2 Microbial Metabolism Yves Comeau 2.1 INTRODUCTION Wastewater originates from residences, institutions, offices and industries, and can be diluted with storm water, groundwater and surface water. Not treating wastewater before its discharge into receiving water bodies results in environmental and human health effects such as the generation of odours, the depletion of dissolved oxygen and the release of nutrients, toxic contaminants and pathogens. biodegradable or non-biodegradable organic matter, some of which can be toxic, and nutrients including the macronutrients nitrogen and phosphorus. While source reduction of contaminants should be encouraged, wastewater treatment by physical, chemical or biological processes remains necessary to minimize the potential impacts of wastewater discharge and to favour the production of valuable end-products such as reusable water, nutrients and biosolids. Wastewater treatment can be achieved by combining a variety of physical (e.g. screening, settling, filtration), chemical (e.g. coagulation, oxidation), thermal (e.g. drying, incineration) and biological (e.g. by suspended or attached biomass) processes. Biological wastewater treatment is based on the natural role of bacteria to close the elemental cycles (e.g. C, N, P) on earth. In a wastewater treatment plant, naturally occurring bacteria are used. By engineering the system, natural limitations for bioconversion such as limited aeration and limited amount of biomass can be overcome. Furthermore, the design of biological processes is based on the creation and exploitation of ecological niches that select for microorganisms best adapted to reproduce under such environmental conditions. Selective pressure may arise from various conditions of availability of electron donor (most often organic matter), electron acceptor (such as oxygen or nitrate), nutrients, pH, temperature, hydrodynamic (washing out non-attached microorganisms) or other conditions. Biological wastewater treatment, the central focus of this book, aims at degrading or adsorbing dissolved, colloidal, particulate and settleable matter into biological flocs or biofilms. Soluble compounds include In this chapter, elements of microbiology are first reviewed to better understand the needs and functions of microorganisms, and then the stochiometry, energetics and kinetics of microbial growth are presented. © 2008 Yves Comeau. Biological Wastewater Treatment: Principles, Modelling and Design. Edited by M. Henze, M.C.M. van Loosdrecht, G.A. Ekama and D. Brdjanovic. ISBN: 9781843391883. Published by IWA Publishing, London, UK. 10 2.2 Biological Wastewater Treatment: Principles, Modelling and Design ELEMENTS OF MICROBIOLOGY Considering the dominant role of bacteria in wastewater treatment, their relationship to other living organisms is first presented followed by their cell structure and components, functions, nutritional requirements, carbon and energy sources, and sensitivity to environmental conditions. 2.2.1 noticeable portion of them, pathogenic to humans. Wastewater pathogens are found among each class of microorganisms from viruses (e.g. Hepatitis A virus causing hepatitis), to bacteria (e.g. Vibrio cholerae causing cholera), to protozoa (e.g. Giardia lamblia causing giardiasis) and even to animals such as helminth worms (e.g. Taenia saginata causing taeniasis). A concise description of pathogenic microorganisms can be found in Chapter 8. Classification of microorganisms There are two types of organisms, prokaryotes and eukaryotes (Figure 2.1). Prokaryotes are mostly unicellular organisms which include bacteria, cyanobacteria (blue-green algae) and archaea (some found in extreme environments) while eukaryotes include unicellular organisms (protozoa, algae, fungi) and multicellular ones (fungi, plants, animals). Recent genetic information has allowed grouping of organisms according to their common evolutionary origins. Coccus Organisms found in wastewater and wastewater treatment plants include mainly microorganisms (viruses, bacteria, protozoa) and some higher organisms (algae, plants, animals). The morphology of various groups of microorganisms which are found in wastewaters and can be observed by microscopy is shown in Figures 2.2 to 2.5. Sarcina (packets of eight) Streptococci Staphylococci Bacillus Chains of bacilli Spirillum Microorganisms are the catalysts of biological wastewater treatment and, for a very small but te Prokaryo Figure 2.2 Morphology of bacteria (adapted from Rittmann and McCarty, 2001) s Archaea Methanogens Extreme halophiles Bacteria Hyperthermophiles Eukarya Gram-positive bacteria Proteobacteria Animals Mitochondrion Slime molds Fungi Plants Cyanobacteria Flagellates Chloroplast Giardia Hyperthermophiles Root of the tree Figure 2.1 Phylogenetic tree of life (adapted from Madigan and Martinko, 2006) Euka ryote s Eukaryotic ‘Crown species’ 11 Microbial Metabolism Zooflagallate Ameba A Paramecium Ciliate B Vorticella Figure 2.3 Morphology of protozoa (adapted from Rittmann and McCarty, 2001) Figure 2.5 Activated sludge floc with good settling properties (A) and with excessive filamentous growth (B) (photos: D. Brdjanovic; Eikelboom, 2000; respectively) 2.2.2 Nematode Rotifers Cell structure and components The structure of prokaryotes and of eukaryotes is presented in Figure 2.6. A Crustacea B Chromatin Endoplasmic Nucleolus reticulum Nucleus (surrounded by nuclear membrane) Nucleoid DNA Cytoplasm includes: RNA, volutin granules Plasmid (polyphosphates, sulfur), storage products (glycogen, lipids) Cell wall and membrame Microtubules Figure 2.4 Morphology of multicellular microorganisms (adapted from Rittmann and McCarty,2001) Flagellum Smooth endoplasmic reticulum Lysosome Mitochondrion Centrioles Cilium Figure 2.6 Structure of (A) prokaryotic (0.5 to 5 microns) and (B) eukaryotic (5 to 100 microns) cells (adapted from Metcalf & Eddy, 2003) 12 Biological Wastewater Treatment: Principles, Modelling and Design One essential difference between these types of organisms is that the genetic material (deoxyribonucleic acid, DNA) is found as a nucleoid in prokaryotes but in a true nucleus surrounded by a membrane in eukaryotes (the greek word karyon means nucleus). Bacteria may also contain extra DNA material as shorter chain plasmids. Energy, in eukaryotes, is mainly generated in mitochondrion. In prokaryotes, the cytoplamic membrane surrounding the cell fluid (or cytoplasm) creates a separation between the intracellular and the extracellular environment which limits the passage of dissolved components and allows the creation of both a pH (more H+ outside) and a charge gradient (more positive charges outside) which is used as a major mechanism to generate energy and to transport metabolites. Internally cells maintain a relatively constant composition. Bacterial polymer compounds of significance in wastewater treatment include poly-β-hydroxyalkanoates (PHAs), glycogen and polyphosphates (Figures 2.8 to 2.10). These compounds play a role as energy reserves as well as organic carbon (PHA, glycogen) and phosphorus (polyphosphates) reserves. A ... CH2 O CH3 O CH3 O CH3 C CH C CH C CH O CH2 O CH2 O CH2 ... beta carbon B Bacterial macromolecules include proteins, nucleic acids (DNA and RNA: ribonucleic acids), polysaccharides and lipids. These compounds are found in various locations in bacteria (Figure 2.7). Flagellum Cytoplasmic membrame Cell wall Cytoplasm A Proteins Nucleoid Ribosomes Figure 2.8 (A) Structure of poly-β-hydroxybutyrate (PHB). In poly-β-hydroxyvalerate (PHV), the -CH3 group is replaced by -CH2CH3 group. PHB and PHV are the two most common poly-βhydroxyalcanoates (PHAs). (B) White granules of PHA stored inside the cells (cell size approximately 1 μm) (photo: M.C.M. van Loosdrecht) B Nucleic Acids: DNA RNA C Polysaccharides Storage granules D Lipids Figure 2.7 Bacterial macromolecules and location in the cell. (A) Proteins are found in the flagellum, the cytoplasmic membrane, the cell wall and the cytoplasm; (B) nucleic acids (DNA and RNA) are found in the nucleoid and ribosomes; (C) polysaccharides are found in