferable to chemical or physical methods, though challenges like slow degradation of persistent organics persist, driving into engineered microbes and optimization. Overall, this process underscores microorganisms' pivotal role in sustaining by carbon and mitigating .

Fundamentals

Definition and Scope

Microbial biodegradation refers to the process by which microorganisms, primarily and fungi, enzymatically decompose complex organic compounds into simpler substances such as (CO₂), (H₂O), (CH₄), or microbial . This biological transformation harnesses the metabolic capabilities of these organisms to break down both naturally occurring materials and synthetic pollutants, often under aerobic or conditions. Unlike abiotic , which involves non-biological physical or chemical processes like or photolysis, microbial biodegradation relies on living cells and their enzymes to initiate and sustain the breakdown, ensuring the process is energy-yielding for the microbes involved. The foundations of understanding microbial biodegradation trace back to the , with Pasteur's pioneering observations on microbial , where he demonstrated that microorganisms drive the of organic substances in nutrient-rich environments. These early insights into laid the groundwork for later applications, with modern recognition emerging in the amid growing concerns over industrial ; by the mid-1900s, researchers began documenting microbial roles in degrading compounds, leading to the formal development of strategies in the 1970s. This evolution marked a shift from basic studies to targeted environmental applications, highlighting microbes' adaptability to novel substrates. The scope of microbial biodegradation encompasses a wide range of substrates, including natural such as and lignocellulosic materials in ecosystems, as well as xenobiotics like pesticides, hydrocarbons, and plastics introduced by human activity. In natural settings, it facilitates the of in soils and aquatic environments, whereas in contaminated sites, it targets persistent pollutants that resist abiotic breakdown. This process is integral to global nutrient cycling, where microbes mineralize organic nitrogen, phosphorus, and other elements, promoting ; it also contributes to by incorporating degraded carbon into stable , potentially mitigating atmospheric CO₂ levels. Furthermore, in , it addresses environmental contamination, with global estimates indicating that microbial activity processes over 1 billion tons of organic waste from alone annually, underscoring its scale in . A fundamental representation of complete aerobic mineralization is given by the stoichiometric for a generalized : \mathrm{C_xH_yO_z + \left(x + \frac{y}{4} - \frac{z}{2}\right) O_2 \rightarrow x\, CO_2 + \frac{y}{2} H_2O} Here, the oxygen requirement n = x + \frac{y}{4} - \frac{z}{2} balances the oxidation of carbon and hydrogen while accounting for oxygen already present in the , ensuring and ; for example, in glucose (\mathrm{C_6H_{12}O_6}), n = 6, yielding 6 CO₂ and 6 H₂O molecules per glucose unit degraded. This illustrates the of aerobic processes in fully mineralizing organics to inorganic end products, releasing energy for microbial growth.

Microorganisms and Enzymes Involved

Microbial biodegradation primarily involves a diverse array of , with genera such as , , , and playing prominent roles in the degradation of s and other organic pollutants. species, including P. putida and P. aeruginosa, are frequently isolated from contaminated sites and exhibit versatile catabolic capabilities for aromatic compounds like polycyclic aromatic s (PAHs). Similarly, strains, such as R. erythropolis, demonstrate robust degradation of aliphatic and aromatic s due to their metabolic diversity and tolerance to toxic substrates. , particularly methanogenic species like those in the orders Methanobacteriales and Methanosarcinales, contribute to biodegradation in anaerobic environments by facilitating syntrophic partnerships that enable breakdown coupled to production. Fungi provide a complementary role through extracellular enzyme , aiding in the of recalcitrant polymers in and habitats. Biodegradation often relies on microbial consortia rather than single strains, as consortia enhance efficiency through metabolic cooperation and division of labor in breaking down complex substrates. Single strains may possess limited catabolic pathways, whereas consortia can sequentially metabolize intermediates produced by initial degraders, leading to more complete removal. For instance, combinations of and isolates have shown synergistic effects in oily sludge degradation compared to individual cultures. Central to these processes are key enzymes, including oxidoreductases such as monooxygenases that initiate the attack on inert hydrocarbons by incorporating oxygen atoms. hydroxylases, a subset of monooxygenases like AlkB and CYP153, catalyze the terminal oxidation of , converting them to alcohols as the first step in degradation. Hydrolases, exemplified by lipases, hydrolyze bonds in and synthetic esters, facilitating the breakdown of derivatives. Lyases contribute by cleaving carbon-carbon bonds in unsaturated compounds, supporting further . The genetic basis for this catabolic versatility often involves s and operons that encode degradation pathways. The TOL (pWW0) in Pseudomonas putida mt-2, for example, carries genes for degradation, including those for meta-cleavage pathway enzymes, enabling aerobic breakdown of aromatic hydrocarbons. Such allow rapid adaptation to pollutants via horizontal transfer. More than 200 genera across , algae, and fungi have been identified as capable of hydrocarbon degradation, reflecting the broad microbial diversity adapted to contaminated environments. Extremophiles, such as thermophilic species like B. licheniformis, extend this capability to harsh conditions, degrading long-chain alkanes at elevated temperatures. Enzyme kinetics in follow the Michaelis-Menten model, where substrate affinity is quantified by the Michaelis constant (Km), typically in the micromolar range for aromatic pollutants like and . For instance, toluene 3-monooxygenase from exhibits a Km of approximately 13–250 μM for , indicating efficient binding at environmentally relevant concentrations. v = \frac{V_{\max} [S]}{K_m + [S]} This equation describes the initial reaction velocity (v) as a function of substrate concentration ([S]), maximum velocity (Vmax), and Km, underscoring how low Km values enhance degradation rates for low-abundance pollutants.

