Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

This article was downloaded by: [UZH Hauptbibliothek / Zentralbibliothek Zürich]

On: 05 July 2014, At: 18:06


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in Environmental Control


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/best19

Methane production from municipal refuse: A review of


enhancement techniques and microbial dynamics
a b c d
Morton A. Barlaz , Robert K. Ham , Daniel M. Schaefer & Ron Isaacson
a
Department of Civil Engineering , North Carolina State University , Raleigh, North Carolina
b
Department of Civil Engineering , University of Wisconsin , Madison, Wisconsin
c
Departments of Meat and Animal Science and Bacteriology , University of Wisconsin ,
Madison, Wisconsin
d
Department of Environment and Safety , Gas Research Institute , Chicago, Illinois
Published online: 09 Jan 2009.

To cite this article: Morton A. Barlaz , Robert K. Ham , Daniel M. Schaefer & Ron Isaacson (1990) Methane production from
municipal refuse: A review of enhancement techniques and microbial dynamics, Critical Reviews in Environmental Control,
19:6, 557-584

To link to this article: http://dx.doi.org/10.1080/10643389009388384

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Volume 19, Issue 6 (1990) 557

METHANE PRODUCTION FROM MUNICIPAL REFUSE: A


REVIEW OF ENHANCEMENT TECHNIQUES AND MICROBIAL
DYNAMICS
Authors: Morton A. Barlaz
Department of Civil Engineering
North Carolina State University
Raleigh, North Carolina

Robert K. Ham
Department of Civil Engineering
Daniel M. Schaefer
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Departments of Meat and Animal Science


and Bacteriology
University of Wisconsin
Madison, Wisconsin

Referee: Ron Isaacson


Department of Environment and Safety
Gas Research Institute
Chicago, Illinois

I. INTRODUCTION
In 1986 it was estimated that an average of 1.64 kg of municipal solid waste were
generated per person per day in the U.S. 62 About 95% of the 133 million tons of municipal
solid waste disposed of in the U.S. in 1984 were buried in sanitary landfills.18 Refuse
disposal is an issue that must be addressed by virtually every community in the country. In
the last 2 decades, the demand for disposal techniques that minimize adverse environmental
impacts has increased. In addition, government officials have become more interested in
energy recovery from wastes as a way to reduce both the cost of waste disposal and national
dependence on foreign energy supplies.
Methane is the primary terminal product of a series of biologically mediated reactions
involved in refuse decomposition in sanitary landfills and is a usable source of energy. In
addition to its energy value, methane production in a sanitary landfill has several advantages.
After the onset of methane production, a reduction in leachate strength is expected, leading
to lower leachate treatment costs and a reduced risk of groundwater contamination. A second
advantage is the potential to reduce costs or even profit from gas control through gas recovery
and utilization equipment. Finally, most of the settlement of a landfill can be expected by
the end of the refuse decomposition period. Thus, enhanced refuse decomposition will lead
to a reduction in long-term care requirements for gas migration and cover maintenance.
In 1984, there were 75 landfill gas-recovery facilities in the planning stage or already
in operation across the U.S. 12 and a more recent study reported that 77 projects are in
operation.39 However, measurement and prediction of methane yields from sanitary landfills
is difficult.48 Where the volume of methane generated from refuse in sanitary landfills has
been reported, it has ranged from 0.00034 to 0.068 m3 CH4 per dry kg of refuse.46 The
methane potential of refuse as measured from biodegradability data44 is 0.13 m3 CH4 per
kg dry refuse. Thus, measured methane yields range from less than 1 to 52% of the expected
methane yield from biodegradability data. Many smaller landfills are bypassed for energy
recovery projects due to uncertain economics. Whether or not methane is recovered, it must
558 Critical Reviews in Environmental Control

TABLE 1
Composition of Municipal Refuse by Component

Percent by* Percent by


wet weight wet weight
Component 1970s* (%) Component 1986C (%)

Food 15 Food 7.9


Garden trimmings 12 Yard 17.9
Paper, cardboard 44 Paper, paperboard 41.0
Plastics 3 Plastics 6.5
Textile, rubber, leather, 5 Textile, rubber, leather. 8.1
wood wood
Glass 8 Glass 8.2
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Tin cans, ferrous, and non- 9 Metal, ceramics 8.7


ferrous metals
Dirt, ashes, brick, etc. 4 Miscellaneous, inorganic 1.6"
waste

* As reported by Tchobanoglous et aI. M for municipal solid waste in the early 1970s.
b
The moisture content of refuse with this composition, based on typical moisture contents for
each component of the refuse, is 21.9% (wet weight).82
c
As reported by Lewis62 for municipal solid waste in 1986.
d
Includes ash and rock.

be vented from landfills so that it does not migrate laterally underground and become an
explosion hazard.82
In order to understand, predict, and possibly control refuse decomposition in sanitary
landfills, one must first understand the microbiology of methane production from refuse.
Research on microbial processes in sanitary landfills is presented in the second section of
this article. With an understanding of the microbiology in hand, one can interpret the effects
of various factors on the decomposition process. In Section ITJ, the effects of parameters
thought to enhance refuse decomposition are discussed. In Section IV, methods to assess
the extent of refuse decomposition and rate of methane generation in full-scale landfills are
presented. Accurate assessment of the methane potential of a landfill is critical for devel-
opment of economically viable energy recovery projects.

A. REFUSE COMPOSITION
Prior to reviewing studies of refuse decomposition, it is important to understand the
composition of municipal solid waste. Traditionally, municipal refuse has been classified
according to visual categories such as glass, paper, metals, etc., as listed in Table 1. Data
presented in Table 1 are based on typical refuse composition reported in both the mid-1970s
and the late 1980s. While the methods used for characterization of municipal solid waste
may not have been precisely the same for the two studies presented in Table 1, it is clear
that paper and paper-related products remain the dominant constituent. It must be recognized
that the typical composition presented in Table 1 is subject to wide variability at certain
times of the year. For example, there are large increases in the concentration of paper from
gift wrapping in late December and garden trimmings in the summer. Given this sort of
variability, it is not possible to state whether the apparent increase in the plastics concentration
of MSW given in Table 1 is significant, as would be expected, or simply a result of refuse
variability.
Recently, data on the chemical composition of refuse have been published, and these
data have been used to calculate a methane potential for each chemical constituent.9 The
chemical constituent data in Table 2 indicate that the cellulose plus hemicellulose fraction
of refuse accounts for 91% of its methane potential. Lignin was considered to have no
Volume 19, Issue 6 (1990) 559

TABLE 2
Composition and Methane Potential
of Municipal Refuse by Chemical
Constituent

Chemical %Dry Methane


constituent weight potential*

Cellulose 51.2% 73.4%


Hemicellulose 11.9 17.1
Protein* 4.2 8.3
Lignin 15.2 0
Starch 0.5 0.7
Pectin' <3.0 —
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Soluble sugars 0.35 0.5

Total
volatile solids'1 78.6

Adapted from Barlaz, M. A., Ph.D. thesis, De-


partment of Civil and Environmental Engineer-
ing, University of Wisconsin, Madison, 1988.

• Data expressed as a percentage of the total


methane potential based on the cellulose hem-
icellulose, protein, sugar, and starch data. The
methane contribution of pectin was not cal-
culated because of the uncertainty associated
with the pectin concentration in refuse. Meth-
ane potential was calculated from the stoi-
chiometry given by Parkin and Owen69 on the
basis of 100% conversion of a constituent to
carbon dioxide and methane. It should be rec-
ognized that 100% of the degradable constit-
uents will not be mineralized, as some fraction
is surrounded by lignin and not accessible for
anaerobic degradation.
b
Determined by multiplication of the total Kjel-
dahl nitrogen by 6.25. The actual protein con-
tent of refuse is probably lower than the value
given here because some of the Kjeldahl ni-
trogen is actually nitrogen containing humic
materials and structural proteins that are not
easily degradable."
c
Actual value is probably less than 3%, but
could not be quantified.
* An independent measure of volatile solids
based on weight loss on ignition at 550°C.
e
The volatile solids concentration is presented
to illustrate that the chemical constituent anal-
yses account for 110% of the volatile solids.
No methane potential is calculated for volatile
solids because this measurement includes both
degradable and nondegradable carbon.
560 Critical Reviews in Environmental Control

BIOLOGICAL
POLYMERS
HYDROLYTIC AND
FERMENTATIVE
MICROORGANISMS

ALCOHOLS, 20%
CARBOXYLIC
ACIDS
(axcapt a e t i a t t )
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

0BLI6ATE
PROTON
REDUCERS

HYDROGENOPHMC ACETOPHILIC
METHANOGENS METHANOGENS

FIGURE 1. Substrate flow in anaerobic ecosystems." Complete multistep methanogenesis as it occurs in lacustrine
and sulfate-depleted marine sediments, bogs, marshes, trees, and digesting sludge. (Reprinted by permission from
author.)

methane potential because it is recalcitrant under the anaerobic conditions required for
methane production.87

H. LANDFILL MICROBIOLOGY
As given in the previous section, municipal refuse typically contains 40 to 50% cellulose,
10 to 15% lignin, 12% hemicellulose, and 4% protein on a dry weight basis. Its decomposition
to methane in sanitary landfills is a microbially mediated process that proceeds along path-
ways similar to those documented for anaerobic sludge digestion. In the first part of this
section, the general pathway for anaerobic decomposition, as it has been documented to
occur in other anaerobic ecosystems, is reviewed. Implications of this pathway for sanitary
landfills are also presented. Following this general pathway review, studies focusing on the
microbiology of refuse decomposition are presented. Finally, efforts to present descriptions
of refuse decomposition that consider both chemical composition and microbial population
dynamics are presented in the third part of this section.

A. THE MICROBIOLOGY OF ANAEROBIC DECOMPOSITION


Three trophic groups of anaerobic bacteria are required for the production of methane
from biological polymers such as cellulose and hemicellulose,8389 as illustrated in Figure 1.
Volume 19, Issue 6 (1990) 561

The first group of microorganisms, referred to as the hydrolytic and fermentative micro-
organisms in Figure 1, is responsible for the hydrolysis of polymers such as carbohydrates,
fats, and proteins. The initial products of polymer hydrolysis are soluble sugars, amino
acids, long-chain carboxylic acids, and glycerol. Hydrolytic and fermentative microorgan-
isms then ferment the intermediates listed above to short-chain carboxylic acids, carbon
dioxide, and hydrogen. Acetate, a direct precursor of methane, and alcohols are also formed.
The second group of bacteria active in the conversion of biological polymers to methane is
the obligate proton-reducing acetogens. They oxidize the fermentation products of the first
group of microorganisms to acetate, carbon dioxide, and hydrogen. The conversion of
fermentation intermediates like butyrate, propionate, and ethanol, which are the primary
substrates of the acetogenic bacteria, is only thermodynamically favorable at very low
hydrogen concentrations. Thus the obligate proton-reducing acetogenic bacteria only function
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

in syntrophic association with a hydrogen scavenger such as a methanogen or a sulfate-


reducing organism. The importance of hydrogen consuming acetogenic bacteria has not been
established in the municipal refuse ecosystem. This microbiological activity probably com-
petes weakly with the hydrogenophilic methanogens for hydrogen.
The third group of bacteria necessary for the production of methane is the methanogens.
Methanogens can utilize only a limited number of substrates, including formate, methanol,
methylamine, hydrogen plus carbon dioxide, and acetate. In sludge digesters it is estimated
that 70% of the methane produced originates from acetate.90 This value has not been in-
vestigated in the landfill ecosystem. The production of methane from acetate yields only 31
kJ/mol CH4 produced. This is barely enough energy for the generation of adenosine tri-
phosphate (ATP), which requires 30.6 kJ/mol, so the growth of methanogens on acetate is
relatively slow. The conversion of hydrogen and carbon dioxide to methane yields 135.6
kJ/mol CH4 produced. Thus, the latter reaction is energetically more favorable. The meth-
anogens are most active in the pH range 6.8 to 7.4.88
The importance of the methanogens in anaerobic digestion has been summarized by
Zeikus.90 As a group, the methanogens:

1. Control the pH of their ecosystem by the consumption of acetate.


2. Regulate the flow of electrons by the consumption of hydrogen, creating thermodyn-
amically favorable conditions for the catabolism of alcohols and acids.
3. Excrete organic growth factors, including vitamins and amino acids, that are used by
other heterotrophic bacteria in the ecosystem.