the cell wall and sometimes in storage granules and (D) lipids are found in the cytoplasmic membrane, the cell wall and in storage granules (adapted from Madigan and Martinko, 2006) Starch, glycogen and cellulose are all polymers of glucose that differ in the type of glycosidic bond between molecules (Figure 2.9). Changing the linkage or geometry of this bond results in polymers that vastly differ in their strength. Cellulose is the strongest polymer and is used as a structural material in plants and trees. It is also the most difficult of these polymers to biodegrade. Polyphosphates are linear chains of phosphates whose negative charge is stabilised by cations. The energy-rich phosphate-ester bond is the same as in the universal energy carrier molecule inside the cell, adenosine triphosphate (ATP) which contains a chain of 13 Microbial Metabolism 3 phosphates. In most bacteria polyphosphate is used as a phosphate reserve and only a limited group of bacteria use it as an energy storage compound. A 6 CH2OH H 5 H 4 OH OH 3 H CH2OH O O H H 2 H 1 H H OH H OH O OH H OH α-1,4-Glycosidic bond CH2 CH2OH O O H H 1 H O 2 H H OH H 2 OH OH 1 O For growth to take place, bacteria must be able to replicate their genetic material and carry-out chemical transformations which allow the synthesis of all the constituents from various precursors and energy (Figure 2.11). Chemical transformations are catalyzed by enzymes which are proteins. The synthesis of any protein requires its genetic expression. The first step is the transcription of DNA (a double strand of nucleic acids) into RNA (a single strand of nucleic acids), followed by its translation into a protein that is then processed to render it functional. With its constituents replicated, a bacterial cell can then divide into two daughter cells. H Coding functions β-1,4-Glycosidic bond α-1,6-Glycosidic bond Functions of bacteria H OH H H OH 2.2.3 Machine functions DNA 1. Energy: ADP +Pi B Replecation Starch α-1,4 bonds Gene expression Transcription ATP 2. Metabolism: generation of precursors of macromolecules (sugars, amino acids, fatty acids, etc.) 3. Enzymes: metabolic catalysts RNA α-1,6 bonds Translation α-1,4 bonds Glycogen Protein β -1,4 bonds Cellulose Figure 2.9 Structure of the polysaccharides. (A) Differences in the glycosidic bonds in the position of linkage between glucose molecules and in geometry (α and β). (B) Structure of starch, glycogen (a bacterial storage polymer) and cellulose (adapted from Madigan and Martinko, 2006) O ... O P O O O P O O O P O ... O Figure 2.10 Structure of polyphosphates. Polyphosphates are polymers of phosphate molecules and are stabilised by cations (e.g. Ca2+, Mg2+, K+) interacting with charged oxygen (-O-) molecule Reproduction (growth) Figure 2.11 Functions of cells. Growth requires both coding and machine functions to be operational. DNA serves for replication and gene expression, first by transcription of DNA into RNA, then translation of RNA into proteins. Note: DNA: deoxyribonucleic acid; RNA: ribonucleic acid (adapted from Madigan and Martinko, 2006) 2.2.4 Characterization of bacteria Microorganisms can be characterized by first isolating single strains from microbial communities with successive dilutions and enrichment cultures and then testing their response to various conditions. More recently, molecular tools have been developed which allow the study of microorganisms without having to isolate and cultivate them. 14 Biological Wastewater Treatment: Principles, Modelling and Design Unique abilities of bacteria to produce a given protein, and store its genetic code, can be used to detect their presence in biological samples. The potential for the expression of a protein is thus given by confirming the presence of the gene in the DNA, while its actual expression would be given by confirming the presence of the associated RNA in the biomass. 2.2.4.1 Fluorescent in situ hybridization (FISH) Fluorescent in situ hybridization (FISH) consists of chemically preparing a short strand of the specific sequence of nucleic acids, an oligonucleotide, and appending a coloured fluorescent marker at its end. Cells are then made porous to the marked oligonucleotide which binds to its complementary strand of RNA. After removing the unbound markers, bacteria containing the target genetic material emit light which can be observed under a fluorescent microscope (Figure 2.12) confirm its presence. In the polymerase chain reaction, three components are added: a high temperature resistant polymerase enzyme, "flanking" oligonucleotides that delimit the extremities of the target gene, and nucleic acids so that copies of the target gene can be made. A temperature cycle is imposed which results in the opening (denaturation) of the DNA and annealing with the added oligonucleotides. The polymerase enzyme then completes the replication of the gene between the two flanking oligonucleotides. As this cycle is repeated the number of copies of the target gene increases exponentially, facilitating its detection. Instead of aiming for only one DNA sequence, many gene sequences can be amplified at once and the fragments of the various amplified genes can be detected by DGGE. In DGGE, an electric current is applied to a gel containing an increasing concentration (gradient) of denaturant. As the various PCR-amplified DNA gene sequences migrate, they start opening, being denatured, which slows down their migration in the gel and yields various bands which are characteristic of the target genes of specific microorganisms. 2.2.5 Bacterial bioenergetics The energy needed for the metabolism of bacteria is obtained from chemical oxidation reduction reactions. Two main pathways of energy generation are the glycolysis and the tricarboxylic acid cycle (TCA; or citric acid cycle or Krebs cycle) in which glucose (a sugar) is degraded into pyruvate and to acetylCoA (AcCoA) which feeds into the TCA cycle (Figure 2.13). Figure 2.12 FISH image of a nitrifying sludge granule. Ammonium oxidising Beta proteobacteria (probe NSO 190): green; Nitrospira-like organism (probe NTSPA 662): red; Eubacteria (probe EUB 338): blue (Eubacteria). Bar indicates 20 μm. (photo: from Kampschreur, 2008) 2.2.4.2 Polymerase chain reaction (PCR) and denaturing gradient gel electrophoresis (DGGE) PCR is used to amplify the number of a specific gene in the DNA. The DNA first needs to be extracted from a biological sample, amplified (multiplied) by a polymerase chain reaction and then identified to Chemical energy is transferred to the energy-rich compound adenosine triphosphate (ATP) and electrons are transferred to the oxidized form of the coenzyme nicotinamide dinucleotide (NAD+) that becomes reduced to NADH. In the presence of an electron acceptor such as oxygen (O2) or oxidized nitrogen (NOx: nitrate, NO3- or nitrite, NO2-), the NADH can transfer electrons via the electron transport chain (E.T.C.) to the electron acceptor. In this electron transport process, protons are transported across the cell membrane to the outside of the cell. The pH and charge gradient thus create a proton motive force (p.m.f.) which is used for the transport of various compounds across the cell membrane and for ATP production by the ATP-ase enzyme. During this transport and ATP generation, protons are transported back to the cell interior. Some toxic chemical compounds such as dinitrophenol (DNP) can neutralize the proton gradient 15 Microbial Metabolism across the membrane and are called uncouplers as they "uncouple" organic carbon consumption and ATP production. Thus, there are three key central metabolites in bacterial bioenergetics, acetylCoA, ATP and NADH. The intracellular level of these compounds acts as a powerful regulator of the metabolism of bacteria. In the absence of an external electron acceptor, the cell cannot regenerate the NADH produced by glycolysis. Under these conditions the TCA cycle will not function to oxidise the substrate further than pyruvate and acetylCoA. By conducting fermentation, however, pyruvate can be reduced with the NADH generated in the glycolysis into products such