Biodegradation Mechanisms

Aerobic Biodegradation

Aerobic biodegradation refers to the oxygen-dependent process by which microorganisms catabolize organic pollutants, utilizing molecular oxygen (O₂) as the terminal electron acceptor in the respiratory electron transport chain to generate energy via oxidative phosphorylation. This enables the complete mineralization of substrates to carbon dioxide (CO₂), water (H₂O), and biomass, distinguishing it from partial degradation in oxygen-limited conditions. The process supports both primary metabolism, where the pollutant directly serves as the carbon and energy source, and cometabolism, in which non-growth substrates are oxidized by enzymes or cofactors produced during the metabolism of a primary substrate, often without providing energy to the microbe. Degradation generally unfolds in sequential stages: initial peripheral oxidation to introduce functional groups, central ring cleavage for aromatic compounds via dioxygenases or monooxygenases, and funneling of intermediates into the tricarboxylic acid (TCA) cycle for full oxidation. A prominent example of aerobic pathways occurs in the degradation of BTEX compounds (, , , and xylenes), common pollutants from releases. In the TOL plasmid-encoded pathway of , undergoes initial oxidation of its through and to form benzoate, followed by conversion to and subsequent meta-cleavage by catechol 2,3-dioxygenase, yielding intermediates like 2-hydroxymuconic semialdehyde that enter the cycle. This pathway exemplifies the efficiency of aerobic ring activation, allowing rapid breakdown under oxic conditions, though alternative routes like the TOD pathway in other strains initiate direct dioxygenase attack on the aromatic ring to form cis-dihydrodiol intermediates before meta-cleavage. Efficiency of aerobic biodegradation is highest in well-oxygenated environments, such as soils maintaining O₂ concentrations above 5% (v/v), where microbial proceeds without limitation; below this threshold, rates decline due to competition with processes or oxygen scarcity. In or aqueous systems, dissolved oxygen levels exceeding 2 mg/L sustain optimal activity, with supplemental oxygen delivery accelerating natural rates by at least an for hydrocarbons. Reported degradation rates for hydrocarbons under enhanced aerobic conditions range from 10 to 100 mg/L/day, depending on bioavailability, microbial inoculum, and availability, though recalcitrant fractions may persist. Beyond BTEX, aerobic processes target diverse pollutants, including pesticides like , which undergoes N-dealkylation by hydrolases in such as Pseudomonas sp. ADP, sequentially removing ethyl groups to form hydroxyatrazine before dechlorination and ring opening. Polycyclic aromatic hydrocarbons (PAHs), such as , are degraded via initial dioxygenase-mediated incorporation of two oxygen atoms to form cis-naphthalene dihydrodiol, followed by dehydrogenation to naphthol and meta-cleavage, integrating into central . A simplified representation of the initial oxidation step in these pathways is: \text{Pollutant} + \text{O}_2 + \text{NADH} \rightarrow \text{oxidized intermediate} + \text{H}_2\text{O} + \text{NAD}^+ with subsequent chain reactions leading to complete mineralization to CO₂. Microbes like Pseudomonas species exemplify versatile utilization of these oxygen-requiring pathways for pollutant remediation.

Anaerobic Biodegradation

Anaerobic biodegradation occurs in environments devoid of molecular oxygen, where microorganisms utilize alternative terminal electron acceptors to facilitate the oxidation of organic pollutants, enabling energy conservation through respiration or fermentation. The sequence of electron acceptors typically follows a redox gradient, starting with nitrate (NO₃⁻) under denitrifying conditions, progressing to manganese(IV) (Mn(IV)) and iron(III) (Fe(III)) reduction, then sulfate (SO₄²⁻) reduction, and finally CO₂ reduction in methanogenic settings. This ordered utilization reflects the thermodynamic favorability of each process, with higher-potential acceptors depleted before lower ones become dominant. In many cases, degradation relies on syntrophic interactions within microbial consortia, where primary degraders produce intermediates like hydrogen (H₂) or formate that partner organisms, such as methanogens or sulfate reducers, consume to maintain low partial pressures and drive the overall reaction forward. Key pathways in anaerobic biodegradation include reductive dechlorination for chlorinated solvents and for hydrocarbons. In reductive dechlorination, sequentially remove chlorine atoms from compounds like perchloroethylene (PCE), transforming it through trichloroethene (TCE) and cis-dichloroethene (cis-DCE) to non-toxic ethene, using H₂ as an . This process is mediated by specialized dehalogenating and is prominent in contaminated aquifers. For hydrocarbons, pathways often involve initial activation via fumarate addition or , leading to breakdown products such as and H₂, which support downstream syntrophic . Efficiency of anaerobic biodegradation is generally lower than aerobic processes due to the lower energy yield from alternative acceptors, with in situ rates for hydrocarbons and chlorinated solvents typically ranging from 0.003 to 0.05 mg/L/day in sediments, though lab enrichments can achieve higher values up to several mg/L/day. These processes are prevalent in anoxic sediments, landfills, and plumes, where oxygen levels are minimal, but even trace O₂ can inhibit dehalogenases and other anaerobic enzymes, halting degradation. Unlike aerobic biodegradation, which yields complete mineralization to CO₂, anaerobic routes often result in partial products like (CH₄) or . Representative examples illustrate these mechanisms. Perchloroethylene (PCE) undergoes complete reductive dechlorination to ethene by Dehalococcoides species, such as D. ethenogenes strain 195, in anaerobic consortia from contaminated sites, with specific dechlorination activities reaching 69 nmol Cl⁻ released/min/mg protein. Similarly, is degraded under sulfate-reducing conditions in petroleum-contaminated sediments, where inoculated sulfate-reducing oxidize to CO₂, coupled to the reduction of SO₄²⁻ to HS⁻, following the 4C₆H₆ + 15SO₄²⁻ + 12H₂O → 24HCO₃⁻ + 15HS⁻ + 9H⁺. A notable process in methanogenic environments is the sulfate-dependent anaerobic oxidation of methane (AOM), which reverses methanogenesis and consumes CH₄ as an energy source: \text{CH}_4 + \text{SO}_4^{2-} \rightarrow \text{HCO}_3^- + \text{HS}^- + \text{H}_2\text{O} This reaction, mediated by consortia of methanotrophic (ANME) and sulfate-reducing , plays a key role in mitigating CH₄ emissions from sediments and contributes to the partial breakdown of biogenic hydrocarbons.

Key Factors

Bioavailability and Transport

Bioavailability refers to the fraction of a pollutant that is accessible for uptake by microorganisms, determined by physical, chemical, and biological interactions in the soil matrix. This accessible portion, often termed the freely dissolved concentration (C_free), is crucial for initiating biodegradation, as only the bioavailable fraction can cross microbial membranes for enzymatic attack. The of pollutants is primarily influenced by to components and their aqueous . , quantified by the distribution coefficient (K_d), describes the partitioning between the aqueous phase and solid phases, such as or minerals; for polycyclic aromatic hydrocarbons (PAHs), K_d values typically range from 10 to 1000 L/kg in soils, leading to strong retention that reduces the mobile fraction available for microbes. Similarly, low water limits dissolution rates; for example, polychlorinated biphenyls (PCBs) exhibit solubilities below 1 mg/L, particularly for highly chlorinated congeners, hindering their transfer into the bioavailable phase. Pollutant transport in soil porous media governs the delivery of bioavailable fractions to microbial communities through mechanisms like and . involves molecular movement from high to low concentration areas within pores or , while transports dissolved via bulk water flow, such as movement or rainfall infiltration. formation by microbial consortia can further enhance local concentrations by creating structured communities that trap and concentrate substrates near degrading cells, improving uptake efficiency in heterogeneous environments. Techniques such as (SPME) are employed to measure by equilibrating a polymer fiber with the sample to quantify the freely dissolved fraction, mimicking passive uptake by organisms. Key environmental factors, including and (TOC) content—typically 1-5% in many s—modulate these processes; higher TOC promotes sorption to , reducing , while shifts can alter binding and desorption kinetics. Low significantly constrains rates, often limiting overall degradation progress, with aged residues showing greater recalcitrance due to in aggregates over time, as seen in weathered oil spills. This limitation arises because microbes cannot access sorbed or undissolved fractions, stalling overall remediation progress despite enzymatic potential. Recent advances, such as palladium nanoparticles (Pd-NPs), have shown promise in enhancing degradation of highly chlorinated PCBs by facilitating C–Cl bond breakage, improving access to recalcitrant pollutants as of 2025. Partitioning models describe the bioavailable fraction quantitatively, where the freely dissolved concentration relates to total via : C_{\text{free}} = \frac{C_{\text{total}}}{1 + K_d \cdot \rho_s \cdot f_s} Here, the bioavailable fraction approximates \frac{C_{\text{free}}}{C_{\text{total}}} = \frac{1}{1 + K_d \cdot \text{solid phase}}, with \rho_s as solid density and f_s as solid fraction, highlighting how high K_d values diminish .