Should the activity of the fermentative organisms exceed that of the acetogens and meth-
anogens, there will be an imbalance in the ecosystem. Carboxylic acids and hydrogen will
accumulate and the pH of the system will fall, thus inhibiting methanogenesis.
In an active anaerobic ecosystem in which a hydrogen-scavenging population, i.e.,
methanogens, is present, glucose will be fermented to acetate, carbon dioxide, and hydrogen,
all products that can be metabolized by the methanogens.64 In the absence of a hydrogen-
scavenging population, other reduced end products such as lactate and ethanol will be formed,
but not further oxidized due to the accumulation of hydrogen. Thus, consumption of fer-
mentation end products by the methanogens allows the fermentation reactions involved in
refuse decomposition to proceed to methane production.60 If fermentative and methanogenic
activities are not balanced, intermediates will accumulate and may percolate from the landfill
as leachate. As the methanogens become active, these soluble organics are metabolized.
Thus, the chemical oxygen demand (COD) of landfill leachate is typically highest before
the onset of refuse methanogenesis.47

B. MICROBIOLOGICAL STUDIES OF THE REFUSE ECOSYSTEM


Much of the initial research on landfill microbiology focused on counts of fecal bacteria
562 Critical Reviews in Environmental Control

and pathogens.2°-23-33 Such work is useful for evaluation of potential health problems as-
sociated with refuse and landfill leachate, but does not elucidate the microbial processes
responsible for landfill raethanogenesis. In more recent studies, researchers have begun to
study characteristics of the refuse ecosystem indicative of biodegradation.6405253-76-77
Cook et al.20 isolated numerous bacteria and fungi from fresh and 3-year-old refuse. In
their study, fresh refuse was sampled from refuse collection vehicles, and old refuse was
excavated from the top layer of refuse at a sanitary landfill. The sampling procedures were
selective for aerobic and aerotolerant anaerobic microorganisms. Pseudomonads, which are
typically aerobic,14 were the major isolates in fresh refuse. Bacillus, Clostridium, Lacto-
bacillus, and Propionibacterium were the four genera isolated from fresh refuse that are
capable of anaerobic metabolism. Bacillus, Lactobacillus, and Propionibacterium were the
primary isolates from old, presumably anaerobic, refuse. Cook observed that most of the
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

isolates required amino acids and very few isolates grew on a medium containing salts,
ammonia, and glucose. Significant decreases were found in the populations of two enteric
organisms, E. coli and S. faecalis, between fresh and old refuse, with both enterics present
at relatively low levels of less than 100 cells per dry gram in 3-year-old refuse.
Riley et al. 78 reported an average total heterotrophic anaerobic population of 3.3 x 105
cells per milliliter in leachate from two locations at the same landfill. The cellulolytic
population ranged from 8.5 to 13.5 x 103 cells per milliliter. Of the genera identified by
Riley, only Bacillus was capable of anaerobic metabolism.
Filip and Kuster33 studied refuse decomposition in a 40-m3 test cell. The anaerobic
proteolytic population did not increase in 20 months relative to the value measured in fresh
refuse. The population of acid-forming and sulfate-reducing bacteria increased immediately
after initiation of the test cell. It was not stated whether the acid-forming bacteria were
grown under anaerobic conditions.
Donnelly and Scarpino23 studied the presence of pathogenic organisms in refuse and
their potential for being released from a landfill in leachate, which could potentially con-
taminate a drinking water supply. Populations of fecal coliforms #nd pathogens in leachate
were not present in sufficiently high numbers to be of concern. However, the refuse itself
contained pathogens, including Clostridium perfringens, at a population of 103 per gram of
refuse. Similar results were obtained at both a commercial landfill and in laboratory-scale
test cells. High numbers of fecal bacteria were found in 1-year-old refuse, but not in the
landfill leachate.
The total anaerobic population, enumerated on blood agar, was as high as 3.5 X 109
cells per gram of fresh refuse for samples from laboratory-scale lysimeters, but only 103
cells per milliliter for refuse samples taken from field-scale test cells and a commercial
landfill. Note that these values are not expressed in similar units. Methanogenic activity
was detected in samples at 1.5-, 2.1-, and 2.6-m depths in samples from a commercial
landfill. Little data on gas composition at the commercial landfill were reported.
Kinman et al.35 performed some microbiological analyses on 5-year-old refuse removed
from laboratory-scale test lysimeters. Clostridium perfringens was present at populations
between 103 and 107 cells per gram of dry refuse. However, these populations may be
misleading as the total clostridia population was reported as 0 to 104 cells per gram, and no
explanation for this apparent discrepancy was reported. The total anaerobic population was
measured at 103 to 10s cells per gram of refuse and was less than the number of C. perfringens
reported for several samples. Each of the six lysimeters sampled contained methanogens,
as evidenced by the presence of methane in the headspace of tubes incubated with 5 g of
refuse. Only two of the six lysimeters showed methane formation in samples from both the
top and bottom.
Rees et al., 76 Rees and Grainger,77 Jones et al., 33 Jones and Grainger,52 and Grainger
et al. 40 presented the first studies in which researchers attempted to measure the microbial
Volume 19, Issue 6 (1990) 563

populations and enzyme activities required for refuse methanogenesis. Jones and Grainger52
measured the activity of several hydrolytic enzymes in both saturated and dry refuse incubated
under laboratory conditions. The protease activity of saturated refuse increased rapidly and
was three times greater than the protease activity of dry refuse after 20 d. The 3-fold increase
in protease activity corresponded to a 30-fold increase in the most probable number (MPN)
of proteolytic bacteria after 20 d. However, protease activity decreased in both saturated
and dry refuse such that its activity on day 80, 5 \x.g Azocoll degraded per hour gram of
dry refuse, was the same in both samples. Changes in the protein concentration in the refuse
were not reported. Protein represents less than 8% of the methane potential of refuse (Table
1), and ammonia does not limit the onset of methane production.8 Thus, under the conditions
of leachate recycle and neutralization used by Barlaz et al., 4 protease activity probably did
not limit refuse methanogenesis.
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Amylase activity was initially lower in saturated refuse but increased to a level above
that of dry refuse after 20 d.52 There was little change in the MPN of the amylolytic bacteria
as the amylase activity increased. The starch concentration of the refuse used in this study
was 2.4%.
Cellulose is the principal biodegradable constituent of refuse under the anaerobic con-
ditions typical of landfills and its degradation in refuse is well documented.9-13 The cellulose
concentration of the refuse used by Jones and Grainger52 was 25.7% and cellulase activity
decreased rapidly after the first 10 d of the experiment. However, in attempts to validate
the procedure used for extraction of cellulase from refuse, it was found that the cellulase
was fully recovered from sterile refuse, but a fixed amount of enzyme activity was measured
in nonsterile refuse regardless of the amount of enzyme added.40 The possibility that the
cellulase enzyme was deactivated by protease was suggested by the authors. Both aerobic
and anaerobic bacteria are capable of cellulose degradation, and the fact that cellulase was
detected for the first 10 d of the experiment suggests that cellulase activity of aerobic
organisms only was detected. During the early stage of a refuse incubation, aerobic conditions
prevail. Once the oxygen present initially is depleted, anaerobic conditions prevail.
In later work, Jones et al. 53 characterized refuse samples from a sanitary landfill as a
function of depth below the surface. Refuse depth would generally be expected to correlate
with the age of the refuse, although this correlation will also depend on the manner in which
the landfill was operated. The total anaerobic population as well as the populations of
proteolytic, amylolytic, and cellulolytic bacteria increased near the water table, suggesting
a stimulatory effect of moisture content. Protease and amylase activity increased sharply in
the water table, which is consistent with the differences in enzyme activity between wet and
dry refuse measured under laboratory conditions.32 Both the protein and starch concentrations
decreased with depth in the landfill, indicating that the buried refuse was decomposing.
However, it is difficult to quantify protein decomposition by measurement of Kjeldahl
nitrogen, as used in this study. Kjeldahl nitrogen does not distinguish between organic
nitrogen present in the refuse and organic nitrogen present in microbial cell mass.
Cellulase activity decreased markedly with depth in the landfill, even below the water
table. Cellulase activity was more prevalent here than in the laboratory-scale tests, at least
over the top 3m. It was suggested that in the top several meters of the landfill the cellulase
activity may be expressed by fungi and that the extraction of fungal cellulase was more
complete than the extraction of bacterial cellulase. The authors suggested that the presence
of oxygen in the upper layers of the landfill could support the growth of fungi, which are
generally aerobic. Oxygen present near the surface of a landfill could support the growth
and activity of aerobic cellulolytic bacteria as well as fungi. However, gas composition data
were not reported.
The cellulose concentration was 30 to 40% in the top 5 m of the landfill, but decreased
to less than 4% at the 7.5-m level, suggesting complete decomposition of the available
564 Critical Reviews in Environmental Control

TABLE 3 TABLE 4
Relationship Between Refuse Depth, Methanogen Populations in
pH, and Methanogen Population17 Landfills

Sample Highest dilution with Methanogen population


depth (m) pH growth on Hj/CO2 highest dilution with
Depth growth on
0.9 5.9 103 (m) Hj/CO, Acetate
3.0 5.5 103
5.0 5.7 104 107
2—3 103
7.0 6.2 103 4 106 104
9.0 6.6 103 4 10' 10'
11.0 5.3 10' 7 103 104
101
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

12.5 5.6 7.3 106 104


1010 10*

From Fielding, E. T. and Archer, D. B.,


Effluent Treatment and Disposal, European
Federation of Chemical Engineering, Ser. No.
53, Pergamon Press, Elmsford, NY, 1986,
331. With permission.

cellulose. Jones et al. 53 did not explain why the lignin concentration remained the same or
decreased with depth. Its concentration would be expected to increase as cellulose decomposes13
since lignin is recalcitrant under anaerobic conditions.87 The trend in the lignin concentration
data of Jones et al. may also be explained by variability in the lignin concentration of the
refuse as buried.
Campbell et al. 17 measured the hydrogen plus carbon dioxide utilizing methanogen
population in refuse samples from a landfill (Table 3). Though qualitative, this is the first
report in which methanogen populations from landfills were measured. Data on total gas
production only were provided, so it is not possible to relate methane production rates to
the population data in Table 3. It is interesting to note that significant methanogen populations
are present at pH levels below those considered optimum for methanogen activity, 6.8 to
7.4. 88 Barlaz et al. 6 measured a four order of magnitude increase in the methanogen population
between refuse sampled at the beginning and end of the anaerobic acid phase of decomposition
(see Figure 2). The pH of the refuse in the anaerobic acid phase remained between 5.7 and
6.2.
The population of methanogens growing on acetate and hydrogen plus carbon dioxide
was measured in six landfill samples (Table 4) by Fielding and Archer.31 Data are reported
as the highest dilution exhibiting the production of methane, and population densities ranged
from 101 to 1010 cells per wet gram. There were typically 100-fold more cells capable of
growth on hydrogen and carbon dioxide compared to growth on acetate. Barlaz et al. 6 also
measured populations of acetate and hydrogen plus carbon dioxide utilizing methanogens
and reported that the two populations never deviated by more than tenfold. The lower ratio
of hydrogen plus carbon dioxide to acetate utilizing methanogens measured by Barlaz may
be a result of efforts to disrupt clumps of cells while diluting the refuse inoculum. Chartrain
et al. 19 reported increases in the most probable number of lactose utilizing bacteria when
efforts were made to disrupt clumps.
A technique for extraction from refuse of coenzyme F 420 , an electron carrier unique to
methanogenic bacteria, was reported by Fielding and Archer.31 The researchers felt that
direct enumeration of methanogens was not reliable as certain species, i.e., Methanosarcina
barkeri, have'a tendency to grow in clumps that are hard to enumerate. The use of F420
concentrations as a monitoring tool was proposed; however, the relationship between F420
concentrations and counts based on direct enumeration remains to be determined. Seven
Volume 19, Issue 6 (1990) 565

TABLES
Microbial Populations in a Sanitary
Landfill

Trophic group Range of population


(cells/wet g)«

Cellulolytics 104—10s
Hemicellulolytics 105—10'
Amylolytics 105—107
Methanogens
-Acetate 102—106
-HJCO2 105—10"
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

• Data are the range of populations in samples


taken between 0.61 and 4 m below the surface.