as acetate and propionate. Bacteria sugar acetate ATP glycolysis NADH pyruvate fermentation AcCoA TCA cycle H+ H+ DNP NADH E.T.C H+ H+ NADH H+ NADH CO2 H+ (e.g.:N2) H2O ATP ADP + Pi NADHtranshydrogenase H+ ATP-ase (reversible) H+ (NOx) O2 – – + + – - + OH- OH charge gradient + H H+ pH gradient pmf Figure 2.13 Overview of bacterial bioenergetics (adapted from Comeau et al., 1986) 2.2.6 Nutritional requirements for microbial growth In addition to energy, microorganisms require sources of carbon and inorganic compounds to synthesize cellular components. Bacteria found in wastewater treatment plants are typically composed of 75-80% water and thus, of 20-25% dry matter. The dry matter content is determined from a liquid sample of known volume by retaining biomass on a glass fiber filter having a nominal pore sizes of about 1.2 micron and evaporating the water to dryness in an oven heated at 105ºC. After cooling, the dried biomass is weighed on an analytical balance and the results expressed as total suspended solids (TSS) in g/m3 (mg/l). The dried glass fibre filter that retained the biomass can then be combusted at 550ºC in a muffle furnace to burn the organic matter (considered to be composed of C, H, O and N). The ash remaining is considered to represent the inorganic components and is termed ash or fixed suspended solids (FSS). By difference, the organic matter is calculated which is termed volatile suspended solids (VSS). The typical composition of the dry matter (TSS) of bacteria is presented in Table 2.1. Table 2.1 Typical composition of bacteria (adapted from Metcalf & Eddy 2003) Constituent or element Major cellular constituents Protein Polysaccharides Lipid DNA RNA Other (sugars, amino acids) Inorganic ions As cell elements Organic (VSS) Carbon Oxygen Nitrogen Hydrogen Inorganics (FSS) Phosphorus Sulfur Potassium Sodium Calcium Magnesium Chlorine Iron Other trace elements %TSS Empirical formula for cells C5H7O2N 55.0 5.0 9.1 3.1 20.5 6.3 1.0 %VSS 93.0 50.0 22.0 12.0 9.0 7.0 2.0 1.0 1.0 1.0 0.5 0.5 0.5 0.2 0.3 53.1 28.3 12.4 6.2 The organic (VSS) and inorganic content of bacteria are thus about 93% and 7%, respectively. Not only should macro nutrients such as nitrogen and phosphorus need to be present for cell growth but other elements are also essential. These compounds are rarely missing in municipal effluents but may be lacking in some industrial effluents such as from sugar or pulp and paper industries. Empirical formulae proposed for cells (active biomass) found in wastewater treatment processes are 16 Biological Wastewater Treatment: Principles, Modelling and Design C5H7O2N and C60H87O23N12P which can be approximated to C5H7O2NP1/12. These formulae give dry matter contents (%TSS) for C, H, O, N and P that are in relatively close agreement with the values presented in Table 2.1. Other trace elements required include Zn, Mn, Mo, Se, Co, Cu and Ni. 2.2.7 Carbon and energy sources and microbial diversity Metabolism is the sum of all chemical processes that take place in living cells (Figure 2.14). It is divided into two categories, catabolism and anabolism. Catabolic reactions are the energy supply of the cell. The catabolic reaction is a redox reaction where the transport of electrons from electron donor to electron acceptor is generating a proton motive force which delivers ATP. Anabolic reactions use this energy for the synthesis of cellular components from carbon sources and other nutrients. If organic carbon compounds are the substrate then they function as well in the catabolic as in the anabolic reactions. The anabolic processes are more or less the same in all bacteria, while the catabolic processes can vary widely between different microbial groups. Energy production requires the presence of an electron donor and an electron acceptor. A reduced compound acts as the electron donor (e.g. organic matter or ammonium) while an oxidized compound acts as the electron acceptor (e.g. oxygen or nitrate). The minimum and maximum oxidation states, with an example of a corresponding molecule, are shown in Table 2.2 for significant elements in microbiology. Carbon sources for biosynthesis are only of two types, organic or inorganic. The energy sources are of three types, organic, inorganic and from light, but the variety of combinations of electron donors and acceptors results in a broad diversity of microorganisms (Table 2.3). The name of these groups come from Greek roots: chemo: chemical; troph: nourishment; organo: organic; litho: inorganic; photo: light; auto: self; hetero: other. Chemotrophs obtain energy from the oxidation of electron donating molecules from their environment. These molecules can be organic (chemo-organotrophs or chemo-organoheterotrophs) or inorganic (chemolithotrophs or chemolithoautotrophs). Chemoorganotrophs are normally heterotrophs and chemolithotrophs are normally autotrophs with these names being used interchangeably. Not every microbial type is presented in this table. Other groups include dehalorespirers which use some types of chlorinated compounds as electron acceptors. Examples of microbial growth reactions with their principal function in wastewater treatment are given below. Neutral molecules are used for reactions even if other ionic species may be dominant. The Eq. 2.1 to 2.6 are given for illustration of metabolism only and are not balanced: • Aerobic heterotrophs: organic matter oxidation C6 H 12O6 + O2 + NH 3 + other nutrients → C5 H7 O2 N + CO2 + H 2O • Denitrifiers: nitrate removal C6 H 12O6 + O2 + HNO3 + NH 3 + other nutrients → C5 H7 O2 N + CO2 + H 2O + N 2 • (2.2) Fermenting organisms: conversion of larger organic compounds: glucose to acetic acid, Anabolism Catabolism (2.1) Energy-yielding metabolism Biosynthetic metabolism Energy sources Biopolymers ATP Biosynthetic intermediates Utilizable energy Heat ADP Intracellular precursor pool Metabolic products External nutrients Figure 2.14 Metabolism as the combination of catabolism and anabolism (adapted from Todar, 2007) 17 Microbial Metabolism C6 H 12O6 + O2 + NH 3 + other nutrients → (2.3) C5 H7 O2 N + CH 3CO2 H + CO2 • Aerobic autotrophic bacteria (ammonia oxidizers): removal of ammonia CO2 + NH 3 + O2 + other nutrients → (2.4) C5 H7 O2 N + HNO3 + H 2O • Hydrogenotrophic methanogens: biogas production H 2 + CO2 + NH 3 + other nutrients → C5 H7 O2 N + CH 4 • (2.5) Plants: O2 production and greenhouse gas reduction CO2 + light + NH 3 + other nutrients → C5 H7 O2 N + O2 (2.6) Table 2.2 Significant elements in microbiology Name and symbol Oxygen Nitrogen Carbon Sulfur Hydrogen Iron Manganese O N C S H Fe Mn Reference oxidation state (=0) and phase O2 (g) N2 (g) C (s) S (s) H2 (g) Fe (s) Mn (s) Electro-negativity (x) 3.50 3.07 2.50 2.44 2.10 1.64 1.60 Oxidation state and state of min x -II H2O -III NH4+ -IV CH4 -II HS0 H2 0 Fe II Mn2+ Oxidation state and state of max x 0 O2 V NO3IV HCO3VI SO42I H+ III Fe3+ IV Mn4+ Oxidation states shown: reference, min, max; phases shown are gas (g) and solid (s); Electro-negativity refers to an atom’s tendency to attract electrons (e-); at a high oxidation state, these elements (except H+) are potential electron acceptors for catabolic reactions (adapted from Heijnen et al., in preparation) Table 2.3 Trophic classification of microorganisms (adapted from Rittmann and McCarty, 2001; Metcalf & Eddy, 2003) Carbon source1 Energy source 2 Electron acceptor Typical products Electron donor Trophic group Microbial group Chemotroph Organotroph Aerobic heterotrophs Denitrifiers Fermenting organisms Iron reducers Sulfate reducers Methanogens (acetoclastic) Lithotroph Nitrifiers: AOB4 Nitrifiers: NOB5 Anammox6 bacteria Denitrifiers Denitrifiers Iron oxidizers Sulphate reducers Sulphate oxidizers Aerobic hydrogenotrophs Methanogens (hydrogenotrophic) Phototroph Algae, plants Photosynthetic bacteria 1 - Type of e donor Organic Organic Organic Organic Acetate Acetate NH4+ NO2NH4+ H2 S Fe (II) H2 H2S, S0,S2O32H2 H2 O2 NO3-, NO2Organic Fe (III) SO42acetate O2 O2 NO2NO3-, NO2NO3-, NO2O2 SO42O2 O2 CO2 CO2, H2O N2, CO2, H2O Organic:VFAs3 Fe (II) H2S CH4 NO2NO3N2 N2, H2O N2, SO42- ,H2O Fe (III) H2S, H2O SO42H2O CH4 Organic Organic Organic Organic Acetate Acetate CO2 CO2 CO2 CO2 CO2 CO2 CO2 CO2 CO2 CO2 H2O H2S CO2 CO2 O2 S (0) CO2 CO2 Carbon source: organic for heterotrophs and inorganic (CO2) for autotrophs; mixotrophs can use both. 2 Typical products: CO2 and H2O are products of catalysis (energy generation) by many micro-organisms. 