Chemotaxis and Microbial Motility

refers to the directed movement of microorganisms, particularly , in response to chemical gradients, enabling them to navigate toward favorable substrates or away from toxins. In the context of microbial biodegradation, this process is primarily mediated by methyl-accepting chemotaxis proteins (MCPs), such as and Tsr in , which act as transmembrane receptors detecting environmental chemicals. These receptors initiate flagella-based motility, where use rotary flagella to propel themselves through run-and-tumble patterns, achieving swimming speeds of 20–50 μm/s depending on the species, such as approximately 20 μm/s in E. coli and up to 40 μm/s in . The underlying mechanism involves via the Che pathway, a conserved two-component system in motile . When an attractant binds to an MCP, it modulates the autophosphorylation of CheA kinase, reducing the frequency of flagellar reversals (tumbles) and promoting smooth "runs" toward higher concentrations; conversely, repellents increase tumbling to reorient away from unfavorable gradients. This biased is powered by the proton motive force for flagellar rotation, though the signaling cascade requires ATP for processes like receptor by and demethylation by , imposing an energetic cost estimated at thousands of ATP molecules per cell cycle to maintain responsiveness. Integration with further modulates motility, as autoinducer signals can coordinate with formation, enhancing collective migration toward pollutants in dense populations. In biodegradation, chemotaxis significantly boosts efficiency by increasing microbial encounter rates with pollutants, often by 10–100-fold compared to passive diffusion in heterogeneous environments, thereby accelerating substrate access and degradation. For instance, PAH-degrading bacteria like Burkholderia sp. exhibit chemotaxis toward compounds such as naphthalene and pyrene, allowing them to concentrate at contamination hotspots and upregulate degradative genes upon arrival. Similarly, Pseudomonas putida G7 demonstrates enhanced naphthalene breakdown in aqueous systems through directed motility, underscoring chemotaxis's role in overcoming spatial barriers to bioremediation. The chemotactic velocity v can be modeled as v = \chi \nabla C where \chi is the chemotactic coefficient reflecting sensitivity to the gradient, and \nabla C is the concentration gradient of the attractant. This equation captures the drift component of bacterial movement, highlighting how steep gradients drive faster directed migration essential for targeting biodegradable pollutants.

Case Studies

Oil and Hydrocarbon Biodegradation

Microbial biodegradation plays a crucial role in the natural attenuation of hydrocarbons, which are complex mixtures released during exploration, transportation, and spills. These substrates primarily consist of alkanes, including linear and branched forms, as well as aromatic compounds such as BTEX (, , , and xylenes) and polycyclic aromatic hydrocarbons (PAHs). Linear alkanes are generally more susceptible to microbial attack than branched alkanes, which degrade more slowly due to steric hindrance, while small aromatics like BTEX break down faster than larger PAHs. Among n-alkanes, chain lengths from C10 to C40 are most readily degradable by common hydrocarbonoclastic bacteria, as shorter chains (C5-C9) may volatilize quickly and longer ones (beyond C40) exhibit reduced . In marine environments, specialized dominate the of these s, with genera such as Alcanivorax and Oceanospirillum emerging as key players during oil spills. Alcanivorax species, particularly A. borkumensis, are obligate hydrocarbon degraders that preferentially metabolize alkanes through the beta-oxidation pathway, converting them into fatty acids and subsequently for energy production. Oceanospirillum, often found in oil plumes, contributes to the degradation of both aliphatic and aromatic fractions in consortia, enhancing overall efficiency via synergistic interactions. Aerobic pathways predominate for alkane breakdown in oxygenated surface waters. These microbes often exhibit toward oil droplets, facilitating initial contact and colonization. A prominent case study is the 2010 in the , where approximately 4.9 million barrels of crude oil were released, prompting a rapid microbial response. The application of chemical dispersants like Corexit 9500 aimed to enhance oil emulsification and bioavailability by increasing the surface area of hydrocarbon droplets. Studies indicate mixed effects on microbial degradation, with an estimated 50-60% of certain hydrocarbons degraded within weeks to months, primarily driven by Alcanivorax for alkanes and Oceanospirillales (including relatives of Oceanospirillum) for aromatics, as identified through stable isotope probing and cultivation studies. Despite these advances, challenges persist in oil biodegradation, particularly with high-molecular-weight fractions such as multi-ring PAHs, which exhibit through carcinogenic and hemotoxic effects, inhibiting microbial activity and reducing degradation rates. Their hydrophobicity further limits bioavailability, prolonging persistence in the environment. strategies address these issues by introducing enriched consortia of degraders, such as halotolerant and species, which can achieve up to 79% degradation of (a high-MW PAH) and 96% of long-chain alkanes like tetracosane in saline conditions. Under favorable aerobic conditions, surface oil spills can achieve substantial , with some studies reporting up to 70-90% removal within 1-2 years, influenced by factors like , availability, and oxygen levels.

Cholesterol and Sterol Biodegradation

Microbial biodegradation of , a with the molecular formula C_{27}H_{46}O, and related bile acids plays a crucial role in , particularly in processing sterol-rich substrates from anthropogenic and natural sources. These compounds originate primarily from effluents, such as those from plants and industrial discharges, as well as animal wastes including fecal matter from and . is present in dairy industry effluents due to processing residues, necessitating targeted microbial interventions for effective treatment. The aerobic degradation pathway in begins with the oxidation of 's 3β-hydroxyl group to a by enzymes, yielding cholest-4-en-3-one as the initial product: \text{[Cholesterol](/page/Cholesterol)} + \text{O}_2 \rightarrow \text{cholest-4-en-3-one} + \text{H}_2\text{O} This step is followed by side-chain cleavage, primarily mediated by monooxygenases such as Cyp125 and Cyp142, which hydroxylate the C-26/C-27 positions, leading to β-oxidation-like breakdown into propionyl-CoA and units. Subsequent ring oxidation involves (Hsd) to form cholest-4-ene-3,17-dione, followed by 3-ketosteroid Δ¹-dehydrogenase (KstD) and 3-ketosteroid 9α-hydroxylase (KshAB), culminating in ring cleavage to and further metabolites. Key microbes capable of this process include of , such as M. smegmatis and M. tuberculosis, which utilize as a sole carbon source. variants, observed in low-oxygen environments like the , involve an oxygen-independent 2,3-seco pathway initiated by cholesterol dehydrogenase/isomerase (AcmA), with such as Sterolibacterium denitrificans oxidizing ring A before side-chain processing. Complete mineralization to CO₂ and is rare and typically incomplete, often halting at central intermediates due to pathway limitations. These degradation capabilities have practical applications in treating cholesterol-laden effluents, particularly in the dairy industry where microbial consortia enhance pollutant removal in systems. Mycobacterium species, abundant in plants, can increase in relative abundance by up to 4.7-fold during cholesterol enrichment, facilitating degradation rates of approximately 1-5 mg/L/day under optimized conditions. Anaerobic processes in Sterolibacterium contribute to breakdown in oxygen-depleted niches, such as anaerobic digesters. A unique aspect of sterol biodegradation is its evolutionary conservation, originating from actinobacterial lineages like Mycobacterium, where gene clusters (e.g., hsa and ksh) have been horizontally transferred to other phyla, enabling widespread microbial adaptation to sterol-rich environments.