Adapted from Sleat, R. et al., Proc. ISWA Symp.


Process, Technol. Environ. Impact Sank. Landfill,
October 20—23, Cagliari, Sardinia, Italy, 1987.

methanogens were isolated from landfill samples and classified to the level of genus. The
isolates included one Methanosarcina, one Methanogenium, and five Methanobacterium
species. All were mesophilic with temperature optima of about 40°C.
In later work, Fielding et al.32 characterized seven methanogenic bacterial isolates from
landfill samples. Using antigenic fingerprinting, the isolates were identified as Methano-
bacterium formicicum, Methanosarcina barken, Methanobacterium bryantii and an uniden-
tified coccus. The optimum pH for all isolates was 7.0 and the optimum temperature was
between 37 and 41CC, which is consistent with the work of Pfeffer,70 Hartz,50 and Mata-
Alvarez and Viturtia.65
The populations of several trophic groups of organisms were measured as a function of
depth in a sanitary landfill.80 Population data are summarized in Table 5. The highest
population of hydrolytic bacteria corresponded with the location of the maximum gas pro-
duction rate, although methane concentrations were not reported. The populations of acetate
and hydrogen plus carbon dioxide utilizing methanogens were higher in more thoroughly
decomposed refuse. The sample that exhibited maximum gas production had a total car-
boxylic acid concentration of at least 30,000 mg/1, although the precise concentration was
not reported.
There has been only one report of a cellulolytic microorganism isolated from a municipal
solid waste landfill.2 The isolate was not previously identified and was named Cellulomonas
fermentans. It is capable of growth under both aerobic and anaerobic conditions, but with
no apparent advantage to aerobic growth. Its major fermentation products under both aerobic
and anaerobic conditions are acetate, formate, and ethanol. The minimum vitamin require-
ments of the organism are for biotin and thiamine, and better growth is exhibited with 0.05%
yeast extract. Optimum growth was at temperatures between 30 and 37°C and at pH 7.4.
Growth was significantly inhibited below pH 7 and above pH 8. Inhibition of cellulose
hydrolysis at a pH of less than seven is useful to refuse methanogenesis. If the pH of the
refuse ecosystem is acidic, it is probably due to an accumulation of volatile fatty acids
arising from cellulose hydrolysis. Additional cellulose hydrolysis while the pH is acidic
would only exacerbate the acidic pH.
Courts et al.21 studied the conversion of hexanoate to methane in multistage chemostats.
The inoculum was a hexanoate-enriched culture from a landfill. It was concluded that there
were four principal groups of microbes required for the conversion of hexanoate to methane.
These groups were the hexanoate-catabolizing proton-reducing acetogens, sulfate reducers,
566 Critical Reviews in Environmental Control

hydrogen-utilizing acetogenic, and methanogenic bacteria. In the presence of excess sulfate,


the sulfate-reducing bacteria utilized a majority of the hydrogen produced from hexanoate
oxidation. The methanogenic and hydrogen-utilizing acetogenic bacteria became more im-
portant as the sulfate concentration became limiting.
Beeman and Suflita11 studied the microbial ecology of an unconfined aquifer below a
sanitary landfill in Oklahoma. Two distinct ecosystems were identified. In a sulfate-rich
area the sulfate-reducing bacteria outcompeted the methanogens for carbon, and in a sulfate-
poor area methanogens predominated. Acridine orange direct counts indicated about 107
cells per dry gram of aquifer soil, including between 102 and 103 methanogens and sulfate
reducers per dry gram. These populations were enumerated in aquifer material below and
adjacent to a landfill and cannot be compared to counts in refuse. It is unclear whether the
methanogenic and sulfate-reducing bacteria migrated from the refuse or were present in the
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

aquifer initially.
Barlaz et al. 6 measured changes in the total anaerobic population as well as the popu-
lations of cellulolytic, hemicellulolytic, acetogenic (based on butyrate catabolism), and
acetate and hydrogen plus carbon dioxide utilizing methanogenic bacteria between the ini-
tiation of refuse incubation and the methane production phase. Concurrently, gas composition
and production as well as changes in the chemical composition of refuse were measured.
They worked with shredded refuse and enhanced decomposition with leachate recycle and
neutralization in laboratory-scale reactors.
In the first part of the study, methods for formation of a liquid inoculum from shredded
refuse were developed.7 Elaborate inoculum formation techniques, including prechilling of
the refuse and multiple extractions, were compared to a simple inoculation procedure that
involved blending the refuse in anaerobic phosphate buffer and formation of a hand-squeezed
extract. The more elaborate inoculum formation procedures did not result in significantly
higher MPNs of anaerobic cellulolytic bacteria, and the simpler procedure was adopted.
There was no evidence for the presence of anaerobic, cellulolytic fungi in the one refuse
sample tested.
Population increases of 2, 4, 5, 5, and 6 orders of magnitude were measured between
fresh refuse and the methane production phase for the hemicellulolytic, cellulolytic, butyrate
catabolizing acetogenic, acetate, and hydrogen plus carbon dioxide, utilizing methanogenic
bacteria, respectively. Seventy two percent by weight of the initial cellulose plus hemicel-
lulose was degraded in a container sampled after 111 d. Mass balances showed that the
mass of nitrate and oxygen entrained in a container initially were sufficient for oxidation to
carbon dioxide of only 2% of the initial sugar concentration of the refuse. This, combined
with the low acid-consuming activity of the acetogenic and methanogenic bacteria and the
low alkalinity of fresh refuse, explained the observed pH decrease in the refuse ecosystem
from 7.5 to 5.7. Initially, acetate utilization, but ultimately polymer hydrolysis, limited
methane production. The cellulolytic and acetogenic populations did not increase until after
the onset of methane production. Prior to the onset of measurable methane production, while
the pH of the refuse was between 5.7 and 6.2, there was a four order of magnitude increase
in the methanogenic population. The microbial population, gas production, and chemical
constituent data are presented in Figure 2 as a four-phase characterization of refuse decom-
position and are discussed in the following section.
Barlaz et al. 8 reported on inhibition that appeared to be characteristic of refuse decom-
position with leachate recycle. Potential causes, including an absence of microorganisms,
toxic carboxylic acid, and cation concentrations, and an ammonia limitation, were discounted
from the available data. Acetate and butyrate concentrations of 9753 and 6956 mg/1, re-
spectively, did not inhibit methane production and neither ammonia, phosphate, nor sulfur
limited the onset of methane production.
Barlaz et al. 10 compared microbial populations in refuse actively producing methane,
Volume 19, Issue 6 (1990) 567

Aerobic Anaerobic Accelerated Decelerated


100 Methane Methane
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

90 111
2027 3441 48
Day

FIGURE 2. Summary of observed trends in refuse decomposition with leachate recycle. Gas volume
data were corrected to dry gas at standard temperature and pressure. The acids are expressed as acetic
acid equivalents. Solids remaining are the ratio of the cellulose plus hemiccllulose removed from a
container divided by the weight of cellulose plus hemicellulose added to the container initially. Meth-
anogen MPN data are the log of the average of the acetate- and H^COj-utilizing populations. (Reprinted
from Barlaz, M. A. et al., Appl. Environ. Microb., 55(1), 55, 1989. With permission.)

incubated with and without leachate recycle and neutralization. Refuse methanogenesis was
stimulated either by leachate recycle and neutralization, or by inoculating fresh refuse with
old, anaerobically degraded refuse at the time of initial incubation. There were no significant
differences in the total anaerobic population or the subpopulations of cellulolytic, hemicel-
lulolytic, acetogenic, or methanogenic (based on acetate or hydrogen plus carbon dioxide
utilization) bacteria between methane-producing refuse, with and without leachate recycle.
The fraction of the total anaerobic population represented by each trophic group in the
inoculated containers was similar to the corresponding fraction in two leachate recycle
containers. It was concluded that leachate recycle may be used to accelerate refuse decom-
position in the laboratory without changing the microbial composition of the refuse ecosystem.
Tibbies and Baecker83 studied population shifts after the addition of phenol to 8-year-
old refuse removed from a landfill and placed in a 0.1 m3 lysimeter. Most isolates from the
568 Critical Reviews in Environmental Control

lysimeter to which phenol was added were capable of phenol degradation, while few species
isolated from the lysimeter to which no phenol was added could degrade phenol. No anaerobic
phenol-utilizing bacteria were isolated. Watson-Craik and Senior84 also reported on the
conversion of phenol to methane in laboratory-scale lysimeters. Leachate recycle was nec-
essary for complete degradation of the 188-mg/l phenol solution.
While the early work on refuse microbiology emphasized populations of fecal indicator
organisms and aerobic bacteria, later work has focused on the anaerobic bacteria responsible
for refuse methanogenesis. As the base of knowledge on refuse decomposition has expanded,
researchers have tried to interrelate gas concentration and production data with leachate
quality and changes in the chemical composition of the refuse. The summaries of refuse
decomposition reviewed in the next section represent the best available synthesis of infor-
mation at a given time.
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

C. DESCRIPTIONS OF REFUSE DECOMPOSITION


There have been numerous attempts to characterize refuse methanogenesis in terms of
different phases of decomposition between the burial of fresh refuse and well-decomposed
refuse. The first such characterization was presented in 1973 by Farquhar and Rovers30 and
it has since been updated by Rees, Pohland, Ehrig and Barlaz. Farquhar and Rovers relied
almost solely on gas composition data to characterize refuse decomposition. The first or
aerobic phase was defined by the depletion of oxygen with carbon dioxide production. In
the anaerobic, nonmethanogenic phase (2), there was high carbon dioxide production, some
hydrogen production, and some nitrogen production due to denitrification. Methane was
first detected in phase 3, according to Farquhar, and its concentration increased to some
terminal value. Phase 3 was termed the unsteady methanogenic phase. Hydrogen was depleted
in phase 3, although there was a lack of knowledge about the importance of hydrogen in
the process of anaerobic decomposition at the time of Farquhar and Rovers's work. The
fourth phase of refuse decomposition was characterized by a steady methane concentration
of at least 50%.
Rees73 added carboxylic acid and cellulose concentrations to the description of refuse
decomposition developed by Farquhar. Rees suggested that the concentration of carboxylic
acids should reach a peak and then decline with the initial detection of methane in the gas
phase. Rees also showed a linear decrease in the cellulose concentration beginning with the
time of initial burial. He proposed a slight increase in the rate of cellulose decomposition
long after the depletion of the carboxylic acids. Rees also added a fifth phase to the four-
phase scheme proposed by Farquhar. In the fifth phase, all of the cellulose was consumed
and the methane concentration decreased as air diffused into the landfill.
Pohland74 proposed a five-phase scheme for refuse decomposition. In the first phase,
the low moisture content of the refuse limits microbial activity and leachate and gas pro-
duction. Measurable microbial activity is observed in the transition phase, 2, and the landfill
shifts from an aerobic to an anaerobic ecosystem as oxygen is depleted. Phase 3, termed
the acid formation phase, is characterized by the production of carboxylic acids and an acidic
leachate. The onset of methane production is observed in phase 4, the methane fermentation
phase. Concurrent with methane production is a reduction in carboxylic acid concentrations
and leachate strength, a pH increase and low oxidation-reduction potentials. In phase 5,
termed the maturation phase, methane production decreases as less available constituents
are degraded.
Ehrig24 described refuse decomposition in four phases (Table 6) that emphasized leachate
characteristics instead of gas production. Refuse decomposition was observed under con-
ditions of leachate recycle, and the refuse was seeded with composted refuse.
The most recent description of refuse decomposition was presented by Barlaz et al. 6
Characterization of refuse decomposition in four phases is presented in Figure 2 and sum-
Volume 19, Issue 6 (1990) 569