3 VFAs: volatile fatty acids (typically acetate, propionate, butyrate). 4 AOB: ammonia oxidizing bacteria. 5 NOB: nitrite oxidizing bacteria. 6 Anammox: anaerobic ammonia oxidizing bacteria. 18 Biological Wastewater Treatment: Principles, Modelling and Design Table 2.4 Oxygen and microorganisms (adapted from Madigan and Martinko, 2006) Group Aerobes Obligate Facultative Microaerophilic Anaerobes Aerotolerant Obligate Type of metabolism Required (e.g. 20%) Better if present, not essential Requires low levels (e.g. 1%) Aerobic respiration Aerobic or nitrate respiration, fermentation Aerobic respiration Not required, not affected by its presence Fermentation or sulphate reduction Fermentation of anaerobic fermentation O2 harmful or lethal Environmental conditions (oxygen, temperature, toxicity) Environmental conditions must be favourable for microorganisms to grow. Major factors affecting growth are oxygen and temperature but pH (typically 6 to 8) and osmotic pressure (depends on the concentration of salts) must also be appropriate. 2.2.8.1 Oxygen The need, tolerance or sensitivity to molecular oxygen (O2) varies widely among micro-organisms (Table 2.4). Aerobes use oxygen and may need it (obligate), function in its absence (facultative) or require it in low levels (microaerophilic). Anaerobes do not use oxygen but may tolerate it (aerotolerant) or not (obligate). In aerobes, enzymes for oxygen reduction (to use O2 as an electron acceptor) are always induced. In contrast, denitrifiers which are facultative aerobes, also have constitutive enzymes for oxygen reduction but enzymes for nitrate (or nitrite) reduction need to be induced, a condition that requires the absence of oxygen. All denitrifying bacteria can also use oxygen, their catabolic processes being relatively similar. Sulphate reducers on the contrary cannot use oxygen, their catabolic process being very different from aerobic respiration. Table 2.5 Engineering definition of some environmental conditions Condition Electron acceptor Present Absent Aerobic OX O2 Anoxic AX NOx Anaerobic AN O2 O2 and NOx - - NOx refers to nitrate (NO3 ) plus nitrite (NO2 ) While the absence of oxygen is referred to as anoxic (without O2) or anaerobic (without air) by microbiologists, engineers make a distinction between these two conditions. Thus, in the absence of oxygen, the presence or absence of oxidized nitrogen (nitrate or nitrite) is referred to as anoxic and anaerobic conditions, respectively (Table 2.5). 2.2.8.2 Temperature Temperature has a significant effect on the growth rate of microorganisms (Figure 2.15). Hyperthermophile Thermophile Growth rate 2.2.8 Relationship to O2 Mesophile Psychrophile 0 20 40 60 80 100 120 Temperature (˚C) Figure 2.15 Effect of temperature on microbial growth rate (adapted from Rittmann and McCarty, 2001) Those operating at a higher temperature range have a higher maximum growth rate than those operating at a lower range. The optimal range of temperature for each group is relatively narrow. With an increasing temperature, a gradual increase in growth rate is observed until an abrupt drop is observed due to the denaturation of proteins at a higher temperature. The generally used terms to describe these microorganisms are psychrophile below about 15ºC, mesophile for 1540ºC, thermophile at 40-70ºC and hyperthermophile which are active above 70ºC up to around 110ºC. 19 Microbial Metabolism 2.3 STOICHIOMETRY AND ENERGETICS 2.3.1 Theoretical chemical oxygen demand (thCOD) and electron equivalents For example (Eq 2.10), the mineralisation of glucose gives, C6 H 12O6 + 6O2 → 6CO2 + 6 H 2O The chemical oxygen demand (COD) determination is commonly conducted in laboratories and involves the oxidation of organic compounds in the presence of an acidic dichromate solution heated at 150ºC for 2 hours. The number of electrons donated by dichromate in the test is expressed as oxygen equivalents in gO2/m3 (or mgO2/l). The electron equivalents of oxygen can be determined by noting that 1 mole of O2 weighs 32 g and contains 4 electron equivalents (2 O molecules • 2 e-/O molecule). Thus, 1 electron equivalent (eeq) corresponds to 8 g of COD (Eq. 2.7) 1 eeq = 8 gCOD (2.7) 180 g 192 g (2.10) Thus, 1 g glucose represents 1.067 g thCOD (192/180). Considering that 8 g of O2 corresponds to 1 eeq, 1 mol of glucose donates 24 eeq. Thus, removing O2 from the above equation, adding 24 electrons as products of the reaction, and as many protons (H+) for charge balance, and water for H balance gives the following half reaction equation (Eq. 2.11). C6 H12O6 + 6 H 2O → 6CO2 + 24e− + 24H + (2.11) For 1 eeq, Eq. 2.11 becomes: Considering that organic matter is an electron donor while O2 is an electron acceptor, dissolved O2 is considered to represent negative COD (Eq. 2.8). 1 1 1 C6 H 12O6 + H 2O → CO2 + e − + H + 24 4 4 1 gO2 = −1 gCOD A similar approach can be used for electron acceptors. For oxygen this gives: (2.8) 0.25O2 + H + + e− → 0.5H 2O (2.12) The theoretical chemical oxygen demand (thCOD) of a substrate can be determined by writing a balanced equation in which O2 is added and the compound is mineralised to end products with ammonia remaining in its NH3 (III) oxidation state. The theoretical COD may deviate from the measured COD when a compound is not reacting in the COD test. Summing the above two equations again gives the full reaction equation for glucose. Eq. 2.9 gives a generalised equation for this purpose. The equation refers to the thCOD of a C, H, N, O containing substrate (adapted from Rittmann and McCarty 2001). HNO3 + 5H + + 5e− → 0.5N 2 + 3H 2O 1 ( 2n + 0.5a − 1.5c − b ) O2 → 2 a − 3c nCO2 + cNH 3 + H 2O 2 (2.9) thCOD/weight ( 2n + 0.5a − 1.5c − b) 16 12n + a + 16b + 14c Similarly for the transformation of nitrate to nitrogen gas (denitrification), the oxidation state of nitrogen is reduced from +V to 0. (2.14) The COD equivalent of this reaction is 5 eeq/mol x 8 gCOD/eeq = 40 gCOD/molHNO3 = 2.86 gCOD/gNO3N. As electrons are accepted and not donated, the COD equivalent of 1 g of nitrate-nitrogen is thus minus 2.86 gCOD (-2.86 gCOD/gNO3-N) = 40/(14 g/mol). Cn H aOb N c + and (2.13) Writing equations with neutral or charged molecules does not change the number of electron equivalents of a reaction as the number of protons (H+) will be adjusted. 20 Biological Wastewater Treatment: Principles, Modelling and Design Table 2.6 Theoretical COD of various compounds by weight Compound Chemical formula Weight (VSS) CHON (g/mol) C/wt (%) N/wt (%) P/wt (%) 113 113 1343 131 393 960 86 53 52 52 55 55 50 56 12 12 12 11 4 10 0 0 2.2 2.3 0 0 3.1 0 160 160 1960 193 560 1369 144 1.42 1.42 1.46 1.48 1.42 1.43 1.67 184 393 282 134 254 320 180 46 60 74 88 16 2 52 55 43 72 85 53 40 26 40 49 55 75 - 15 4 0 0 0 9 0 0 0 0 0 0 - 0 0 0 0 0 0 0 0 0 0 0 0 - 256 560 320 272 880 384 192 16 64 112 160 64 16 1.39 1.42 1.13 2.03 3.46 1.20 1.07 0.35 1.07 1.51 1.82 4.00 8.00 thCOD COD/VSS (g/mol) (g/g) Biomass C5H7O2N C5H7O2NP1/12 C60H87O23N12P C6H7.7O2.3N C18H19O9N C41.3H64.6O18.8N7.04 C4H6O2 Organic substances Casein Average organics Carbohydrates Fats, oils Oils: oleic acid Proteins Glucose Formate Acetate Propionate Butyrate Methane Hydrogen C8H12O3N2 C18H19O9N C10H18O9 C8H6O2 C18H34O2 C14H12O7N2 C6H12O6 CH2O2 C2H4O2 C3H6O2 C4H8O2 CH4 H2 Lag Exponential Stationary phase growth phase phase For substrates, however, the thCOD/VSS ratio varies greatly according to the degree of reduction of the substrate. Ratios range between 0.35 for formate, a highly oxidized substrate, to 4.00 gram COD per gram substrate for methane, and to 8 gram COD per gram for hydrogen. An average municipal wastewater would have a typical COD to volatile solids (filtered plus particulate) of 1.2 gCOD/gVS. 2.3.2 Cell growth Cell growth in a batch test is characterized by four phases during which the substrate and biomass concentration evolve (Figure 2.16). Concentration The theoretical COD of a number of compounds is presented in Table 2.6. Various biomass equations give thCOD to dry weight ratios varying between 1.37 and 1.48 gCOD/gVSS, with 1.42 being considered typical for municipal biological wastewater treatment. Decay phase Substrate Biomass Time Figure 2.16 Biomass growth in batch mode (adapted from Metcalf & Eddy, 2003) The four phases are: (1) The lag phase during which there is little biomass increase and little substrate consumed as the cells acclimate to the new situation. 