Plastics and Synthetic Polymers Biodegradation

Plastics and synthetic polymers, such as (PET), (PE), and polyurethane, exhibit low natural degradability primarily due to their hydrophobic nature, which limits microbial access and enzymatic attack. These synthetic materials, widely used in , textiles, and foams, persist in the environment for decades because their high molecular weight and crystalline structure resist and oxidation under ambient conditions. In natural settings, biodegradation rates for these polymers are typically less than 1% per year, constrained by poor and the absence of specialized degraders. Microbial degradation of these plastics proceeds via enzymatic hydrolysis, breaking ester or carbon-carbon bonds to depolymerize polymers into monomers that can enter central metabolic pathways. A landmark example is the PETase enzyme from the bacterium Ideonella sakaiensis, discovered in 2016, which initiates PET breakdown by hydrolyzing ester linkages. The reaction can be represented as: \text{PET} + \text{H}_2\text{O} \xrightarrow{\text{PETase}} \text{MHET} + \text{EG} where MHET is mono(2-hydroxyethyl) terephthalic acid and EG is ethylene glycol; subsequent ring cleavage of terephthalate by enzymes like MHETase enables complete mineralization to CO₂ and water. For PE, oxidation precedes hydrolysis, while polyurethane degradation involves urethanases targeting urethane bonds. Key microbes include bacteria such as Rhodococcus ruber for PE, which forms biofilms and reduces polymer weight through laccase activity, and fungi like Aspergillus species for polyvinyl chloride (PVC), which secrete oxidases to initiate dechlorination. However, single strains often achieve only partial degradation, necessitating microbial consortia for synergistic complete mineralization via complementary enzyme profiles. Recent advances in the have focused on to accelerate these processes. The FAST-PETase variant, developed using machine learning-guided mutations, exhibits up to sixfold higher catalytic efficiency on at 50°C compared to the wild-type , enabling near-complete of post-consumer in under a week. As of 2025, advances include machine learning-optimized enzymes achieving higher efficiency and scalable systems for . with such engineered microbes or consortia can elevate field degradation rates to around 10% per year in optimized conditions, far surpassing natural rates, though scalability remains limited. A major challenge is the persistence of derived from these polymers, which, due to their small size and increased surface area, evade complete biodegradation and accumulate in ecosystems, complicating remediation efforts.

Applications and Advances

Waste Biotreatment Analysis

Waste biotreatment analysis encompasses a range of techniques and metrics used to evaluate the efficiency of microbial degradation processes in managing organic waste streams. Respirometry is a primary method that measures oxygen uptake or production to assess microbial rates in aerobic systems, providing insights into the biodegradable fractions of (COD) in . This approach quantifies the readily and slowly degradable COD components by monitoring O2/CO2 fluxes, enabling optimization of treatment conditions such as aeration rates. Complementing respirometry, labeling techniques, particularly with 14C-labeled substrates, track the mineralization of s by measuring the release of radiolabeled CO2, which indicates the percentage of complete degradation to inorganic products. For instance, in systems, 14C tracking has been employed to determine the extent of nonvolatile organic compound mineralization, distinguishing between into and full breakdown. Key metrics in waste biotreatment include degradation efficiency, often expressed as COD removal percentages, and half-life calculations that model pollutant persistence. In activated sludge processes, typical COD removal efficiencies range from 80-95%, reflecting robust microbial activity under optimal conditions like sufficient nutrient balance and hydraulic retention times. Half-life (t_{1/2}) is derived from kinetic models to predict the time required for 50% degradation, aiding in design and performance forecasting; for readily biodegradable substances in waste environments, half-lives can be as short as 1-7 days. These metrics are applied across aerobic and anaerobic digesters, where microbial densities in municipal treatments typically range from 10^6 to 10^9 colony-forming units (CFU) per milliliter, supporting high-rate breakdown. Aerobic digesters rely on oxygen-dependent consortia for rapid COD reduction, while anaerobic systems produce through methanogenic pathways, both scalable for handling large volumes like urban . Degradation kinetics are frequently modeled using first-order equations to describe substrate concentration decline over time: \ln\left(\frac{C}{C_0}\right) = -kt where C is the concentration at time t, C_0 is the initial concentration, k is the rate constant (typically 0.1-1 day^{-1} for many organic pollutants in biotreatment systems), and t is time. This model assumes driven by microbial activity and is validated through experimental data from respirometry or studies. However, challenges such as inhibition by like or lead can reduce degradation rates by disrupting function or microbial viability, necessitating pre-treatment or adaptive consortia. To monitor and mitigate these issues, quantitative (qPCR) targets functional genes such as alkB, which encodes alkane monooxygenases involved in degradation, allowing real-time assessment of degradative potential in digester communities. For example, elevated alkB copy numbers correlate with enhanced breakdown in contaminated waste streams, guiding operational adjustments.

Metabolic Engineering and Biocatalysis

involves the targeted modification of microbial genomes and metabolic pathways to enhance the biodegradation of recalcitrant pollutants, leveraging tools like -Cas systems to assemble novel degradation pathways. For instance, editing has been applied to to integrate and express and MHETase enzymes from , enabling efficient of () into monomers like and . This approach, demonstrated in the 2020s, allows for surface display or fusion constructs that improve enzyme stability and substrate access, achieving reported PET depolymerization efficiencies up to approximately 63% under optimized conditions compared to lower efficiencies in wild-type systems. Synthetic biology extends these efforts by designing microbial consortia or single strains capable of simultaneous degradation of multiple pollutants, such as heavy metals and organic xenobiotics, through the construction of artificial genetic circuits. A versatile synthetic Vibrio natriegens strain, VCOD-15, engineered via plasmid-based pathway integration, has shown concurrent removal of aromatic organic pollutants like naphthalene, phenol, and toluene in hypersaline environments, reducing residual levels to below 2% after treatment. These designs often incorporate genetic circuits for efficient expression, ensuring targeted activation in contaminated sites and minimizing off-target effects. In biocatalysis, enzymes are isolated and immobilized on supports like nanoparticles or membranes for ex situ of effluents, enhancing reusability and . Laccase enzymes, immobilized via covalent binding to silica or magnetic frameworks, have been used for dye decolorization, retaining over 90% activity for 10-20 cycles in batch reactors treating azo dyes like Cibacron Blue. This immobilization reduces enzyme leaching and allows continuous operation, with turnover numbers exceeding 500 per cycle for oxidative breakdown of chromophores. Specific examples highlight the practical impact of these strategies. Engineered strains, modified through of biphenyl dioxygenase genes, exhibit up to 2-fold faster degradation of polychlorinated biphenyls (PCBs) in microcosms, converting highly chlorinated congeners to less toxic chlorobenzoic acids. Commercially, ' BioRemove™ series employs engineered microbial blends for in wastewater, achieving over 80% reduction in contaminants in industrial applications. These engineered systems offer benefits including heightened substrate specificity, which minimizes incomplete mineralization and byproduct toxicity, alongside improved yields—such as 85-90% pollutant removal versus 40-50% in unmodified microbes—through optimized enzyme cascades. (FBA) models aid pathway optimization by solving steady-state constraints, where the flux vector \mathbf{v} satisfies S \mathbf{v} = 0, with S as the stoichiometric matrix, enabling prediction of maximal rates under limitations. As of 2025, advances include AI-guided mutations in enzymes via , enhancing degradation rates for plastics in environmental applications.