TABLE 6
Characterization of Refuse Decomposition According to Ehrig24

- Phase Characteristics

1 High strength leachate, pH less than 6.5, increasing methane concentration


with low methane production
2 High methane concentration and pH, high strength leachate
3 Stable pH, high methane production, maximum methane concentration,
decreased leachate strength
4 Constant pH, low leachate strength, low methane production, and even-
tually a decrease in the methane concentration
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

marized here. In the aerobic phase (phase 1), both oxygen and nitrate are consumed with
soluble sugars serving as the energy source for microbial activity. All of the trophic groups
required for refuse methanogenesis are present in fresh refuse (cellulolytics, acetogens, and
methanogens), although there is little change in their populations. In the anaerobic acid
phase (phase 2), carboxylic acids accumulate, the pH decreases, and there is some cellulose
and hemicellulose decomposition. The methanogen population increases, and methane is
detected in the landfill gas. In phase 3, the accelerated methane production phase, there is
a rapid increase in the rate of methane production to some maximum value. Characteristic
of this phase is a methane concentration of 50 to 60%, a decrease in carboxylic acid
concentrations, an increase in the pH of the ecosystem, little solids hydrolysis, and increases
in the populations of cellulolytic, acetogenic, and methanogenic bacteria. The fourth phase
is termed the decelerated methane production phase. The methane concentration, pH and
cellulolytic, and methanogenic populations remain at levels similar to those in phase three.
Concurrently, the methane production rate decreases, the acetogen population increases,
carboxylic acids are depleted, and there is an increase in the rate of cellulose plus hemi-
cellulose hydrolysis.

HI. EFFECTS OF ENHANCEMENT VARIABLES ON METHANE


PRODUCTION
Research on the decomposition of solid waste in sanitary landfills was first reported by
Merz and Stone,68 and since that time numerous researchers have tried to enhance refuse
methanogenesis by manipulation of the landfill ecosystem. Research on enhancement of
methane production preceded research on landfill microbiology. With the advances in landfill
microbiology presented in the previous section, it is appropriate to reanalyze previous efforts
to enhance methane production. Parameters that have been studied to promote methane
production include moisture content, particle size, sludge addition, and leachate recycle.
The effects of these and other parameters are discussed in this section. Areas in which
studies have provided conflicting data are presented and, where possible, resolved. The
research covered herein is restricted to work in which refuse decomposition was studied
under conditions approximating a sanitary landfill. Studies in which refuse was degraded in
well-mixed anaerobic digesters37'8870-71-91 are not reviewed here. Halvadakis et al. 44 compiled
a list comparing methane yields from digester and refuse lysimeter studies. Yields from
digesters, in which conditions for methane production are optimized, tend to be at least
twice those of lysimeters. However, there is so much variability in lysimeter data that
definitive statements are inappropriate. Barlaz et al.9 and Buivid et al.15 both worked with
leachate recycle and neutralization and reported methane yields similar to those reported for
digester studies.

A. MOISTURE CONTENT
Merz and Stone67 concluded that higher moisture contents stimulated methane production
570 Critical Reviews in Environmental Control

in field-scale test cells (1376 m3) filled with nonshredded refuse. Moisture contents in the
range of 26 to 52% were tested. The time to the onset of methane production and the volume
of gas produced were not reported.
Emcon27 and Leckie et al.61 reported the effects of moisture on refuse decomposition in
field-scale test cells (770 m3) in Sonoma County, CA. Methane was not detected after 2
years in a cell initiated at field capacity (45.2%), the highest moisture content tested.
However, the gas in this cell contained 20% CH4 by volume after 2.5 years. In a second
cell, where the moisture content was increased from an initial value of 25.6%, to about
60%, the methane concentration increased to 15% within 6 months and 50% in 2 years. It
was concluded that adjusting the refuse to field capacity initially or continuous flow of water
through refuse accelerated decomposition.
Wujcik and Jewell86 studied the effect of moisture content on the batch fermentation of
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

wheat straw and dairy manure. Methane yields decreased at moisture contents below 70%
and the yield at 30% moisture was 22% of the yield at 70% moisture. Fresh refuse typically
contains 20% moisture.82
Barlaz et al.4 concluded that higher moisture contents (55%) in shredded refuse lead to
more rapid acid accumulations. However, methane production was not always observed at
low moisture contents (35%). Halvadakis et al.43 found moisture content to be one of the
most critical variables controlling gas production in field-scale test cells. Using a modified
biochemical methane potential assay, Bogner et al. I2a stimulated methane production by
increasing the moisture content of various refuse samples.
Jones and Grainger52 reported higher protease and amylase activities in saturated refuse
compared to dry refuse, although no differences could be detected with respect to cellulase
activity. Since the methane potential of protein and starch in municipal refuse is about 10%
of that of cellulose (Table 1), it is not possible to conclude from these data that higher
moisture contents stimulate refuse decomposition.
The results of Barlaz et al. 4 are not consistent with other studies presented. 276167 With
the exception of Barlaz, others interpreted their results based on gas concentration, which
is not always a reliable indicator of methane production.4 The unprocessed refuse used by
Merz and Stone and Leckie was compacted to densities of 177 to 338 kg/m3 compared to
469 to 613 kg/m3 in the work of Barlaz, who worked with shredded refuse. The discrepancy
between these studies may be explained by consideration of the effects of moisture, density,
and combinations thereof, on refuse decomposition.
In a heterogeneous system such as a landfill, the opportunity for contact between mi-
croorganisms, their substrates, and other necessary growth factors may limit biodegradation.
As the refuse moisture content in increased, the opportunity for contact is increased, which
should enhance microbial activity. Halvadakis et al.44 believe that in most landfills, the
hydrolysis rate is slow enough that the rates of acid production and consumption are balanced.
If the hydrolysis rate was high or stimulated to too great an extent, or if a great deal of
soluble substrates were available, fermentative organisms would grow at a much faster rate
than the methanogens and there would be increased potential for acid accumulation and a
pH decrease.
At higher densities each particle of refuse is in closer contact with adjacent particles.
Thus, for the same moisture content, there will be more water to refuse contact per unit
volume of landfill as the density increases. If moisture contact stimulates the rate of hydrolysis
arid too high a rate of hydrolysis is inhibitory, then as the density increases the optimum
moisture content would be expected to decrease. Shredded refuse may be compacted to
higher densities than unprocessed refuse. This explains why the optimal moisture content
determined by Merz and Stone,67 Emcon,27 and Leckie et al.61 was not applicable for Barlaz
et al., 4 who worked with shredded refuse.
The broadest data sets where moisture content can be evaluated are those of Emberton26
Volume 19, Issue 6 (1990) 571

and Jenkins and Pettus.31 Emberton evaluated methane production rate data for landfills
across the U.S. and categorized the landfills based on annual precipitation. Jenkins tested
the effect of moisture content in refuse sampled from landfills. In both studies, the methane
production rate exhibited an upward trend with increasing moisture contents, confounding
factors such as density, refuse age, and refuse composition notwithstanding.

B. LEACHATE RECYCLE AND MOISTURE FLOW


The flow of moisture through a landfill may be expected to stimulate microbial activity
by providing better contact between insoluble substrates, soluble nutrients, and the micro-
organisms. If the rate of polymer hydrolysis is stimulated and the acetogen and methanogen
populations cannot metabolize the end products of the hydrolytic and fermentative micro-
organisms, carboxylic acids will accumulate. This will cause the pH of the refuse to decrease,
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

unless it is artificially buffered. Assuming the refuse can conduct adequate volumes of water,
the ability of increased water flow, as used with leachate recycle, to stimulate methane
production will depend on balanced activity among the populations of hydrolytic, acetogenic,
and methanogenic bacteria.
Pohland,72 working with shredded refuse, showed leachate recycle to be stimulatory,
and leachate recycle plus neutralization further stimulated decreases in leachate carboxylic
acid concentrations. Methane production was inferred from the decreases in carboxylic acid
concentration. In later work, Pohland173 observed very little methane production from refuse
with leachate recycle and no neutralization.
Buivid et al.15 showed that leachate recycle was most stimulatory when anaerobic sludge
and calcium carbonate were added to the refuse initially. The sludge served as an inoculum
and the calcium carbonate was intended to neutralize the refuse and sludge. The calcium
carbonate plus sludge addition was typically 10% of the refuse dry weight, although the
proportion of sludge to calcium carbonate was not specified. Leachate recycle was also
stimulatory in refuse to which calcium carbonate but no sludge was added. However,
enhanced refuse methanogenesis was not observed with leachate recycle and no calcium
carbonate addition. Buivid used shredded refuse with a paper content of 77%, which is
significantly greater than the typical value given in Table 1 (41%). The high paper content
probably decreased the initial soluble sugar concentration associated with food waste. This
in turn decreased the potential for a rapid accumulation of carboxylic acids from sugar
fermentation. Nevertheless, the enhancement attributed to calcium carbonate addition sug-
gests that low pH limited methane production. Tests were performed in 0.056 m3 (15 gal)
reactors at a density of 406 kg/m3.
Barlaz et al. 4 ' 8 concluded that leachate recycle without neutralization stimulated the
accumulation of carboxylic acids. Daily leachate neutralization stimulated methane produc-
tion with depletion of carboxylic acids. Mata-Alvarez and Viturtia65 showed that leachate
recycle stimulated methane production in shredded refuse to which pig manure and calcium
carbonate were added. The cellulose plus hemicellulose fraction of the refuse used by Mata-
Alvarez (35%) and the low lignin content (3%) were unusual. Kinman et al.56 evaluated
leachate recycle alone and in test reactors to which a calcium carbonate buffer had been
added at the time of reactor loading. Where added, the calcium carbonate was 18.5% of
the refuse dry weight. Leachate recycle did not enhance methane production under either
condition.
Halvadakis et al. 43 concluded that leachate recycle increased the initial rate of gas
production in a field-scale test cell that included anaerobic sewage sludge and calcium
carbonate. Both Buivid and Kinman used calcium carbonate to neutralize refuse. While
calcium carbonate was effective for Buivid, it did not enhance methane production in
Kinman's work. Barlaz et al.3 observed that the leachate pH from refuse to which calcium
carbonate was added at the time of filling never exceeded 5.0, and there was no methane
production.
572 Critical Reviews in Environmental Control

Klink and Ham58 concluded that moisture flow increased the rate of methane production
by 25 to 50% relative to refuse at the same moisture content with no moisture flow. They
worked with refuse in an active state of methane production. Rovers and Farquhar30 monitored
methane concentration and leachate strength in field cells that received either controlled
water addition or naturally occurring rainfall. The cell that received the least moisture
produced leachate with the lowest COD. Apparently, refuse decomposition in this cell was
moisture limited, as fermentative activity and the associated high COD leachate is typically
observed even when methanogenesis is inhibited. In field-scale work with unprocessed
refuse, Leckie et al.61 showed that leachate recycle was stimulatory, even without leachate
neutralization. However, only gas composition and not gas production was measured.
In summary, Barlaz and Buivid are similar in describing the importance of leachate
neutralization. Buivid and Mata-Alvarez both observed rapid refuse decomposition under
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

conditions of leachate recycle with additions of an inoculum and calcium carbonate. Although
Pohland73 appeared to have stimulated methane production with leachate recycle in the
absence of neutralization, the reported methane yield indicates very little conversion of refuse
to methane. Even with buffer addition, leachate recycle was not stimulatory in Kinman's
study. The existing data suggest that leachate recycle and neutralization will usually stimulate
refuse methanogenesis from fresh refuse. In the absence of neutralization, leachate recycle
does not appear to stimulate methane production from fresh refuse. Moisture flow alone
stimulated methane production from refuse in an active state of methane production.