21 Microbial Metabolism (2) The exponential growth phase follows during which the biomass grows at its maximum rate consuming much of the substrate which is readily available. (3) The stationary phase is next during which little external substrate is available and the biomass concentration remains relatively constant. (4) Finally, the decay phase is associated with biomass decay due to the consumption of the internal carbon and energy reserves for its maintenance needs, and due to predation and lysis. These growth conditions may be found in wastewater treatment plants at start-up (lag phase), in highly loaded plants or the front part of plug flow process (exponential growth phase), in the mid and end section of a plug flow process (stationary phase) and in a facultative lagoon or aerobic sludge digester (decay phase). 2.3.3 Yield and energy 2.3.3.1 Energy from catabolism Microbial metabolism requires energy for cell synthesis. Depending on the electron acceptor and donor couple and the associated energy production, a varying proportion of the electrons available from the electron donor will be available for biomass synthesis. For example, aerobic oxidation of glucose generates much more energy than the transformation of glucose into methane explaining why the cell yield of the first reaction is greater than that of the second. Bioenergetics provides a tool to quantify the amount of energy available for various biological reactions which can then be used to determine the biomass yield of a reaction. Energy production by catabolism depends on the oxidation and reduction of chemicals available to microorganisms. In a given reaction, the electron donor (ED) is oxidized while the electron acceptor (EA) is reduced. The electron donor is considered to be the high energy substrate or "food" of the reaction and a large variety of compounds can play this role. The electron acceptor, conversely, is an oxidized form and a more limited number is available for biological systems (mainly oxygen, nitrate, nitrite, iron (III), sulfate, carbon dioxide). The change in Gibbs energy (ΔG0) is a useful thermodynamic property of a reaction which characterizes the maximum amount of energy (work) obtainable for a given reaction. The superscript indicates that the compounds involved are at standard conditions (1 mole, 1 atmosphere) and 25ºC. For biological processes often the standard Gibbs energy is given for pH 7, which is then denoted by adding a prime (’) to the symbol for the Gibbs energy. Some half reactions for biological systems and Gibbs energy changes per electron equivalent (ΔG0 kJ/eeq) are listed in Table 2.7. In combining electron donor and electron acceptor reactions it should be noted that all reactions in Table 2.7 are presented as electron acceptors with the electron on the left hand side. Thus, for an electron donor reaction, the reagents and products of the reaction should be exchanged and the sign of the Gibbs energy change should be changed. If the net reaction results in a negative ΔGo’, this means that energy can be released and the reaction can occur spontaneously, an exergonic reaction. Conversely, if the net reaction results in a positive ΔGo’, energy input would be needed for the reaction to take place and it will not occur spontaneously, an endergonic reaction. The energy available from the transformation of glucose (electron donor) by aerobic oxidation (with O2 as electron acceptor) and by methanogenesis (with carbon dioxide as electron acceptor) is illustrated in Table 2.8. These two oxidation reactions of glucose illustrate that aerobic metabolism provides nearly 7 times more energy than anaerobic methanogenesis. Consequently, the cell yield would be expected to be much higher with oxygen than with carbon dioxide as electron acceptors. Other biological reactions are illustrated on Figure 2.17. 2.3.3.2 Synthesis fraction and biomass yield A portion of the electron-donor substrate is used for cell synthesis (fs0: true synthesis fraction) and the rest for energy production (fe0: true energy fraction) (Figure 2.18). On an electron equivalent (eeq) basis, the sum of fs0 plus fe0 equals 1. The electron balance, and thus the COD balance, is maintained. f s 0 + f e0 = 1 (2.44) 22 Biological Wastewater Treatment: Principles, Modelling and Design Table 2.7 Half reactions for biological systems (Metcalf & Eddy, 2003 a), (unit for ΔG0’ is kJ per electron equivalent b) Parameter Half reaction Reactions for bacterial cell synthesis (Rcs) 1 CO + 1 HCO + 1 NH + + H − + e − = 1 C H O N + 9 H O Ammonia as nitrogen 2 20 3 20 4 5 20 5 7 2 20 2 source 5 CO + 1 NO − + 29 H + + e − Nitrate as nitrogen = 1 C5 H7 O2 N + 11 H 2O 2 28 3 28 28 28 28 source ΔG0’ Nitrite -93.23 (2.17) -78.14 (2.18) -71.67 (2.19) = 1 H 2 S + 1 HS - + 1 H 2O 12 12 2 13.60 (2.20) = 1 H 2 S + 1 HS - + 1 H 2O 21.27 (2.21) 24.11 (2.22) Organic donors (heterotrophic reactions) 9 CO + 1 NH + + Domestic wastewater 31.80 (2.23) Proteins + − 9 1 1 2 50 4 50 HCO3 + H + e = 50 C10 H 19 O3 N + 25 H 2O 50 8 CO + 2 NH + + 31 H + + e − = 1 C16 H 24 O5 N 4 + 27 H 2O 2 33 4 33 33 66 66 1 HCO - + H + + e − 1 HCOO - + 1 H O = 3 2 2 2 2 1 CO + H + + e − 1 = C6 H 12O6 + 1 H 2O 2 4 24 4 1 CO + H + + e − = 1 CH 2O + 1 H 2O 2 4 4 4 32.22 (2.24) 48.07 (2.25) 41.96 (2.26) 41.84 (2.27) 1 CO + H + + e− 2 6 1 CO + 1 HCO - + H + + e − 2 10 3 5 1 CO + H + + e− 2 6 = 1 CH 3OH + 1 H 2O 37.51 (2.28) 35.78 (2.29) 31.79 (2.30) 1 CO + 1 HCO - + H + + e − 2 14 3 7 1 CO + 1 HCO - + H + + e − 2 8 3 8 = 1 CH 3CH 2COO - + 5 H 2O 27.91 (2.31) 27.68 (2.32) 27.61 (2.33) Eq. (2.15) (2.16) Reactions for electron acceptors (Ra) Oxygen Nitrate Sulfite Sulfate Carbon dioxide (methane fermentation) 1 NO 2 − + 4 H + + e − 2 3 3 1 O + H + + e− 2 4 1 NO − + 6 H + + e− 3 5 5 1 SO 2 − + 5 H + + e − 3 6 4 : 1 SO 2 − + 19 H + + e − 4 8 16 1 CO + H + + e − 2 8 = 1 N 2 + 2 H 2O 6 3 = 1 H 2O 2 = 1 N 2 + 3 H 2O 10 5 16 16 = 1 CH 4 + 1 H 2O 8 4 2 Reactions for electron donors (Rd) Formate Glucose Carbohydrates Methanol Pyruvate Ethanol Propionate Acetate: Grease (fats and oils) 4 CO + H + + e − 2 23 6 6 = 1 CH 3COCOO - + 2 H 2O 10 5 = 1 CH 3CH 2OH + 1 H 2O 12 4 14 14 = 1 CH 3COO - + 3 H 2O 8 8 = 1 C8 H 16 O + 15 H 2O 46 46 Inorganic donors (autotrophic reactions) Fe3 + + e − = Fe 2 + -74.40 (2.34) 1 NO − + H + + e − 3 2 1 NO − + 5 H + + e − 3 4 8 1 NO - + 4 H + + e − 2 3 6 1 SO 2- + 4 H + + e − 4 6 3 1 SO 2- + 19 H + + e − 4 8 16 1 SO 2- + 5 H + + e − 4 4 4 = 1 NO2- + 1 H 2O -40.15 (2.35) = -34.50 (2.36) -32.62 (2.37) 19.48 (2.38) 21.28 (2.39) 21.30 (2.40) 27.47 (2.41) 40.46 (2.42) 44.33 (2.43) 1 N + 4 H + + e− 6 2 3 H + + e− 1 SO 2- + H + + e − 4 2 a = = = = 2 2 1 NH + + 3 H O 4 8 2 8 1 NH + + 1 H O 4 3 2 6 1S+2H O 6 3 2 1 H S + 1 HS - + 1 H O 16 2 16 2 2 1 S O2− + 5 H O 8 2 3 8 2 = 1 NH 4+ 3 = 1 H2 2 = SO32 − + H 2O Adapted from McCarty (1975) and Sawyer et al. (1994). b Reactants and products at unit activity except [H+] = 10-7 M 23 Microbial Metabolism Table 2.8 Energy available from the transformation of glucose Aerobic oxidation of glucose Anaerobic oxidation of glucose (methanogenesis) ΔG0’ ED: glucose to CO2; EA: CO2 to CH4 (kJ/eeq) 1 1 1 C6 H 12O6 + H 2O → CO2 + H + + e − -41.96 Donor: 24 4 4 ED: glucose to CO2; EA: O2 to H2O Donor: 1 1 1 C6 H 12O6 + H 2O → CO2 + H + + e − 24 4 4 1 1 O2 + H + + e − → H 2O 4 2 Acceptor: -78.14 1 1 1 1 C6 H 12O6 + O2 → CO2 + H 2O 24 4 4 4 On a 1 mole basis for glucose, the net equation would become (• 24): C6 H 12O6 + 6O2 = 6CO2 + 6 H 2O Net: 60 58 Acceptor: 1 1 1 CO2 + H + + e − → CH 4 + H 2O 8 8 4 1 1 1 C6 H 12O6 = CH 4 + CO2 24 8 8 On a 1 mole basis for glucose, the net equation becomes (• 24): C6 H 12O6 = 3CH 4 + 3CO2 -120.10 Net: -2882 ΔG0’ (kJ/eeq) -41.96 24.11 -17.85 -428 -0.6 50 -0.5 30 -0.3 20 -0.2 + CO2 + 101 HCO3 + H + e = 1 10 CH3COCOO + Glucose/CO2 + Hydrogen/H 2 5 H2O + 1 6 1 8 CO2 + H +e = 121 CH3CH2OH + 41 H2O + CO2 + 81 HCO3 + H + e = 81 CH3COO + 83 H2O 1 8 CO2 + H + e = 1 8 SO4 + 1619 H + e = 161 H2S + 161 HS + 21 H2O + 1 8 CH4 + 41 H2O + 2 Pyruvate/CO2 Ethanol/CO2 Acetate/CO2 Denitrification 1 5 C6H12O6 + 14 H2O Methanogenesis -0.4 1 24 Fermentation 40 CO2 + H + e = + H + e = 12 H2 Methane/CO2 Sulfide/Sulfate 2+ 3+ 10 -0.1 0 0.0 CH4 + CO2 CO2 + H2O + N2 CO2 + H2O Denitrification: Aerobic oxidation: 0.4 0.5 0.6 -60 0.7 -70 -77 Acetate -50 0.3 Methanogenesis: -40 0.2 Product: -30 0.1 Fermentation: -20 NO3 + H + e = N2 + H2O Fe + e = Fe O2 + H + e = H2O Reaction: -10 E0’(volts) ΔG0’ (kJ/eeq) + + 1 1 CO2 + H + +e = C 6H12O6 + H 2O 24 4 H ++e = H21 2 + 1 2 1 1 CO 2 + HCO 3 + H + e = CH 3 COCOO + H 2O 10 5 10 5 + 1 1 CO 2 + H +e = CH 3CH2OH + H41 2O 6 12 + 1 1 3 1 CO 2+ HCO 3 + H + e = CH 3COO + H8 2O 8 8 8 + 1 1 2+ H + e = CH 4 + H41 2O CO 8 8 2 1 19 + 1 1 1 SO 4+ H + e = H 2S + HS + H 2O 8 16 2 16 16 + 1 6 3 1 5 NO3 + 5 H + e = 10 N2 + 5 H2O 3+ Fe + e = Fe2+ 0.8 1 4 + 1 2 O2 + H + e = H2O N2/Nitrate Fe(II)/Fe(III) H2O/O2 -80 Figure 2.17 Energy scale for redox couples with glucose as electron donor (adapted from Rittmann and McCarty, 2001) Aerobic oxidation Substrate: Glucose + 1 4 24 Biological Wastewater Treatment: Principles, Modelling and Design The active bacterial cells generated by growth using the initial electron donor then undergo decay due to maintenance, predation and cell lysis. During decay, a portion of the active bacterial cells become the electron donor to generate more energy and more reaction end products. The global split of electron equivalents between active residual cells (fs: observed synthesis fraction) and reaction end products (fe: observed energy fraction) remains equal to 1. f s + fe = 1 The fraction fs and fs can be expressed in mass units, rather than on an eeq basis, and are then called true yield (or maximum theoretical yield; Y) and observed yield (Yobs), respectively. The fraction fs0 can be used to estimate the true yield Y: f s0 M c 8ne where: Mc 8 (2.46) gram cells per empirical mol of cells number of gram thCOD per eeq (see half reaction Eq.2.18 in Table 2.7) number of eeq per empirical mol of cells ne With C5H7O2N as the empirical formula for cells, the molecular weight is 113 g/mol. With ammonia as the nitrogen source for its synthesis, there are 20 eeq per empirical mole of cells (Table 2.7, reaction Eq. 2.15) and the above equation can be simplified to: f s0 Y = f s 0.706 = 1.42 gCOD/gCells 0 fe0 1 (2.47) fs0 fs Growth Yobs = fsM c 8ne (2.48) Active residual cells Decay Figure 2.18 Use of electron donor for energy production and cell synthesis. Note: f: fraction of electrons donated; e: energy; s: synthesis (adapted from Rittmann and McCarty, 2001). 2.3.3.3 Observed yield from stoichiometry If an empirically balanced stoichiometric equation can be obtained for biomass synthesis from a given wastewater, the biomass observed yield can be calculated. Using the protein casein to represent wastewater in laboratory experimentation with activated sludge, Porges et al. (1956) proposed the following equation: C8 H 12O3 N 2 + 3O2 → casein C5 H7 O2 N + NH 3 + 3CO2 + H 2O (2.49) bacterial cells g weight C8H12O3N2 3O2 184 96 Sum C5H7O2N NH3 3CO2 H2O 113 280 g /gCasein 1.00 gCOD/mol 1.42 g COD 256 g COD/gCOD Similarly, fs can be used to estimate the observed yield Yobs, Active bacterial cells Cell synthesis Sum where the ratio of 1.42 gram COD per gram cells was also calculated in Table 2.6. Reaction end products fe Electron donor (2.45) 0 Y= Energy production 18 0.61 (Yobs) -1.00 1.39 -96 160 160 1.00 17 132 280 -0.38 (-fe) 0 0 0 0 0 0 160 0.62 (fs) Thus, consuming 184 g of casein requires 96 g of oxygen and produces 113 g bacterial cells and other reaction end products. Similar proportions would be expected for a full-scale wastewater treatment plant treating this compound (which is of comparable composition to typical domestic wastewater). The biomass true yield (Y) is thus, 0.61 g biomass per g substrate consumed (= 113/184). Note that the mass of products equals that of reactants (280 g/mol of casein consumed). 25 Microbial Metabolism On a COD basis, the thCOD of casein being 1.39 gCOD/gCasein (Table 2.5) gives 256 gCOD/molCasein, and the thCOD of bacterial cells of composition C5H7NO2 being 1.42 gCOD/gVSS, gives 160 gCOD/molCells. The observed synthesis fraction (fs) is thus 0.62 gCOD/gCOD (0.61x1.42/1.39). The oxygen requirement is 96 g O2 per mole of casein consumed, corresponding to 0.52 gO2/gCasein (96/184). Thus, the energy production fraction (fe) is 0.38 g COD of O2 per g COD of casein (0.52/1.39). Note that oxygen has a negative COD (-1.0 gCOD/gO2) and the COD balance is maintained. f s + fe = 0.62 + 0.38 = 1.00 (2.50) The experimentally reported observed (and not “true”) synthesis fraction (fs) of 0.62 is quite high in comparison to other values published in the literature for wastewater treatment. Thus, the true synthesis fraction (fs0) should only be a little higher and the cells were probably close to their exponential growth phase, a condition in which the fraction of energy obtained from endogenous decay is minimal. Indeed, using the methodology presented in the next section, and the half reaction and free energy change value presented in Table 2.7 for protein, which has a very similar chemical structure to that of casein, a true synthesis fraction (fs0) of 0.64 can be calculated. The nitrogen and phosphorus requirements for cell growth can be evaluated by considering that they constitute 12.0 and 2.0%, respectively, of the volatile fraction of the biomass produced (the CHON fraction) as can be estimated in the empirical equation C5H7NO2P1/12 (Table 2.5). In the above example, for 113 g of biomass produced (corresponding to 184 g of casein degraded), 13.4 g of nitrogen would need to be added either from organic (e.g. casein) or inorganic sources (e.g. ammonia). Similarly, 2.26 g of phosphorus would need to be added per 113 g of biomass produced. 2.3.3.4 True yield estimation from bioenergetics Bioenergetics can be used as an alternative to conducting careful laboratory scale experimentation to determine the true (or maximum) yield of a reaction. The approach presented below is adapted from that of Metcalf & Eddy (2003) which is a simplification of that of Rittmann and McCarty (2001) which was recently updated by McCarty (2007). An alternative approach has been developed by Heijnen et al. (in preparation) which mainly differs from the above ones in its estimation of anabolic energy need by an energy dissipation function instead of an efficiency factor. These references provide additional details to those presented below for the development of other half reactions and their free energy changes, complex fermentation reactions, autotrophic reactions and non standard conditions. The simplified procedure presented below is divided into 4 steps which consist in determining, (i) the energy provided from catabolism knowing the electron donor, the electron acceptor and the source of nitrogen for growth, (ii) the energy needed for cell synthesis (anabolism), (iii) the energy needed for the overall growth reaction (metabolism) and (iv) the true yield (Y) coefficient. A. Energy providing reaction (catabolism) The methodology to develop the reaction and associated Gibbs energy production for the catabolic reaction of the electron donor (ED) and electron acceptor (EA) was presented in section 2.3.3.1. The method of Rittman & McCarty (2001) assumes that only a fraction (40 to 80%, typically 60%) of the energy available from an oxidation-reduction reaction is used in the anabolism while the rest is lost as heat. ΔGcata = K ΔGR where: ΔGcata K ΔGR (2.51) Gibbs energy available for catabolism from 1 eeq of ED (kJ/eeq) fraction of energy transfer captured (typically 0.60) Gibbs energy released from 1 eeq of ED (kJ/eeq) B. Energy needed for cell synthesis (anabolism) The energy needed to synthesise heterotrophic biomass from an electron donor is estimated by considering pyruvate as a central metabolic intermediate and a source of nitrogen for biomass synthesis. ΔGana = where: ΔGana ΔGp ΔGP Km + ΔGc + ΔGN K (2.52) Gibbs energy required for anabolism from 1 eeq of ED (kJ/eeqED) Gibbs energy required to convert 1 eeq of ED to pyruvate (kJ/eeqED) 26 Biological Wastewater Treatment: Principles, Modelling and Design m ΔGc ΔGN constant: +1 if ΔGp is positive (endergonic) and -1 if ΔGp is negative (exergonic) Gibbs energy required to convert 1 eeq of pyruvate to cells = 31.41 kJ/eeqCells free energy required per eeq of cells to reduce nitrogen to ammonia (kJ/eeqCells) = 17.46, 13.61, 15.85, 0.00 for NO3-, NO2-, N2 and NH4+, respectively. The first term of the equation describing the conversion of the electron donor to pyruvate has an exponent m on the efficiency fraction K. Should ΔGp be positive, as would be the case for acetate being transformed to pyruvate, this reaction would require energy (endergonic) and the positive value to m results in a greater value (more energy needed) for this first term. Should ΔGp be negative, as would be the case for glucose being transformed to pyruvate, this reaction would release energy (exergonic) and the negative value to m would result in a lower value (less energy needed) for this first term. C. Energy for the overall growth reaction (metabolism) Two mass balance equations can be written, one that was already presented for the electron donor for which its electrons are used for energy and synthesis fe0 + fs0 = 1 From the ED mass balance, fs0 and fe0 can be found f s0 = 1 ⎛ f0⎞ 1 + ⎜ e0 ⎟ ⎝ fs ⎠ (2.57) and f e0 = 1 − f s 0 (2.58) D. True yield (Y) The true yield, Y, can then be expressed in mass fraction once fs0 has been determined by using Eq. 2.47 presented earlier for an empirical biomass equation of C5H7O2N produced with ammonia as the nitrogen source. Y = f s 0 0.706 = f s0 1.42 gCOD/gCells (2.59) An example is presented below for estimating the true yield from bioenergetics. 2.3.3.5 Example: Estimating true yield from bioenergetics for the aerobic oxidation of glucose with ammonia as nitrogen source (2.53) A. Energy providing reaction (catabolism) and one for energy where as much energy is consumed for anabolism as provided by catabolism. The negative sign accounts for the fact that anabolism consumes rather than produces energy: -fs0 ΔGana = fe0 ΔGcata (2.54) This equation can be rewritten to visualise that the energy required for cell growth (anabolism) is provided by the energy released from catabolism times the ratio of ED oxidised to ED used for cell synthesis. − ΔGana = The reaction and energy available from the aerobic oxidation of glucose were developed above from half reactions 1 1 1 1 C6 H 12O6 + O2 = CO2 + H 2O 24 4 4 4 -120.10 kJ/eeq and ΔGcata = K ΔGR = 0.6 × ( −120.10) = −72.06 kJ/eeq 0 fe ΔGcata f s0 (2.55) It can also be rewritten to isolate the unknowns (fe0/fs0). fe ΔGana =− ΔGcata f s0 0 (2.56) 1 1 1 C6 H 12O6 + H 2O = CO2 + H + + e − 24 4 4 B. Energy needed for cell synthesis (anabolism) The reaction and Gibbs energy required to convert 1 eeq of glucose to pyruvate is: 27 Microbial Metabolism ED: 1 1 1 C6 H 12O6 + H 2O = CO2 + H + + e − 24 4 4 (ΔG0’ = -41.96 kJ/eeq) EA: 1 1 CO2 + HCO3 + H + + e− = 5 10 1 2 CH 3COCOO − + H 2O 10 5 (ΔG0’ = +35.78 kJ/eeq) Net: 1 1 1 1 C6 H 12O6 + H 2O + CO2 + HCO3 − = 24 20 5 10 1 1 CH 3COCOO − + H 2O 10 4 (ΔG0’ = -6.18 kJ/eeq) and: ED: EA: ΔGp M also: K ΔGc ΔGN thus: ΔGana glucose to CO2; (ΔG0’, kJ/eeq) CO2 to pyruvate, (ΔG0’, kJ/eeq) -6.18 kJ/eeq -1 (since ΔGp is negative) 0.6 31.41 kJ/eeqCells 0.00 kJ/eeqCells with NH4+ as the nitrogen source = (ΔGp / Km ) + ΔGc + (ΔGN/K) = (-6.18 / 0.6-1) + 31.41 + 0 = +27.70 kJ/eeq C. Overall reaction for growth (metabolism) The ratio of the fractions fe0 / fs0 can now be calculated. ⎛ ΔGana f e0 / f s 0 = − ⎜ ⎝ ΔGcata ⎞ ⎛ 27.70 ⎞ ⎟ = −⎜ ⎟ = 0.38 ⎝ −72.06 ⎠ ⎠ and fs0 = 1 / (1 + (fe0 / fs0)) = 1 / (1 + 0.38) = 0.72 gCellCOD/gCOD consumed fe0 = 1 - fs0 = 0.28 gCOD/gCOD consumed D. True yield in mass units. The true yield in mass units, considering an empirical biomass equation of C5H7O2N is: Y= f s0 =0.51 gVSS/gCOD consumed 1.42 2.4 KINETICS 2.4.1 Substrate utilisation rate The rate of substrate utilisation by bacteria depends on a number of factors that are characteristic of a given microbial group. The most important parameters are the maximum substrate utilisation rate and half saturation and inhibition constants. 2.4.1.1 Saturation function The microbial substrate utilisation rate mainly depends on its maximum substrate utilisation rate, the amount of biomass present and the concentration of substrate used for growth. rs = k M s X (2.60) where: substrate utilisation rate (g COD/m3.h) maximum specific substrate utilisation rate (g COD/gVSS.h) saturation function for soluble substrate SS (gCOD/gCOD) biomass concentration (gVSS/m3) rs k MS X The effect of substrate concentration on the rate of reaction is considered by the saturation function. MS = SS ( K S + SS ) where: SS KS (2.61) substrate concentration (gCOD/m3) substrate half saturation constant (gCOD/m3) The saturation function (MS) varies from 0 to 1 as a function of the concentration of substrate available in solution near the biomass (Figure 2.19). The substrate utilisation rate is null in the absence of substrate. At the half saturation constant concentration, the saturation function value is 0.5 and the substrate utilisation rate is half of the maximum value. At nine times the half saturation value, the substrate utilisation rate is 90% of its maximum and at an infinitely high concentration, the saturation function reaches a value of 1.0 and the substrate utilisation rate is at its maximum value. 28 One form of inhibition function that is commonly used is the following. Figure 2.19 Effect of substrate concentration on the saturation function and kinetic of substrate utilisation. Constants used were: KS = 5 gCOD/m3, k = 4 gCOD /gVSS.d, X = 250 gVSS/m3 The effect of varying the inhibitory concentration from 0 to 10 times its half saturation value is illustrated in Figure 2.20. 1.0 1000 0.8 800 0.6 600 0.4 400 0.2 200 KS 0 0 0 10 20 30 40 50 II = KI ( K I + SI ) where: KI half saturation constant of the inhibitory compound (g/m3) concentration of the inhibitory compound (g/m3) SI rS = k M S M SO2 M SNH 3 M SPO4 X (2.62) where MSO2, MSNH3 and MSPO4 represent the saturation functions for oxygen, ammonia and phosphate, respectively. Inhibition function , II Substrate concentration, SS (gCOD/m3) The effect of other limiting nutrients (e.g. oxygen, ammonia, phosphate) could also be considered in this substrate utilisation rate formulation by multiplying with the various saturation functions (also called switching functions). (2.65) 1.0 1000 0.8 800 0.6 600 0.4 400 0.2 200 KI 0 0 100 200 300 400 0 500 Substrate utilisation rate, rS (gVSS/m3.d) Saturation function (MS) Substrate utilisation rate, rS (gVSS/m3.d) Biological Wastewater Treatment: Principles, Modelling and Design Inhibiter concentration, SI (g/m3) According to Liebig's law of minimum, however, growth is considered to be limited by only one nutrient. Thus, a more appropriate formulation would be to consider only the minimum of the various saturation functions in the above equation. Eq. 2.62 needs adjustment with the MIN operator which applies to the functions between parentheses and not k: rs = k ⋅MIN ( M S M SO2 M SNH3 M SPO4 ) ⋅ X 2.4.1.2 (2.63) Inhibition function In the presence of an inhibitory compound, a saturation function can be used to slow down the substrate utilisation rate. rS = k I I X where: II (2.64) Figure 2.20 Inhibition kinetics. Constants used were: KI = 50 g/m3, k = 4 gCOD/gVSS.d, X = 250 gVSS/m3 The inhibition function considered here has a mirror effect to that of the saturation function. No effect on the substrate utilisation rate is seen at a null value of inhibitor concentration. At the half saturation constant concentration, the inhibition function value is 0.5 and the substrate utilisation rate is half of the maximum value. At nine times the half saturation value, the substrate utilisation rate is only 10% of its maximum and at an infinitely high inhibitor concentration, the substrate utilisation rate is completely inhibited. More details on inhibition are provided in Chapter 10. 