Fungal Biodegradation

Fungi, particularly within the phylum such as white-rot species including chrysosporium, are prominent in microbial biodegradation due to their production of extracellular enzymes that target insoluble, recalcitrant substrates like lignocellulosic materials. These enzymes enable the fungi to access and break down complex polymers externally, facilitating degradation in natural environments where substrates are embedded in solid matrices. fungi, such as certain endophytic species, also participate in biodegradation, often contributing to the hydrolysis of synthetic materials through similar extracellular mechanisms. A key pathway in fungal biodegradation involves the of , primarily by white-rot basidiomycetes via the enzymes lignin peroxidase (), manganese peroxidase (MnP), and , which generate non-specific radicals to cleave aromatic structures. , in particular, catalyzes the oxidation of using , initiating the formation of phenoxy radicals that lead to ring opening and eventual mineralization to CO₂. This process can be represented as: \text{Lignin (phenolic unit)} + \text{H}_2\text{O}_2 \xrightarrow{\text{LiP}} \text{phenoxy radical} + \text{H}_2\text{O} \quad \rightarrow \quad \text{depolymerization products} \quad \rightarrow \quad \text{CO}_2 + \text{H}_2\text{O} White-rot fungi achieve lignin degradation rates of up to 50% in substrates like tobacco stalk over 30 days under solid-state fermentation conditions. Representative examples of fungal biodegradation include the degradation of pesticides such as DDT by the basidiomycete Pleurotus ostreatus, which achieves approximately 19% removal in 7 days through extracellular oxidation. For plastics, the ascomycete Pestalotiopsis microspora degrades polyurethane as its sole carbon source, even under anaerobic conditions, via esterase activity that hydrolyzes polymer bonds. These capabilities highlight fungi's versatility in addressing xenobiotic compounds. Fungi offer advantages in biodegradation due to their inherent tolerance to high concentrations of toxins, such as and organic pollutants, which allows them to thrive in contaminated environments where may falter. Additionally, in symbiotic mycorrhizal associations, fungi enhance breakdown by accelerating through exudation and mobilization, contributing to carbon cycling in ecosystems. Unlike bacterial processes that often involve intracellular , fungal strategies emphasize extracellular radical-based attacks on insoluble substrates.