C. PARTICLE SIZE
DeWalle et al., 22 working with shredded refuse at 44% moisture in 0.21 m3 (55 gal)
drums, suggested that a large particle size (25 cm) stimulated methane production relative
to drums with refuse particle sizes of 2.5 and 12.5 cm. However, the average rate of methane
production in the supposedly stimulated drum was only 0.0005 m3 CH4/kg per year, which
is well below the rates reported by others for enhanced methane production.444 Using
shredded refuse and 0.23 m3 containers, Fungaroli and Steiner36 concluded that decreasing
particle size stimulated the rate at which the maximum methane concentration was achieved.
Buivid et al.15 found that refuse with a 25- to 35-cm particle size produced 32% more
methane after 90 d than refuse with a 10- to 15-cm particle size. Similarly, in a comparison
of refuse with particle sizes of 10 to 15 cm and 1.25 to 2.5 cm, the refuse with the larger
particle size produced 16 times more methane after 90 d. The refuse was inoculated with
sludge in all tests. The density for all four containers mentioned above was 406 kg/m3 and
the moisture content was 75%.
Ham42 found that the onset of methane production, as measured by gas concentration,
was stimulated by shredded refuse (90% by weight passing a 7.5-cm sieve) relative to
unprocessed refuse in 204-m3 field cells covered with soil.
The use of refuse with a reduced particle size relative to unprocessed refuse infers some
mixing of the refuse associated with the size reduction operation. The well-mixed, shredded
refuse permits greater contact between the key refuse constituents required for methane
production: moisture, substrate, and microorganisms. Thus, a smaller particle size could
result in an increase in the rate of hydrolysis. Under the conditions used in the work of
Ham, the smaller particle size did not cause stimulation of the hydrolysis rate to the point
where the refuse soured. While Ham concluded that a smaller particle size was stimulatory,
Buivid concluded the opposite. This may be explained by the range of particle sizes tested
by Buivid (1 — 2 cm, 10 — 15 cm, and 25 — 35 cm), while Ham compared unprocessed
refuse with a 7.5-cm particle size refuse. Other differences between Ham and Buivid include
the use of a sludge inoculum by Buivid and higher moisture content and density used by
Buivid relative to Ham, who relied on rainwater infiltration. The higher density should, by
itself, stimulate hydrolysis as previously explained. Thus, further stimulation by a reduced
Volume 19, Issue 6 (1990) 573

particle size could cause too rapid a rate of hydrolysis, leading to a build-up of acidic end
products and a lower pH.

D. INOCULUM ADDITION
Anaerobically digested sewage sludge can serve as a seed of microorganisms as well as
a source of nitrogen, phosphorus, and other nutrients. Barlaz et al. 6 concluded that all of
the organisms required for refuse methanogenesis were present in fresh refuse. Barlaz also
reported that acids accumulate and the pH decreases due to an imbalance between fermen-
tative and methanogenic activity. Once the pH decreases, there is little mechanism within
a sanitary landfill for refuse neutralization. Several researchers have evaluated the use of
sludge to bring fermentative and methanogenic activity into balance, which would prevent
the initial pH decrease. When such sludge is added to refuse to enhance methane production,
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

the effects of seed and nutrient addition are confounded. In Buivid et al's. study15 there
were 26 reactors, and methane production was nil or low in 4 of the 5 reactors in which
there was no sludge inoculum. In addition, each reactor with sludge inoculum contained a
calcium carbonate buffer. Augenstein et al.1 also stimulated refuse methanogenesis by ad-
dition of sludge and calcium carbonate. Mata-Alvarez and Viturtia65 observed rapid refuse
decomposition in containers seeded with pig manure. In addition to pig manure, calcium
carbonate was added to the refuse and leachate was recycled. Sludge addition without buffer
addition did not stimulate methane production for Barlaz et al.4 Pohland's72 data also suggests
that sludge may stimulate an accumulation of carboxylic acids.
Leckie et al. 61 found the addition of septic tank bottoms to a field-scale test cell stimulated
acid production that inhibited methane production due to a pH decrease. Kinman et al. 56
found that the addition of sludge to 3-year-old shredded refuse was stimulatory.
The addition of anaerobic sewage sludge to fresh refuse has not been shown to enhance
methane production by itself. Both Barlaz et al. 4 and Pohland72 observed carboxylic acid
accumulations and decreases in pH associated with sludge addition to fresh refuse. This
observation suggests that the hydrolytic and fermentative populations in the sludge inoculum
are either better able to survive in or adapt more rapidly to, the refuse ecosystem, than are
the methanogenic bacteria. The hydrolytic and fermentative activity results in an accumu-
lation of carboxylic acids, which is not immediately consumed by the methanogens; con-
sequently, the pH decreases. The addition of sludge and buffer has been more successful
in promoting the onset of methane production from fresh refuse.1565 Perhaps the added buffer
holds the pH of the refuse ecosystem near neutral, allowing the methanogenic bacteria from
sludge to acclimate to the refuse ecosystem more rapidly than in the absence of buffer.
There are insufficient data to determine whether there is an advantage to the addition of
anaerobic sewage sludge to fresh refuse in which buffer is also added. Barlaz et al. 6 showed
that all of the trophic groups required for methane production are present in fresh refuse.
This suggests that conditions for growth, such as pH, are more important than inoculum
addition.
Another potential inoculum for fresh refuse is old refuse. Barlaz et al.4 stimulated the
onset of methane production in lysimeters inoculated with old, anaerobically degraded refuse,
relative to control containers containing fresh refuse only.
Old refuse could stimulate methane production by acting as a diluent against the ac-
cumulation of toxic compounds in the refuse ecosystem. Although Barlaz et al. 4 tested this
hypothesis, no firm conclusions on the effect of old refuse as a diluent were possible.
Steggman81 showed the addition of 2-month-old refuse to fresh refuse stimulated methane
production. It is not known whether this refuse contained a methanogen population signif-
icantly higher than that present in fresh refuse. The 2-month-old refuse may have been
stimulatory because of its effects as a diluent.
574 Critical Reviews in Environmental Control

E. pH
Kasali et al.54 evaluated the effect of pH on refuse methanogenesis by addition of a
phosphate solution (0.2 M) buffered to pH levels between 5.3 and 8.3 to refuse sampled
from a landfill. The refuse tested was greater than 1 month old at the time of sampling, and
it is not known whether it was in an active state of methane production at that time. It was
reported that the buffer stimulated acid production, causing large accumulations of carboxylic
acids that inhibited methane production. Although the pH was buffered, the final pH differed
from the initial pH by one or more pH units in three of the five lysimeters.
On analysis of refuse sampled from ten or more landfills, pH has been found to be the
strongest predictor of the methane generation rate.79 To our knowledge, the work by Kasali
et al.54 is the only study in which the effect of pH on refuse methanogenesis has been studied
in a controlled experiment. It is generally accepted that the pH optimum for methanogenic
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

bacteria, 6.8 to 7.4, 88 is applicable to the methane production rate in the refuse ecosystem.

F. NUTRIENT ADDITION
Filip and Kuster33 tested whether ammonia, glucose, ammonia plus glucose, or peptones
could increase microbial activity in refuse. Glucose, peptones, and ammonia plus glucose
stimulated the carbon dioxide evolution rate, whereas ammonia alone had no effect, indicating
that availability of carbon but not nitrogen limited microbial activity. Methane evolution
was not measured.
Barlaz et al. 8 concluded that ammonia, phosphate, and sulfur did not limit the onset of
methane production and that the concentrations of these nutrients present in the accelerated
methane production phase supported a methane production rate of at least 929 1 of methane
at standard temperature and pressure per kilogram of dry refuse per year. However, it was
observed that ammonia and phosphate were nearly depleted late in the decomposition process,
and more research is needed to determine whether nutrients limit methane production in the
final stage of refuse decomposition. Bogner et al.12a found nutrients to stimulate certain
refuse samples using modified biochemical methane potential assays. However, simple
moisture was nearly as stimulatory in most cases.

G. TEMPERATURE
The optimum temperature for methane production in the mesophilic range has been
reported as 41 and 42°C by Hartz et al. 30 and Pfeffer,70 respectively. Whereas Hartz worked
with laboratory-scale lysimeters, Pfeffer's experiments were conducted in completely mixed
reactors. Mata-Alvarez and Viturtia65 found the maximum methane production rate occurred
at 42°C, although maximum cumulative methane production occurred at 34 to 38°C. Leckie
et al. 6t observed temperatures of about 20°C in the middle of test cells in California. Temporal
variation in landfill temperature reflected variation in ambient temperature with some lag
and damping effect.
The optimum temperature for thermophilic refuse decomposition is at least 60°C.70
Temperatures above 60°C were not tested by Pfeffer.

H. SUMMARY COMMENTS
The effects of seven variables were discussed in this section, including

1. Moisture content
2. Leachate recycle and moisture flow
3. Particle size
4. Inoculum addition
5. pH
6. Nutrient addition
7. Temperature addition
Volume 19, Issue 6 (1990) 575

Moisture content26-51 and pH79 were the variables that best correlated with the .methane
production rate in refuse samples from sanitary landfills. The onset of methane production
from fresh refuse was most consistently enhanced by leachate recycle and neutralization4'8-65'73
and addition of old, anaerobically degraded refuse to fresh refuse.4 Moisture flow was also
shown to stimulate methane production from refuse already in an active state of methane
production.58
There are insufficient data to identify an optimum particle size for refuse decomposition.
As discussed, the optimum particle size can be expected to vary with moisture content.
Anaerobic sludge addition with buffer consistently stimulated methane production from fresh
refuse.15 Without buffer, sludge addition was not stimulatory.415 Nutrients do not limit the
onset of methane production from refuse and the nutrients available in refuse support sig-
nificant rates of methane production.8
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

It has been possible to explain many of the inconsistencies in the conclusions of different
researchers by examination of their experimental conditions. Many researchers have eval-
uated their results on the basis of methane concentration. This is not as accurate an indicator
of enhancement as is the methane production rate, which only recently has become a standard
measurement.
Research on the microbiology of refuse decomposition reviewed in Section II, and the
work on enhancement of refuse methanogenesis discussed in this section, have as their goal
an improved understanding of refuse decomposition and, hopefully, improved methane yields
from sanitary landfills. Increased methane yields will improve the economic feasibility of
sanitary landfills as an energy source. In the next section methods to assess the methane
potential of refuse buried in landfills are reviewed.

IV. ASSESSMENT OF METHANE PRODUCTION POTENTIAL OF


SANITARY LANDFILLS
In order to evaluate the economic feasibility of a landfill gas recovery project, it is
necessary to estimate both the rate of methane production and the total volume of methane
that can be expected from the landfill. The few data that are available on methane yields
from full-scale landfills46 indicate that methane yields are between 1 and 50% of the yields
calculated from stoichiometry.44 In addition, it is difficult to extrapolate methane production
rate data from laboratory lysimeters to field-scale landfills. Thus, it is best to measure
methane production rates in full-scale landfills wherever possible. Segal79 presented the
program used by GSF Energy, Inc. for evaluation of landfill gas-recovery projects. Current
practice is to estimate the rate of methane production by pump or drawdown tests. Cumulative
methane production can be estimated from mass balances on samples of buried refuse or
from theoretical models. Methods to measure gas production rates in landfills, the use of
mass balances, and efforts to model landfill gas production are reviewed in this section.