2.4.2 Growth rate When the rate of substrate utilisation is at its maximum, the growth rate is also at its maximum and their ratio is, theoretically, that of the true yield. μmax = Y k inhibition function compound (g/g) for the inhibitory where: (2.66) 29 Microbial Metabolism μmax maximum growth (gVSS/gVSS.d) rate of biomass The growth rate of a biomass depends on its rate of substrate utilisation for cell synthesis and on its decay rate which is proportional to the concentration of biomass present. rg = Y rs − b X where: rg b (2.67) biomass growth rate (gVSS/m3.d) specific biomass decay rate (gVSS/gVSS.d) Substituting in equations presented earlier gives: rg = Y k M S X − b X (2.68) rg = μmax M S X − b X (2.69) ⎛ SS ⎞ rg = μ max ⎜ ⎟X −b X ⎝ K S + SS ⎠ (2.70) The specific growth rate is obtained by dividing the growth rate by the biomass concentration. rg where: μ specific biomass growth rate (gVSS/gVSS.d) or ⎛ SS K ⎝ S + SS μ = μ max ⎜ ⎞ ⎟−b ⎠ (2.72) (2.74) μmax = Y k − b (2.75) (ii) The minimum substrate concentration required at which the rate of cell synthesis just equals its rate of decay is when the specific growth rate (μ) is zero which gives: S S min = b KS Y k −b where: SSmin (2.76) minimum concentration required to achieve a null growth rate (gCOD/m3) (iii) At a null substrate concentration (SS = 0 gCOD/m3), the specific growth rate becomes negative and is equal to the rate of decay. μ = −b (2.71) X Ss =1 ( K S + SS ) and (2.77) 2.5 Specific growth rate, μ (gVSS/gVSS.d) μ= MS = 2.0 μmax = 2.3 gVSS/gVSS.d 1.5 SSmin = 0.22 gCOD/m3 1.0 0.5 -b = -0.1 gVSS/gVSS.d 0 0 or 1 2 3 4 5 Substrate concentration, SS (gCOD/m3) ⎛ SS K ⎝ S + SS μ =Y k⎜ ⎞ ⎟−b ⎠ (2.73) The effect of substrate concentration on the specific growth rate, as calculated from the above equation, is illustrated in Figure 2.21. The following aspects are apparent from this graph: (i) The maximum specific growth rate is obtained at a high (infinite) substrate concentration at which point: Figure 2.21 Effect of substrate concentration on biomass growth rate. Constants used were b = 0.1 gVSS/gVSS.d, k = 4 gVSS/gVSS.d, KS = 5 gCOD/m3, Y = 0.6 gVSS/gCOD 2.4.3 Stoichiometric and kinetic parameter values Typical values of stoichiometric and kinetic parameters for various bacterial groups are presented in Table 2.9. In general, a higher fS0 (or true yield, Y) results in a higher maximum specific growth rate (μmax) which results in higher specific removal rates (k = μmax / Y). 30 Biological Wastewater Treatment: Principles, Modelling and Design Table 2.9 Typical values of stoichiometric (fS0, Y) and kinetic (qmax, μmax ) parameters for various bacterial groups, (adapted from Rittmann and McCarty 2001) Electron acceptor Electron donor fS 0 Y µmax K - Microbial group e donor Chemotrophic organotrophs Aerobic heterotrophs Sugar No sugar Aerobic heterotrophs Denitrifiers Organic Fermenting organisms Sugar Sulphate reducers Acetate Methanogens Acetate (acetoclastic) Chemotrophic lithotrophs Nitrifiers :AOB NH4Nitrifiers :NOB NO2Methanogens H2 (hydrogenotrophic) O2 O2 NO3-, NO2Organic SO42Acetate 0.70 0.60 0.50 0.18 0.08 0.05 0.49 gVSS/gbCOD 0.42 gVSS/gbCOD 0.25 gVSS/gbCOD 0.18 gVSS/gbCOD 0.057 gVSS/gbCOD 0.035 gVSS/gbCOD 13.2 8.4 4.0 1.2 0.5 0.3 27.0 g bCOD/gVSS.d 17.0 g bCOD/gVSS.d 16.0 g bCOD/gVSS.d 10.0 g bCOD/gVSS.d 8.7 g bCOD/gVSS.d 8.4 g bCOD/gVSS.d O2 O2 CO2 0.14 0.10 0.08 0.34 gVSS/gNH4-N 0.08 gVSS/gNO2-N 0.45 gVSS/gH2 0.9 0.5 0.3 2.7 g NH4-N /gVSS.d 1.1 g NO2-N/gVSS.d 1.1 g H2/gVSS.d bCOD: biodegradable COD µmax in gVSS /gVSS d k = µmax /Y= specific rmax (per unit biomass) REFERENCES Comeau Y., Hall K.J., Hancock R.E.W. and Oldham W. K. (1986) Biochemical model for biological enhanced phosphorus removal. Wat. Res. 20, 1511-1521. Eikelboom D.H. (2000) Process Control of Activated Sludge Plants by Microscopic Investigation ISBN: 9781900222297, pg.156 Heijnen J.J., Kleerebezem R. and van Loosdrecht M.C.M. (in preparation) A generalized method for thermodynamic state analysis of environmental systems. Kampschreur M.J. Tan N.C.G., Kleerebezem R., Picioreanu C., Jetten M.S.M., van Loosdrecht M.C.M. (2008) Effect of dynamic process conditions on nitrogen oxides emission from a nitrifying culture. Environ. Sci. Techn., 42(2), 429-435. McCarty P.L. (2007) Thermodynamic electron equivalents model for bacterial yield prediction: Modifications and comparative evaluations. Biotech. Bioeng. 97(2), 377388. Metcalf & Eddy Inc. (2003) Wastewater Engineering Treatment and Reuse, 4th ed., McGraw-Hill, New York. Madigan M.T. and Martinko J.M. (2006) Brock Biology of Microorganisms (11th ed.). San Francisco, CA,Pearson Education, Inc. Rittmann B.E. and McCarty P.L. (2001) Environmental Biotechnology - Principles and Applications. New York, McGraw-Hill. Todar K. (2007) Microbial metabolism (Electronic version). Retrieved September 17, 2007 from http://www.bact.wisc.edu/themicrobialworld/metabolis m.html NOMENCLATURE Symbol Description Unit b Specific biomass decay rate gVSS/VSS.d fe Observed energy fraction of COD used gCOD/gCOD fe0 True energy fraction of COD used gCOD/gCOD fs Observed synthesis fraction of COD used gCOD/gCOD fs0 True synthesis fraction of COD used gCOD/gCOD II Inhibition function for the inhibitory compound g/g 31 Microbial Metabolism k Maximum specific substrate utilization rate gCOD/gVSS.h K Fraction of energy transfer captured kJ/kJ KI Half saturation constant of the inhibitory compound g/m3 KS Substrate half saturation constant gCOD/m3 m Constant: +1 if ΔGp is positive and -1 if ΔGp is negative Mc Weight of cells per empirical mole of cells g/mol MS Saturation function for soluble substrate SS gCOD/gCOD ne Number of electron equivalents per empirical mol of cells eeq/mol rg Biomass growth rate gVSS/m3.d rS Substrate utilisation rate g COD/m3.h SI Concentration of the inhibitory compound g/m3 SS Substrate concentration gCOD/m3 SSmin Minimum concentration required to achieve a null growth rate gCOD/m3 X Biomass concentration gVSS/m3 Y True yield gVSS/gCOD Yobs Observed yield gVSS/gCOD ΔGana Gibbs energy required for anabolism from 1 eeq of electron donor (ED) kJ/eeqED ΔGc Gibbs energy required to convert 1 eeq of pyruvate to cells kJ/eeqED ΔGcata Gibbs energy available for catabolism from 1 eeq of ED kJ/eeq ΔGN Free energy required per eeq of cells to reduce nitrogen to ammonia kJ/eeqED o’ o ΔG Change in Gibbs free energy at standard conditions (25 C, 1 M, 1 atm) but pH 7 kJ/mol ΔGp Gibbs energy required to convert 1 eeq of electron donor (ED)to pyruvate kJ/eeqED ΔGR Gibbs energy released from 1 eeq of ED kJ/eeq Abbreviation Description ADP Adenosine diphosphate AMP Adenosine monophosphate AN Anaerobic AOB Ammonia oxidizing bacteria ATP Adenosine triphosphate AX Anoxic bCOD Biodegradable COD COD Chemical oxygen demand DGGE Denaturing gradient gel electrophoresis DNA Deoxyribonucleic acid ETC Electron transport chain FISH Fluorescent in situ hybridization FSS Fixed (inorganic) suspended solids GAO Glycogen accumulating organisms NOB Nitrite oxidizing bacteria OX Aerobic PAO Phosphorus accumulating organism PCR Polymerase chain reaction PHA Polyhydroxyalcanoates 32 Biological Wastewater Treatment: Principles, Modelling and Design Pi Inorganic phosphate pmf Proton motive force RNA Ribonucleic acid thCOD Theoretical chemical oxygen demand TSS Total suspended solids VSS Volatile suspended solids Greek symbols Explanation Unit μ Specific growth rate of biomass gVSS/gVSS.d μmax Maximum specific growth rate of biomass gVSS/gVSS.d Colony of protozoa in an activated sludge ecosystem: (photo: D. Brdjanovic)