References

  1. [1]
    Systemic approaches to biodegradation | FEMS Microbiology Reviews
    Biodegradation: a systemic process · Alexander, 2001). One of the bioremediation strategies is biostimulation, which aims to enhance the local bacterial ...
  2. [2]
    Microbial Biodegradation - an overview | ScienceDirect Topics
    Biological treatment of MPs including “in vivo degradation”, “micro-biological degradation”, and “enzymatic degradation” is known as the Biodegradation process.
  3. [3]
    Bioremediation of environmental wastes: the role of microorganisms
    Bioremediation involves using biological agents such as plants and microbes to remove or lessen the effects of environmental pollutants.
  4. [4]
    Introduction to Microbiology - GMU
    French chemist, Louis Pasteur, first recognized fermentation as a microbial ... Biodegradation and bioremediation involves on-site use of microbial activities to ...
  5. [5]
    Thirty Years and Counting: Bioremediation in Its Prime? | BioScience
    Mar 1, 2005 · Bioremediation in its formal sense, meaning any use of living organisms to degrade wastes, has been practiced since humans first populated ...
  6. [6]
    Microbial carbon use efficiency promotes global soil carbon storage
    May 24, 2023 · A classical paradigm emphasizes the roles of plant carbon inputs and soil organic matter decomposition in driving SOC storage and persistence.
  7. [7]
    Trends in Solid Waste Management - World Bank
    Sep 20, 2018 · The world generates 2.01 billion tonnes of municipal solid waste annually, with at least 33 percent of that—extremely conservatively—not ...
  8. [8]
    Stoichiometry of the Aerobic Biodegradation of the Organic Fraction ...
    The amount of oxygen required for biodegradation of organic fraction of MSW was estimated on the basis of stoichiometric equations and increased from 0.92 moles ...
  9. [9]
    Insights into polyethylene biodegradative fingerprint of ...
    Dec 12, 2024 · The main bacterial genera involved in PE biodegradation are Bacillus, Pseudomonas, and Rhodococcus (Soleimani et al., 2021; Montazer et al., ...
  10. [10]
    Pseudomonas and Pseudarthrobacter are the key players in ... - NIH
    May 25, 2024 · Members of the genus Pseudomonas are the dominant PAH-degrading bacteria and are cold-adapted indigenous bacteria in Antarctic soils. These ...
  11. [11]
    Microbial Degradation of Petroleum Hydrocarbon Contaminants
    Microbial consortium consisting of two isolates of Pseudomonas aeruginosa and one isolate Rhodococcus erythropolis from soil contaminated with oily sludge was ...
  12. [12]
    Comparative Genomics of the Rhodococcus Genus Shows Wide ...
    May 21, 2020 · The genus Rhodococcus exhibits great potential for bioremediation applications due to its huge metabolic diversity, including biotransformation of aromatic and ...
  13. [13]
    Membrane transport systems and the biodegradation potential ... - NIH
    Apr 4, 2014 · The Rhodococcus genus contains species with remarkable ability to tolerate toxic compounds and to degrade a myriad of substrates.
  14. [14]
    Non-syntrophic methanogenic hydrocarbon degradation by ... - Nature
    Dec 22, 2021 · The methanogenic degradation of oil hydrocarbons can proceed through syntrophic partnerships of hydrocarbon-degrading bacteria and methanogenic archaea.Missing: biodegradation | Show results with:biodegradation
  15. [15]
    Anaerobic hydrocarbon biodegradation by alkylotrophic ...
    This process was previously assigned to the syntrophy of hydrocarbon-degrading bacteria and methanogenic archaea. ... biodegradation by alkylotrophic methanogens ...
  16. [16]
    Diverse Metabolic Capacities of Fungi for Bioremediation - PMC
    Fungi play a major role as decomposers and symbionts in all ecosystems including soil and aquatic habitats owing to their robust morphology and diverse ...
  17. [17]
    Microbial Consortia Are Needed to Degrade Soil Pollutants - PMC
    Compared to the culture of a single strain, the performance of the consortium of microorganisms is better. The microbial consortium showed apparent effects in ...
  18. [18]
    Construction of microbial consortia for microbial degradation of ...
    Dec 5, 2022 · Many plastics are biodegraded by microbial consortia rather than individual strains, possibly because of the limited metabolic capacity of ...
  19. [19]
    Pollutant Degrading Enzyme: Catalytic Mechanisms and Their ... - NIH
    Aug 6, 2021 · Therefore, the study of new and effective degradation mechanisms for pollutants has become an international focal point. Bioremediation was ...
  20. [20]
    Diverse alkane hydroxylase genes in microorganisms and ... - NIH
    May 15, 2014 · AlkB and CYP153 are important alkane hydroxylases responsible for aerobic alkane degradation in bioremediation of oil-polluted environments ...Missing: oxidoreductases lyases
  21. [21]
    Potential for and Distribution of Enzymatic Biodegradation of ... - MDPI
    Major enzyme classes that may be involved in PS degradation such as hydrolases and monooxygenases ... alkane biodegradation mechanisms of alkane hydroxylases ...
  22. [22]
    Molecular and functional analysis of the TOL plasmid pWWO from ...
    The genetic organization of the Pseudomonas putida plasmid pWWO-161, which encodes enzymes for the degradation of toluene and related aromatic hydrocarbons,Missing: basis | Show results with:basis
  23. [23]
    Complete Nucleotide Sequence of TOL Plasmid pDK1 Provides ...
    The genes associated with the degradation of man-made xenobiotic compounds (e.g., atrazine, 2,4-dichlorophenoxyacetate, and haloacetates) have predominantly ...
  24. [24]
    Hydrocarbon-Degrading Bacteria and the Bacterial Community ...
    Indeed, more than 200 bacterial, algal, and fungal genera, encompassing over 500 species, have been recognized as capable of hydrocarbon degradation (see ...
  25. [25]
    Biotechnological Potential of Extremophiles: Environmental ... - PMC
    licheniformis MN6 pure isolates and a consortium on the degradation of long-chain n-alkanes (C32 and C40). Biodegradation efficiencies were higher for C32 (90%) ...
  26. [26]
    Toluene 3-monooxygenase of Ralstonia pickettii PKO1 is ... - PubMed
    TG1/pBS(Kan)T3MO cells expressing T3MO oxidized toluene at a maximal rate of 11.5 +/- 0.33 nmol/min/mg of protein with an apparent Km value of 250 microM and ...
  27. [27]
    Expression and purification of the recombinant subunits of toluene
    and 13.3 ± 1.3 lM, respectively. It should be noted that the. Km value is in good agreement with that determined using. E. coli cells expressing the entire Tomo ...
  28. [28]
    Bacterial Metabolism - Medical Microbiology - NCBI Bookshelf - NIH
    In aerobic respiration, molecular O2serves as the terminal acceptor of electrons. ... In respiration, the electron transfer reaction is the primary mode of ...Missing: biodegradation O2 cometabolism
  29. [29]
    Cometabolic Bioremediation
    Under aerobic conditions, the contaminant is oxidized by an enzyme or cofactor produced during microbial metabolism of another compound with oxygen. Under ...Missing: O2 terminal
  30. [30]
    Metabolism of toluene and xylenes by Pseudomonas (putida (arvilla ...
    These compounds are degraded by oxidation of one of the methyl substituents via the corresponding alcohols and aldehydes to benzoate and m- and p-toluates, ...
  31. [31]
    Aerobic and Anaerobic Toluene Degradation by a Newly Isolated ...
    On the other hand, Pseudomonas putida mt-2 oxidizes the methyl group of toluene to form benzoic acid, which is further metabolized through a meta cleavage ...
  32. [32]
    [PDF] Chapter XII Enhanced Aerobic Bioremediation
    aerobic bioremediation approaches follows the stoichiometric relationships shown ... Aerobic Biodegradation. Successful oxygen delivery and distribution is.
  33. [33]
    Micro-aerobic conditions based on membrane-covered improves the ...
    Haug (1993) showed that an oxygen concentration of <5 % during composting could affect the biodegradation rate of aerobic microorganisms; while an oxygen ...
  34. [34]
    Biodegradation - Hydrocarbons - Enviro Wiki
    Apr 1, 2024 · Hydrocarbon degradation under anaerobic conditions is often slower compared to aerobic degradation, due to less favorable reaction energetics ...<|separator|>