A. TESTS TO MEASURE GAS PRODUCTION RATES FROM LANDFILLS


Pump testing is the most widely used method to assess gas flow rates at full-scale
landfills. The concept is to pump gas from the site, measuring the volume and composition
of the gas, and then to use pressure sensing probes to determine the volume of the landfill
affected by pumping. This is, presumably, the volume of refuse producing the measured
gas. Multiple wells or trenches are used to withdraw the gas from the entire site to get total
gas flow rates, or results from the portion of the site affected by pumping can be used to
project total site flow rates.
Lu and Kunz63 developed a model to predict methane production from sanitary landfills
based on field measurements of changes in landfill gas pressure caused by pumping gas
from the landfill. Measurement of landfill gas pressure at three points radially outward from
576 Critical Reviews in Environmental Control

a withdrawal well are required to use the model. The model also calculates the fraction of
gas produced in a landfill that can be recovered from a withdrawal well as opposed to gas
that migrates vertically from the landfill through the cover. Lu and Kunz found that at the
Fresh Kills landfill in New York the horizontal conductance for gas flow was 37.5 times
greater than the vertical conductance. This ratio of horizontal to vertical conductance is
favorable for gas recovery in that landfill gas moves freely toward withdrawal wells (hor-
izontally), while loss of gas by vertical movement through the landfill cover is relatively
low.

1. Design of a Pump Test


Gas may be withdrawn by either a well or trench. If a well is used, then a vertical pipe,
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

perforated at the bottom, is placed in the landfill. Typically, a 15-cm pipe, perforated over
one to two thirds of the depth of the landfill, is placed in a hole 0.7 to 1.0 m in diameter.
The perforated depth is backfilled with gravel and the remainder of the annular space is
backfilled with a sealant such as bentonite. The gas flow rate pumped from a landfill is
monitored as a function of the suction applied to the production well or trench. Pressure-
sensing probes are typically placed in 15-cm holes located radially outward at varying
distances from the production well. Often, multiple probes are placed at each distance to
permit averaging. The pressure-sensing probe is typically a 1-cm pipe located in the 15-cm
hole. It is surrounded by gravel over the slotted portion of the probe and a sealant over the
nonperforated portion.
Trenches may be installed during landfill construction and used in place of wells for
gas-testing purposes. A trench will consist of a horizontal section of perforated pipe placed
in gravel. Pressure-sensing probes are placed on either side of the trench. Pipe within 35 to
70 m from the sides of the landfill would be nonperforated in order to prevent air intrusion
if gas is withdrawn by pumping.
Emcon Associates28 presented information on the design of pumping tests using one or
more withdrawal wells. They reviewed the equations and assumptions used for prediction
of landfill gas recovery from pumping tests at two landfills.
While conducting a pumping test, it is important to check for indications of air flow
into the landfill.28 If the rate of gas withdrawal from a landfill exceeds the rate of gas
production, air will be drawn into the landfill, inhibiting anaerobic activity. Air intrusion
should be apparent from an increase in the nitrogen concentration of the recovered gas.
Oxygen may not appear in the recovered gas even if it is drawn into a landfill, as it can be
rapidly consumed by microbial activity. Other evidence of air intrusion would be an increase
in the ratio of carbon dioxide to methane as carbon dioxide will be produced in either aerobic
or anaerobic reactions, while methane is produced only under anaerobic conditions.
A given landfill includes refuse of different ages, as well as other heterogeneities. Thus,
it is advantageous to perform a pump test over as much of the landfill as possible.
The major problem in gas testing is defining the area of influence. Accurate measures
of the pressure prior to pumping, termed the static pressure, are critical. However, static
pressures change hourly, as well as from day to day and from location to location within a
landfill, reflecting barometric pressure changes, moisture changes at the surface of the
landfill, and heterogeneities within the landfill. A second problem with pump testing is the
accuracy of the pressure-sensing device. A third problem is to determine how long pumping
must continue before pressure readings are stable. The net result of the various problems
listed above is that there is a fair amount of uncertainty in interpreting pump test results.

2. Flux Boxes to Estimate the Rate of Gas Production


Flux box testing consists of placing a device over a portion of the landfill surface to
capture gas flow. The device can range in capture area from a few to several square meters,
Volume 19, Issue 6 (1990) 577

and can range from simple inverted cans with plastic bag collectors to large plastic sheets.
These devices are usually sealed by attachment to a metal ring pressed into the landfill cover
or by placement in a small trench dug around the test area that is filled with water.
One problem with flux boxes is nonuniform gas flow through the cover, so it is advisable
to use many test areas. Other problems include wind effects, changes in soil moisture content,
leaks and diffusion through capture devices, gas utilization by soil microorganisms, gas
dissolution, adsorption, and the effect of vegetation. Lu and Kunz63 reported that the gas
production rate predicted by their model matched well with the rate of gas released through
the landfill cover as measured with a flux box.

B. USE OF REFUSE CHEMICAL COMPOSITION TO ESTIMATE METHANE


POTENTIAL
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Pumping tests discussed in the previous section are useful for estimation of the rate of
gas production in a landfill and, by measurement of the gas concentration, for confirmation
that there is steady methane production. However, the other important quantity necessary
for economic evaluation of a landfill gas-recovery project, cumulative gas production, cannot
be estimated from pumping tests. In this section a method is described for estimation of the
methane potential of refuse using mass balances and chemical composition data.
The mass of methane that would be produced if all of a given constituent were converted
to carbon dioxide, methane, and ammonia may be calculated from Equation I. 69

CnHaObNc + [n - (a/4) - (b/2) + 3(c/4)]H2O - *


[(n/2) - (a/8) + (b/4) + 3(c/8)]CO2 +
[(n/2) + (a/8) - (b/4) - 3(c/8)]CH4 + cNH3 (1)

Cellulose and hemicellulose comprise 91% of the methane potential of fresh refuse (Table
1). These two constituents along with carboxylic acids will comprise at least 90% of the
methane potential of refuse in virtually any state of decomposition. Procedures are available
for measurement of these refuse constituents.6 Based on Equation 2, 373 1 of methane at
standard temperature and pressure would be expected for every kilogram of carbohydrate
degraded (CnH2nOn).
The methane potential of carboxylic acids can be calculated from the reactions governing
the conversion of valerate to propionate and acetate, butyrate and propionate to acetate and
hydrogen, the conversion of hydrogen and carbon dioxide to methane, and the conversion
of acetate to methane and carbon dioxide.66 Calculated methane potentials are 720.8, 643.7,
537.0, and 373.0 1 CH4/kg of valerate, butyrate, propionate, and acetate, respectively.
Mass balances may be used to estimate the methane potential remaining in a landfill by
sampling the refuse, performing the appropriate chemical analyses, and calculating the
methane potential. Ideally, the chemical composition and methane potential of the refuse at
burial would also be known. If not, then typical cellulose and hemicellulose values must be
assumed. Comparison of the initial methane potential of the refuse with that at the time of
sampling will provide information on the fraction of the refuse that has been degraded.
As described here, the calculated methane potential assumes complete mineralization of
degradable organics. Some fraction of these constituents is surrounded by lignin and not
readily available for anaerobic decomposition. Thus, the actual methane potential of a refuse
sample will be less than that calculated. Bookter and Ham13 measured cellulose concentrations
as low as (8.2%) for 9-year-old shredded refuse that was well decomposed, as evidenced
by its humus-like appearance. As part of this same study, various landfills around the country
were sampled and cellulose concentrations as low as 6.6% were reported for nonshredded
refuse. Refuse sampled from a landfill in Wisconsin had cellulose concentrations of 10.0,
578 Critical Reviews in Environmental Control

6.1, and 3.1% for refuse buried in 1948, 1954, and 1957, respectively. Kinman et al.S7
found between 1.6 and 62.0% paper in 20-year-old refuse excavated from the Mallard North
landfill in Chicago, IL. The average paper concentration was 32.0%, and no cellulose
concentrations were reported. More data on the composition of refuse several years after
burial are needed to estimate the fraction of the cellulose and hemicellulose not available
for anaerobic decomposition.
Bookter and Ham13 were the first to measure cellulose and lignin concentrations in
landfills. They found the cellulose-to-lignin ratio to be a useful index of the degree of refuse
decomposition and remaining methane potential. Laquidara et al. 59 proposed estimation of
the methane potential of landfills based on measurement of the total solids, total volatile
solids, total organic carbon, and lignin concentrations in refuse.
In summary, chemical analyses and mass balances are useful for assessment of the
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

degree to which refuse has decomposed and the maximum remaining methane potential.
Using mass balances, Barlaz et al.9 documented the loss of 71% of the cellulose and 77%
of the hemicellulose present in fresh refuse. In this work, carbon recoveries of 70 to 100%
were obtained. In using the mass balance approach to project yields from landfills, it is
necessary to make assumptions regarding the maximum extent of cellulose plus hemicellulose
degradation. Mass balances are most useful for projection of methane production when used
in concert with drawdown tests to measure the current rate of methane production.59
Once a rate of methane production has been established by a pumping test, and the
methane potential has been estimated, models may be employed to determine the period
over which methane will be produced. This determination still requires assumptions on the
shape of the methane production rate curve.

C. MODELS FOR PROJECTION OF LANDFILL GAS PRODUCTION


Theoretical approaches to projection of the rate of gas generation in a landfill generally
involve development of models based on first-order kinetics as given in Equation 2. 48

where c = concentration of decomposable matter, t = time, k = first-order reaction rate


constant. This equation states that the rate of loss of decomposable matter is proportional
to the amount of decomposable matter remaining. It assumes that the factor limiting the rate
of methane generation at a landfill is the amount of degradable material remaining. In a
landfill, however, many factors other than substrate availability may contribute to reduce
gas generation. For example, moisture, nutrients, or inhibitory compounds may limit the
rate of methane production.
A zero-order kinetic model seems to most accurately describe the rate of gas production
at full-scale landfills during the most active period of gas generation. Many full-scale landfills
where gas is recovered produce approximately the same volume of gas on an annual basis
over the period of active gas generation.
Laquidara et al. 59 developed a model for prediction of methane production in landfills
that are no longer receiving refuse. Methane production is projected form a first-order model
using total volatile solids as the substrate concentration. Total volatile solids are categorized
as easy, moderate, or hard to degrade with different decay coefficients assigned to each
fraction. At landfills where the model does not accurately predict methane generation, the
model is to be calibrated to field data by adjustment of the decay coefficients. No procedure
for independent measurement of decay coefficients was suggested.
Table 7 presents the range of typical gas generation rates for different landfill situations
in the literature, along with information gathered from experience. Although the data include
Volume 19, Issue 6 (1990) 579

TABLE 7
Methane Production Rates Reported for Municipal Refuse

Conditions Methane production


rate
(I/kg dry refuse-year)

Lysimeters, average methane production rate observed 6.7 x 10~4—20.1


during periods of methane generation
Pilot or test landfills, rate during periods of active meth- 10—40
ane generation
Full-scale landfills, rate determined by pumping tests 0.54—26.8 (typically
6.7—13.4)
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Adapted from Ham, R. K. and Barlaz, M. A., Proc. ISWA Symp. Process,
Technol. Environ. Impact Sanit. Landfills, October 20—23, Cagliari, Sardinia,
Italy, 1987.

only lysimeters producing reasonable volumes of methane, there is still much variation in
the data. Gas production rates measured in full-scale landfills actually reflect the gas re-
covered, as any gas lost to the atmosphere was not measured. Gas production rates outside
of the range of 6.7 to 13.4 1 methane/dry kg per year tend to be from landfills subjected to
dry conditions, which reduce the rate of methane production, or from landfills containing
unusual wastes or subjected to large amounts of water. Segal79 reported the gas production
rates at 23 landfills and there was a factor of 7 difference between the landfills producing
the lowest and highest volumes of gas. There was significant variability in gas generation
rates even among landfills from the same geographical region.
From estimates of the total gas production potential, discussed in the previous section,
and the rate of gas production (Table 7), it is possible to project future production rates at
a landfill using either a zero- or first-order model. A typical approach is to assume a lag
phase or start-up period over which little methane is produced, followed by a period of
active methane generation, and, finally, a period of exponentially decreasing methane
production.
Each year's production of methane is modeled separately, and the total gas production
for each annual increment of refuse placed in the landfill is computed to give total annual
landfill gas projections. The useful life of methane gas generation by such modeling appears
to be 5 to 20 years, depending on the model used and the landfill under study. Note that if
a landfill continues to take refuse, portions of the site older than 5 to 20 years would be of
marginal interest for gas generation, but this would be offset by gas production from new
refuse so the total landfill gas production would remain constant if the yearly tonnage of
refuse remains constant.
Findikakis and Leckie34 developed a model of gas production and flow in sanitary
landfills. Required input data include the fraction of the refuse that is easily, moderately,
and poorly degradable, along with the corresponding half-lives, total gas generation potential,
and a detailed physical description of the landfill. The model predicts pressure gradients in
the landfill and gas losses through the surface of the landfill.
Recently, Findikakis et al.35 presented data to suggest that a first-order model did not
accurately predict methane production. They developed an empirical model, fitted to field
data from the controlled landfill project.2945 In the model the methane production rate
increased to a peak with a hyperbolic function and then decreased exponentially. This type
of behavior was also reported by Barlaz et al.6 Input parameters for the model include the
fraction of refuse that is readily, moderately, and slowly degradable, a generation time, a
lag time, a total generation potential, and a 90% conversion time.
580 Critical Reviews in Environmental Control

El-Fadel et al.25 added equations describing the biological reactions involved with refuse
methanogenesis to the gas flow model of Findikakis and Leckie.34 The model predicted that
the biokinetic constants for the acidogenic and methanogenic reactions would not significantly
effect the predicted methane generation rate. It also predicted that methane production would
be very sensitive to the rate of polymer hydrolysis and the concentration of soluble organic
carbon, particularly acetate. Barlaz et al. 6 also concluded that hydrolysis rate and acid
availability controlled the rate of methane generation based on laboratory-scale lysimeters.
The data reported by Sleat et al.80 also support the predictions generated by the model of
El-Fadel et al.

V. SUMMARY AND CONCLUSIONS


Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

There are ample data to document that refuse will decompose to methane in sanitary
landfills with measurable losses of cellulose, the principal biodegradable constituent of
refuse.13-46 Recent research on the microbiology of refuse decomposition has provided a
better understanding of the refuse ecosystem. Barlaz et al.6 showed that all of the trophic
groups required for refuse methanogenesis are present in fresh refuse and that initially acetate
utilization, but ultimately polymer hydrolysis, limits the rate of methane production. The
work of Jones et al., 53 in which it was shown that hydrolytic enzyme activities increase near
the water table, confirms laboratory results on the importance of moisture for refuse de-
composition. Using the increased understanding of the microbiology of refuse decomposition,
it was possible to evaluate studies on enhancement of methane production.
A major factor limiting the onset of methane production is a decrease in the pH of the
refuse ecosystem soon after refuse burial. While it was originally thought that this pH
decrease, or souring, was due to rapid cellulose hydrolysis,4 it was later suggested that the
initial pH decrease resulted from soluble sugar concentrations in refuse in excess of the
amount of oxygen and nitrate available for microbial oxidation of these sugars to carbon
dioxide.6 After depletion of the oxygen and nitrate, sugars are fermented to carboxylic acids.
These acids accumulate because the low populations of methanogenic bacteria in fresh refuse
cannot metabolize acids at the rate at which they are produced. Coupled with the low alkalinity
of fresh refuse,5 the acid accumulation causes a pH decrease that inhibits methane production.
There are several potential solutions to the souring problem. Two solutions were noted
as successful enhancement parameters earlier. In leachate recycle and neutralization, the pH
of the refuse ecosystem is maintained near neutrality, providing a more favorable environment
for methanogenic activity. The addition of old, anaerobically degraded refuse to fresh refuse
provides a source of methanogenic bacteria able to metabolize acids as they are produced,
thus circumventing an inhibitory pH decrease.
Food waste is the most likely source of soluble sugars in refuse. Thus, it may be possible
to reduce the soluble sugar concentration of fresh refuse by separation of food waste from
commercial sources. This separated waste could be composted prior to burial. Plant cells
are reported to undergo spontaneous lysis under anaerobic conditions and release soluble
sugars.41^3 It is conceivable that this source of soluble sugars would play a significant role
in the souring of landfills even if sugars in fresh refuse were to be eliminated. Clearly,
additional research is needed to explore potential solutions to the souring problem.
The most consistent enhancement techniques are leachate recycle and neutralization and
the use of old, anaerobically degraded refuse as a seed. Addition of anaerobic sewage sludge
and calcium carbonate also stimulated the onset of methane production.15 However, it is
unclear whether the sludge was stimulatory beyond the stimulation attributed to the addition
of calcium carbonate as a buffer.
Our knowledge of leachate characteristics as related to the state of refuse decomposition
is good, however, accurate projection of the methane production potential of landfills remains
Volume 19, Issue 6 (1990) 581

difficult. The number of landfill gas-recovery projects is increasing, and both gas recovery
wells and trenches have been used for gas collection. Nevertheless, assumptions are still
necessary to predict both the onset of methane production and total yields, and there is little
reliable field data that can be used to validate available models. Mass balances can provide
information on the state of refuse decomposition in a landfill that is not available from pump
tests. With time, more data on gas production from full-scale landfills should become
available. Analysis of these data will increase the ability to predict the behavior of the future
sanitary landfills with respect to gas production.
Historically, little attention has been given to fundamental research on refuse decom-
position. Such research is necessary to improve landfill design for enhanced methane pro-
duction and minimization of adverse environmental impacts resulting from poorly designed
or operated landfills. Only recently have fundamental studies led to an understanding of the
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

refuse ecosystem, and only recently Ham and Noble proposed an agenda for future funda-
mental research.49 They point out that traditionally field-scale demonstrations have been the
backbone of landfill experimentation. Clearly, such research will remain of high importance.
However, under field conditions without enhancement, refuse decomposition is a slow
process. The long periods of time, high cost, and inability to rigorously control experimental
conditions make large-scale tests inappropriate for testing of many effects. In many cases,
small-scale tests can be used effectively to answer specific questions. The time required for
smaller scale tests can be reduced by enhancement of refuse decomposition. Once proven
in small-scale tests, key concepts of refuse decomposition would then be verified in field-
scale test lysimeters.

ACKNOWLEDGMENTS
The authors wish to acknowledge the following organizations that supported research
on refuse decomposition over the past 10 years: The Wisconsin Alumni Research Foundation,
GSF Inc., Wisconsin Power and Light, and the American Public Power Association. In
addition, we thank Mark Milke for his early collaboration on refuse decomposition research
and the staff of the Madison Energy Recovery Plant for providing refuse in support of this
research.

REFERENCES
1. Augenstein, D. C., Cooney, C. L., Wise, D. L., and Wentworth, R. L., Fuel gas recovery from
controlled landfilling of municipal wastes, Resour. Recovery Conserv., 2, 103, 1976.
2. Bagnara, C. et al., Isolation and characterization of a cellulolytic mircoorganism, Cellulomonas fermentans
sp. nov., Int. J. Syst. Bacteriol., 35(4), 502, 1985.
3. Barlaz, M. A., Ham, R. K., and Milke, M. W., Parameters affecting refuse methanogenesis and the
solids composition of anaerobically degraded refuse, Proc. 9th Ann. Madison Waste Conf., Department of
Engineering Professional Development, University of Wisconsin, Madison, September 9 - 1 0 , 1986.
4. Barlaz, M. A., Milke, M. W., and Ham, R. K., Gas production parameters in sanitary landfill simulators,
Waste Manage. Res., 5, 27, 1987.
5. Barlaz, M. A., Microbial and Chemical Dynamics During Refuse Decomposition in a Simulated Sanitary
Landfill, Ph.D. thesis, Department of Civil and Environmental Engineering, University of Wisconsin,
Madison, 1988.
6. Barlaz, M. A., Schaefer, D. M., and Ham, R. K., Bacterial population development and chemical
characteristics of refuse decomposition in a simulated sanitary landfill, Appl. Environ. Microbiol., 55(1),
55, 1989.
7. Barlaz, M. A., Schaefer, D. M., and Ham, R. K., Effects of pre-chilling and sequential washing on the
enumeration of microorganisms from refuse, Appl. Environ. Microbiol., 55(1), 50, 1989.
582 Critical Reviews in Environmental Control

8. Barlaz, M. A., Ham, R. K., and Schaefer, D. M., Inhibition of methane formation from municipal
refuse in laboratory scale lysimeters, Appl. Biochem. Biotechnol., 20, 181, 1989.
9. Barlaz, M. A., Ham, R. K., and Schaefer, D. M., Mass balance analysis of decomposed refuse in
laboratory scale lysimeters, ASCE J. Environ. Eng., 115(6), 1088, 1989.
10. Barlaz, M. A., Ham, R. K., and Schaefer, D. M., Microbiological, chemical and methane producing
characteristics of refuse with and without leachate recycle, submitted.
11. Beeman, R. E. and Suflita, J. M., Microbial ecology of a shallow unconfmed ground water aquifer
polluted by municipal landfill leachate, Microb. Ecol., 14, 39, 1987.
12. Berenyi, E. B. and Gould, R. N., Methane recovery from municipal landfills in the U.S.A., Waste
Manage. Res., 4, 189, 1986.
12a. Bogner, J. E. et al., Modified biochemical methane potential (BMP) assays to assess biodegradation
potential of landfilled refuse, Proc. Fifth Int. Conf. Solid Wastes, Sludges, and Residual Materials: Char-
acterization, Technology, Management and Public Policy, Rome, Italy, April 26 to 29, 1989.
13. Bookter, T. J. and Ham, R. K., Stabilization of solid waste in landfills, ASCE J. Environ. Eng. Div.,
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

108(EE6), 1089, 1982.