  35. [35]
    Atrazine biodegradation efficiency, metabolite detection, and trzD ...
    However, atrazine can be degraded mainly by biological processes N-dealkylation, dechlorination, and ring cleavage (Struthers et al., 1998). Atrazine can be ...
  36. [36]
    Substrate Specificity of Naphthalene Dioxygenase - NIH
    The three-component naphthalene dioxygenase (NDO) enzyme system carries out the first step in the aerobic degradation of naphthalene by Pseudomonas sp.
  37. [37]
    New Frontiers of Anaerobic Hydrocarbon Biodegradation in the Multi ...
    Our objective is to review the anaerobic hydrocarbon biodegradation processes, the most important hydrocarbon degraders and their diverse metabolic pathways.
  38. [38]
    Degradation of BTEX by anaerobic bacteria: physiology and ...
    Sep 7, 2010 · Toluene can be biodegraded with nitrate, Mn(IV), Fe(III), sulfate or CO2 as terminal electron acceptors (Langenhoff et al. 1997b; Chakraborty ...
  39. [39]
    Microbial interspecies interactions: recent findings in syntrophic ...
    May 13, 2015 · This syntrophic interaction is based on the transfer of reducing equivalents, such as hydrogen and formate, between these microbes and is also ...Missing: biodegradation | Show results with:biodegradation
  40. [40]
    Reductive Dechlorination of Chlorinated Ethenes and 1,2 ... - NIH
    “Dehalococcoides ethenogenes” 195 can reductively dechlorinate tetrachloroethene (PCE) completely to ethene (ETH). When PCE-grown strain 195 was transferred ...
  41. [41]
    Assessing in situ rates of anaerobic hydrocarbon bioremediation
    Deuterated metabolites formed in one aquifer at rates that ranged from 3 to 50 µg l−1 day−1, while the comparable rates at another aquifer were slower and ...
  42. [42]
    Anaerobic biodegradation of hazardous organics in groundwater ...
    Nov 1, 1994 · Sanitary landfills constructed without engineered liners release leachate into the subsurface, resulting in groundwater contamination. The ...
  43. [43]
    Anaerobic Benzene Degradation in Petroleum-Contaminated ... - NIH
    The finding that supplementing aquifer sediments with benzene-oxidizing sulfate reducers can greatly accelerate anaerobic benzene degradation not only provides ...
  44. [44]
    A biochemical framework for anaerobic oxidation of methane driven ...
    Apr 24, 2018 · The anaerobic oxidation of methane (AOM) requires reduction of electron acceptors (Fe(III), Mn(IV), nitrate or sulfate) to be thermodynamically ...
  45. [45]
  46. [46]
    [PDF] eco-ssl_pah.pdf - Environmental Protection Agency (EPA)
    Jun 5, 2007 · This document provides the Eco-SSL values for polycyclic aromatic hydrocarbons (PAHs) and the documentation for their derivation. This ...Missing: typical | Show results with:typical
  47. [47]
    [PDF] 4. CHEMICAL AND PHYSICAL INFORMATION
    In general, PCBs are relatively insoluble in water, and the solubility decreases with increased chlorination (see Table 4-3). PCBs are also freely soluble in ...
  48. [48]
    [PDF] 5.0 CONTAMINANT FATE AND TRANSPORT - EPA
    These mechanisms include diffusion, advection, mechanical dispersion, and hydrodynamic dispersion. Diffusion is the process by which a contaminant in water ...
  49. [49]
    Solid-phase microextraction to predict bioavailability and ... - PubMed
    Solid-phase microextraction to predict bioavailability and accumulation of organic micropollutants in terrestrial organisms after exposure to a field- ...
  50. [50]
    [PDF] UNDERSTANDING VARIATION IN PARTITION COEFFICIENT, Kd ...
    Thus the Kd has units of volume/mass, such as typically ml/g. Investigators have often found that Kd values measured for many contaminants for a given soil-.
  51. [51]
    Bacterial Chemotaxis toward Environmental Pollutants: Role in ... - NIH
    Bacterial chemotaxis enhances naphthalene degradation in a heterogeneous aqueous system. ... Summary review of health effects associated with naphthalene.
  52. [52]
    Speed-dependent bacterial surface swimming - ASM Journals
    May 8, 2024 · Moreover, P. aeruginosa cells swim much faster than E. coli cells, with speeds of approximately 40 µm/s and 20 µm/s, respectively (40).
  53. [53]
    Requirement of ATP in bacterial chemotaxis.
    Dec 28, 1981 · Evidence is presented that chemotaxis requires ATP or a closely related metabolite, in addition to its known.
  54. [54]
    Spatial structure, chemotaxis and quorum sensing shape bacterial ...
    Jan 2, 2024 · Our findings demonstrate that microscale medium structure and complex flow coupled with bacterial quorum sensing and chemotaxis control the heterogenous ...
  55. [55]
    (PDF) Chemotaxis of Burkholderia sp. Strain SJ98 towards ...
    Aug 10, 2025 · Chemotaxis against polycyclic aromatic hydrocarbons (e.g., pyrene, naphthalene, anthracene, and phenanthrenes), polychlorinated biphenyls ...
  56. [56]
    [PDF] The ecological roles of bacterial chemotaxis | The Stocker Lab
    In the absence of growth, the energy expenditure of the cell reduces to its maintenance level, estimated59,60 to be. ~104 ATP s−1 for E. coli and species with ...
  57. [57]
    Drift velocity of bacterial chemotaxis in dynamic chemical ... - Journals
    Sep 11, 2025 · v d = 𝜒 ∇ c , (1.1). where 𝜒 is the chemotactic sensitivity parameter and c is the concentration of an attractant or repellent ...
  58. [58]
    Hydrocarbon-Degrading Bacteria Alcanivorax and Marinobacter ...
    Oct 28, 2020 · Marine hydrocarbon-degrading bacteria play an important role in natural petroleum biodegradation processes and were initially associated with ...Missing: Oceanospirillum oxidation
  59. [59]
    Hydrocarbonoclastic Alcanivorax Isolates Exhibit Different ... - Frontiers
    The aldehyde is then converted to fatty acid, which can be degraded to acetyl-CoA through β-oxidation, constitute cellular membranes and exopolymeric ...Missing: Oceanospirillum | Show results with:Oceanospirillum
  60. [60]
    Hydrocarbon-degrading bacteria enriched by the Deepwater ... - NIH
    Jun 20, 2013 · We identified several aliphatic (Alcanivorax, Marinobacter)- and polycyclic aromatic hydrocarbon (Alteromonas, Cycloclasticus, Colwellia)-degrading bacteria.
  61. [61]
    Biodegradation of high molecular weight hydrocarbons under saline ...
    Aug 2, 2022 · This study used heavy crude oil to enrich petroleum-degrading bacteria from oil-contaminated saline soils.
  62. [62]
  63. [63]
  64. [64]
  65. [65]
    Microbial and Enzymatic Degradation of Synthetic Plastics - Frontiers
    PS are extremely stable polymers with high molecular weight and strong hydrophobic character, which makes these polymers highly resistant to biodegradation ( ...
  66. [66]
    Degradation Rates of Plastics in the Environment - ACS Publications
    Feb 3, 2020 · This Perspective summarizes the existing literature on environmental degradation rates and pathways for the major types of thermoplastic polymers.
  67. [67]
    Review Polyethylene terephthalate degradation under natural and ...
    Aug 5, 2020 · In the natural environment, PET has shown extremely slow degradation rates, resulting in bio-accumulation for decades if it escapes waste ...
  68. [68]
    A bacterium that degrades and assimilates poly(ethylene ... - Science
    Mar 11, 2016 · The new species, Ideonella sakaiensis, breaks down the plastic by using two enzymes to hydrolyze PET and a primary reaction intermediate.
  69. [69]
    The reaction mechanism of the Ideonella sakaiensis PETase enzyme
    Mar 27, 2024 · In 2016, Yoshida et al. reported the discovery of a soil bacterium, Ideonella sakaiensis 201-F6, that secretes a two-enzyme system for the ...
  70. [70]
    Polyurethane degradation by extracellular urethanase producing ...
    Jun 15, 2024 · The present study isolated polyurethane (PU) degrading bacterial species from soil dumped with plastic wastes. Four bacterial isolates, RS1, RS6 ...Missing: microbial biodegradation
  71. [71]
    Polyethylene Degradation by a Rhodococcous Strain Isolated from ...
    Microbial degradation of PE has been reported, but the underlying mechanisms are poorly understood. Here, we isolated a Rhodococcus strain A34 from 609 day ...Introduction · Results · Discussion · Supporting Information
  72. [72]
    Optimizing Eco-Friendly Degradation of Polyvinyl Chloride (PVC ...
    Oct 22, 2023 · Indeed, the environmental strain genus Aspergillus would be the perfect fungus for developing commercial levels of plastic biodegradation [49].
  73. [73]
    Microbial consortia for multi-plastic waste biodegradation: Selection ...
    This study evaluates the effectiveness of microbial consortia in degrading various types of plastics, including low-density polyethylene (LDPE), low linear- ...
  74. [74]
    Machine learning-aided engineering of hydrolases for PET ... - Nature
    Apr 27, 2022 · We demonstrate that untreated, postconsumer-PET from 51 different thermoformed products can all be almost completely degraded by FAST-PETase in 1 week.Missing: original | Show results with:original
  75. [75]
    Bio-augmentation - Effective Method of Treating Plastic Waste
    So, increase in bacterial load had a correlation with the degradation of polymer. The maximum degradation was found to be 8.8% within a period of 3 months ...
  76. [76]
    Biological Degradation of Plastics and Microplastics - NIH
    Jun 26, 2023 · Microbial enzymatic degradation will be based on the degradation ability of bacterial enzymes and appropriate biodegradation efficiency [14].
  77. [77]
    Respirometry tests in wastewater treatment: Why and how? A critical ...
    Nov 1, 2021 · Respirometry methods aim to evaluate the rapidly and slowly biodegradable COD (Chemical Oxygen Demand) fractions in wastewater streams, ...Missing: O2 flux
  78. [78]
    Respirometry tests in wastewater treatment: Why and how? A critical ...
    Aug 7, 2025 · Since oxygen consumption is an key indicator of the aerobic microbial activity, respirometry is often used to measure the oxygen uptake rate ...Missing: flux | Show results with:flux
  79. [79]
    Estimating the Removal and Biodegradation Potential of ... - PubMed
    A two-step procedure is described to characterize the removal and biodegradation potential of nonvolatile 14C-labeled organic compounds in activated sludge.Missing: microbial | Show results with:microbial
  80. [80]
    Radiolabeling for polymers degradation studies - ScienceDirect.com
    The radiolabeling technique (14C) allows the quantification of the degradation process of biopolymers by easily being able to distinguish between CO2 which is ...
  81. [81]
    [PDF] Evaluation of the Efficiency of Surface Aerator in the Activated ...
    ... COD. Especially in the. Activated Sludge Process (#ASP), the BOD removal is up to 80-95 percent and bacteria removal is up to 90-95%. The type of suitable ...
  82. [82]
    [PDF] Biodegradation Default Half-Life Values in the Light of ... - ECETOC
    Based on 'real world' half-life data for readily biodegradable substances, an overall half-life as low as 1.95 days can be calculated.
  83. [83]
    US7879593B2 - Fermentation systems, methods and apparatus
    The inoculum provided at the start of a fermentation batch is optimally about 10 6-10 7 cfu/ml, and may be in the range of about 10 3 to 10 8 cfu/ml. By ...
  84. [84]
    [PDF] epa-625-r-92-013.pdf
    Microbial characterization of municipal wastewater at a spray irrigation site: The Lubbock infection surveillance study. J. WPCF. 60(7):1222-1230. Moore ...
  85. [85]
    Degradation Kinetics Equations | US EPA
    Feb 6, 2025 · EPA and PMRA have developed a consistent approach for evaluating and calculating degradation kinetics in environmental media.
  86. [86]
    Determination of soil biodegradation half-lives from simulation ...
    A kinetic model approach for determination of biodegradation half-lives from soil simulation testing is presented.
  87. [87]
    Bioavailability of Heavy Metals in Soil: Impact on Microbial ... - MDPI
    In this review, we discuss the problems associated with the degradation of chlorinated organics in co-contaminated environments, owing to metal toxicity.2. Metal Toxicity And... · 3. Metal Speciation And... · References And Notes
  88. [88]
    Abundance and diversity of n-alkane and PAH-degrading bacteria ...
    Mar 1, 2022 · Quantifying functional genes from broad diversity of bacteria for monitoring oil pollution at sea. qPCR revealed limited reliability of degenerate primers in ...
  89. [89]
    Abundance and diversity of n-alkane and PAH-degrading bacteria ...
    Dec 9, 2021 · The amplicons generated through the qPCR assays were sequenced to reveal the diversity of oil-degradation related genes captured by the ...
  90. [90]
    PlastiCRISPR: Genome Editing-Based Plastic Waste Management ...
    May 8, 2025 · Similarly, in E. coli, CRISPR has been used to create fusion enzymes that combine PETase with MHETase or laccases, improving substrate ...
  91. [91]
    Development of a versatile synthetic microbe for the concurrent ...
    May 23, 2025 · We proposed the development of a synthetic strain, through synthetic biology approaches, capable of degrading multiple pollutants in saline environments.
  92. [92]
    Immobilized laccase mediated dye decolorization and ...
    Apr 25, 2015 · In batch decolorization, 90-95% decolorization was achieved of the simulated dye effluent for up to 10–20 cycles. Continuous decolorization in a ...Missing: turnover | Show results with:turnover
  93. [93]
    Enhanced rhizoremediation of polychlorinated biphenyls by ...
    ... PCBs degradation, which were 1.1–3.2 times faster than control microcosms. Although soil-inoculated LS1 maintained the PCB-degrading activity, higher PCBs ...
  94. [94]
    [PDF] Novozymes Water & Waste Management
    Apr 12, 2021 · BioRemove™ HC is the most effective biological product for hydrocarbon reduction. It is a blend of microorganisms developed for petroleum ...
  95. [95]
    Metabolic modeling of synthetic microbial communities for ...
    May 31, 2023 · Here, we review different types of microbial interactions and their research progress in environmental remediation.
  96. [96]
    Biodecomposition with Phanerochaete chrysosporium: A review - NIH
    Nov 25, 2024 · Because of their unique ability to break down lignin and aromatic compounds, fungi, especially white rot fungi (WRF), are used in the treatment ...
  97. [97]
    A Novel Extracellular Multicopper Oxidase from Phanerochaete ...
    Lignin degradation by the white rot basidiomycete Phanerochaete chrysosporium involves various extracellular oxidative enzymes, including lignin peroxidase, ...
  98. [98]
    Biodegradation of Polyester Polyurethane by Endophytic Fungi
    Several organisms demonstrated the ability to efficiently degrade PUR in both solid and liquid suspensions. Particularly robust activity was observed among ...
  99. [99]
    Microbial degradation of lignin: Role of lignin peroxidase ... - NIH
    These results indicated that the lignin polymer is really degraded by the LiP of white-rot fungi. ... White rot fungi constitutively produce laccase during ...Manganese Peroxidase (mnp) · Laccase (la) · 4. Aromatic Ring Cleavage<|separator|>
  100. [100]
    Mechanistic Features of Lignin Peroxidase-catalyzed Oxidation of ...
    ... pH. Oxidation of phenolate leads to formation of the corresponding phenoxyl radical (Equation 6). P h O − → P h O · + e. (Eq. 6). Redox potentials ...
  101. [101]
    Lignin Degradation by Fungal Pretreatment: A Review - ResearchGate
    Mar 11, 2017 · The degradation rate of lignin in the fermented tobacco stalk was 37.70, 51.56 and 53.75% with Trametes versicolor, Trametes hirsute and ...<|separator|>
  102. [102]
    Evaluation of the Synergistic Effect of Mixed Cultures of White-Rot ...
    Jul 28, 2017 · DDT was degraded to approximately 19% by P. ostreatus during the 7-day incubation period. The principal result of this study was that the ...
  103. [103]
    Biodegradation of polyester polyurethane by endophytic fungi
    Two Pestalotiopsis microspora isolates were uniquely able to grow on PUR as the sole carbon source under both aerobic and anaerobic conditions. Molecular ...
  104. [104]
    Fungal bioremediation: An overview of the mechanisms ...
    Fungal species such as Phanerochaete chrysosporium and Pleurotus ostreatus play a crucial role in degrading hydrocarbons found in contaminated soils and waters.
  105. [105]
    Fine roots and mycorrhizal fungi accelerate leaf litter decomposition ...
    Dec 20, 2020 · Mycorrhizal fungi may accelerate decomposition directly, through the exudation of enzymes, or indirectly by stimulating the free-living ...