14. Brock, T. D., Smith, D. W., and Madigan, M. T., Biology of Microorganisms. Prentice Hall, Englewood
Cliffs, NJ, 1984.
15. Buivid, M. G. et al., Fuel gas enhancement by controlled landfilling of municipal solid waste, Resour.
Recovery Conserv., 6, 3, 1981.
16. Burns, R. G. and Martin, J. P., Biodegradation of organic residues in soil, in Microflora and Faunal
Interactions in Natural and Agro-Ecosystems, Mitchell, M. J. and Nakas, J. P., Eds., Martinus Nijhoff,
The Hague, 1986.
17. Campbell, D. J. V. et al., Understanding refuse decomposition practices to improve landfill gas energy
potential, Proc. Energy Biomass 3rd EC Conf., Palz, W., Ed., Venice, Italy, Elsevier, London, March
2 5 - 2 9 , 1985.
18. Carra, J. S. Design criteria for sanitary landfills — the American concept, Proc. ISWA Symp. Process
Technol. Environ. Impact Sanit. Landfills, Cagliari, Sardinia, Italy, October 2 0 - 2 3 , 1987.
19. Chartrain, M. J. and Zeikus, J. G., Microbial ecophysiology of whey biomethanation: characterization
of bacterial trophic populations and prevalent species in continuous culture, Appl. Environ. Microbiol.,
51(1), 188, 1986.
20. Cook, H. A. et al., Microorganisms in household refuse and seepage water from sanitary landfills, Proc.
W. VA. Acad. Sci., 39, 107, 1967.
21. Courts, D. A. P. et al., Multi-stage chemostat investigation of interspecies interactions in a hexanoate-
catabolizing microbial association isolated from anoxic landfill, J. Appl. Bacteriol., 62, 251, 1987.
22. DeWalle, F. B., Chian, E. S. K., and Hammerburg, E., Gas production from solid waste in landfills,
ASCE J. Environ. Eng. Div., 104(EE3), 415, 1978.
23. Donnely, F. A. and Scarpino, P. V., Isolation, characterization and identification of microorganisms from
laboratory and full scale landfills, EPA Proj. Summary, EPA-600/S2-84-119, 1984.
24. Ehrig, H. J., Laboratory scale tests for anaerobic degradation of municipal solid waste, Proc. Int. Solid
Waste Assoc. Conf., Philadelphia, PA, 1984.
25. El-Fadel, M., Findikakis, A. N., and Leckie, J. O., A numerical model for methane production in
managed sanitary landfills, Waste Manage. Res., 7, 31, 1989.
26. Emberton, J. R., The biological and chemical characterization of landfills, Proc. Energy Landfill Gas,
Solihull, West Midlands, UK, October 3 0 - 3 1 , 1986.
27. Emcon Association, Sonoma County Solid Waste Stabilization Study, EPA 530/SW 65d.l, 1975.
28. Emcon Associates, Methane Generation and Recovery from Landfills, Ann Arbor Science Publishers, Ann
Arbor, MI, 1980.
29. Emcon Association, Controlled Landfill Project, Mountain View California, Fifth Annual Report, 1985.
Emcon Association, San Jose, CA, 1987.
30. Farquhar, G. J. and Rovers, F. A., Gas production during refuse decomposition, Water, Air, Soil Pollut.,
2, 483, 1973.
31. Fielding, E. T. and Archer, D. B., Microbiology of landfill. Identification of methanogenic bacteria and
their enumeration, in Effluent Treatment and Disposal, European Federation of Chemical Engineering, Ser.
No. 53, Pergamon Press, Elmsford, NY, 1986, 331.
32. Fielding, E. R., Archer, D. B., de Macario, E. C., and Macario, A. J. L., Isolation and characterization
of methanogenic bacteria from landfills, Appl. Environ. Microbiol., 54(3), 835, 1988.
33. Filip, A. and Kuster, E., Microbial activity and the turnover of organic matter in a municipal refuse
disposed of in a landfill, Eur. J. Appl. Microbiol. Biotechnol., 7, 371, 1979.
34. Findikakis, A. N. and Lecki, J. O., Numerical simulation of gas flow in sanitary landfills, ASCE J.
Environ. Eng. Div. 106(EE5), 927, 1979.
35. Findikakis, A. N., Papelis, P., Halvadakis, C. P., and Leckie, J. O., Modelling gas production in
managed sanitary landfills, Waste Manage. Res., 6, 115, 1988.
Volume 19, Issue 6 (1990) 583

36. Fungaroli, A. A. and Steiner, R. L., Investigation of sanitary landfill behavior, Municipal Environmental
Research Laboratory, Environmental Protection Agency, Cincinnati, OH, EPA-600/2-79-053a, 1979.
37. Gizven, H. J . et al., Anaerobic digestion of cellulosic fraction of domestic refuse by rumen microorganisms,
Biotechnol. Bioeng., 32, 749, 1988.
38. Gossett, J. M. et al., Heat treatment and anaerobic digestion of refuse, ASCEJ. Env. Eng. Div., 108(EE3),
437, 1982.
39. Governmental Advisory Associates (GAA), 1989, Methane Recovery from Landfills Yearbook 1988-1989,
New York, NY, 1989.
40. Grainger, J. M. et al., Estimation and control of microbial activity in landfill, in Microbiological Methods
for Environmental Biotechnology, Grainger, J. M. and Lynch, J. M., Eds., The Society for Applied
Bacteriology Technology, (9th series), Academic Press, Orlando, FL, 259, 1984.
41. Greenhill, W. L., Plant juices in relation to silage fermentation. I. The role of the juice, J. Br. Grassl.
Soc., 19, 30, 1964.
42. Greenhill, W. L., Plant juices in relation to silage fermentation, II. Factors effecting the release of juices,
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

J. Br. Grassl. Soc., 19, 231, 1964.


43. Greenhill, W. L., Plant juices in relation to silage fermentation. III. Effect of water activity of juice, J.
Br. Grassl. Soc., 19, 336, 1964.
44. Halvadakis, C. P. et al., Landfill methanogenesis: literature review and critique, Technical Report No.
271, Department of Civil Engineering, Stanford Univesity, 1983.
45. Halvadakis, C. P., Findikakis, A. N., Papelis, C., and Leckie, J. O., The Mountain View controlled
landfill project field experiment, Waste Manage. Res., 6, 103, 1988.
46. Ham, R. K., Hekimian, K. K., Katten, S. L., Lockman, W. J., Lofy, R. J., McFaddin, D. E., and
Daley, E. J., Recovery, processing, and utilization of gas from sanitary landfills, EPA - 600/2-79-001,
Municipal Environmental Research Laboratory, Cincinnati, OH, 1979.
47. Ham, R. K. and Bookter, T. J., Decomposition of solid waste in test lysimeters, ASCE J. Environ. Eng.
Div., 108(EE6), 1147, 1982.
48. Ham, R. K. and Barlaz, M. A., Measurement and prediction of landfill gas quality and quantity, in
Sanitary Landfilling: Process. Technology, and Environmental Impact, Christensen, T. H., Cossu, R., and
Stegmann, R., Eds., Academic Press, Orlando, FL, 1989, 155.
49. Ham, R. K. and Noble, J. J., The future of landfills: an agenda for fundamental research, presented at
the USEPA Conference on Municipal Solid Waste Technology, San Diego, CA, January 30 to February
1, 1989.
50. Hartz, K. E. et al., Temperature effects: methane generation from landfill samples, ASCE J. Environ.
Eng. Div., 108(EE4), 629, 1982.
51. Jenkins, R. L. and Pettus, J. A., The use of in vitro anaerobic landfill samples for estimating gas generation
rates, in Biotechnological Advances in Processing Municipal Wastesfor Fuels and Chemicals, Antonopoulos,
A. A., Ed., Argonne National Laboratory Report ANL/CNSV - TM - 167, 1985, 419.
52. Jones, K. L. and Grainger, K. M., The application of enzyme activity measurements to a study of factors
affecting protein, starch and cellulose fermentation in a domestic landfill, Eur. J. Appl. Microbiol. Bio-
technol., 18, 181, 1983.
53. Jones, K. L. et al., Methane generation and microbial activity in a domestic refuse landfill site, Eur. J.
Appl. Microbiol. Biotechnol., 18, 242, 1983.
54. Kasali, G. B. et al., Preliminary investigation of the influence of pH on the solid-state refuse methanogenic
fermentation, J. Appl. Bacteriol., 65, 231, 1988.
55. Kinman, R. N. et al., Gas characterization, microbiological analysis, and disposal of refuse in GRI landfill
simulators, EPA Project Summary EPA/600/S2-86/041, 1986.
56. Kinman, R. N. et al., Gas enhancement techniques in landfill simulators, Waste Manage. Res., 5, 13,
1987.
57. Kinman, R. N. et al., Analysis of 20-year-old refuse from the Mallard North Landfill in Chicago, Illinois,
Proc. Purdue Ind. Waste Conf., May 9 - 1 1 , West Lafayette, IN, 1989.
58. Klink, R. E. and Ham, R. K., Effects of moisture movement on methane production in solid waste
samples, Resour. Corner., 8, 29, 1982.
59. Laquidara, M. J., Leuschner, A. P., and Wise, D. L., Procedure for determining potential gas quantities
in an existing landfill, Water Sci. Tech., 18(12), 151, 1986.
60. Laube, V. M. and Martin, S. M., Conversion of cellulose to methane and carbon dioxide by triculture
of Acetivibrio cellulolyticus, Desulfovibrio sp., and Methanosarcina barkeri, Appl. Environ. Microbiol.,
42, 413, 1981.
61. Leckis, J. O., Pacey, J. G., and Halvadakis, C., Landfill management with moisture control, ASCE J.
Environ. Eng. Div., 105(EE2), 337, 1979.
62. Lewis, J., What's in the solid waste stream, EPA J., 15(2), 15, 1989.
63. Lu, A. and Kunz, C. O., Gas-flow model to determine methane production at sanitary landfills, Environ.
Sci. Tech., 15(5), 436, 1981.
584 Critical Reviews in Environmental Control

64. Mah, R. A., Interactions of methanogens and non-methanogens in microbial ecosystems, Proc. 3rd Int.
Symp. Anaerobic Dig., August 1 4 - 1 9 , Boston, MA, 1983.
65. Mata-Alvarez, J. and Martinex-Viturtia, A., Laboratory simulation of municipal solid waste fermentation
with leachate recycle, J. Chem. Tech. Biotechnol., 36, 547, 1986.
66. McInerney, M. J. and Bryant, M. P., Basic principles of bioconversions in anaerobic digestion and
methanogenesis, in Biomass Conversion Processes for Energy and Fuels, Sofar, S. S. and Zaborsky, O.,
Eds., Plenum Press, New York, 1981.
67. Merz, R. C. and Stone, R., Gas production in a sanitary landfill, Public Works, 95(2), 84, 1964.
68. Merz, R. C. and Stone, R., Landfill settlement rates, Public Works, 93(9), 103, 1962.
69. Parkin, G. F. and Owen, W. F., Fundamentals of anaerobic digestion of wastewater sludges, ASCE J.
Environ. Eng. Div., 112(5), 867, 1986.
70. Pfeffer, J. T., Temperature effects on anaerobic fermentation of domestic refuse, Biotechnol. Bioeng., 16,
771, 1974.
71. Pfeffer, J. T. and Khan, K. A., Microbial production of methane from municipal refuse, Biotechnol.
Downloaded by [UZH Hauptbibliothek / Zentralbibliothek Zürich] at 18:06 05 July 2014

Bioeng., 18, 1179, 1976.


72. Pohland, F. G., Sanitary Landfill Stabilization with Leachate Recycle and Residual Treatment, Georgia
Institute of Technology, EPA Grant No. R-801397, 1975.
73. Pohland, F. G., Leachate recycle as a landfill management option, ASCEJ. Environ. Eng. Div., 107(EE6),
1057, 1980.
74. Pohland, F. G., Dertien, J. T., and Ghosh, S. B., Leachate and gas quality changes during landfill
stabilization of municipal refuse, in 3rd Int. Sympj Anaerobic Dig., August 1 4 - 1 9 , Boston, MA, 1983,
185.
75. Rees, J. F., The fate of carbon compounds in the landfill disposal of organic matter, Chem. Tech.
Biotechnol., 30, 161, 1980.
76. Rees, J . F . et al., Microbial transformation processes in sanitary landfills, in Gas- und Wasserhaushalt
von Mulldenponien, Technische Universitat Braunschweig, 1982, C l .
77. Rees, J. F. and Grainger, J. M., Rubbish dump or fermenter? Prospects for the control of refuse
fermentation to methane in landfills, Process Biochem., November/December, 41, 1982.
78. Riley, R. D. et al., Chemical and microbiological studies on leachate from a waste tip, J. Appl. Bacterial.,
42, 285, 1977.
79. Segal, J. P., Testing large landfill sites before construction of gas recovery facilities, Waste Manage. Res.,
5, 123, 1987.
80. Sleat, R. et al., Activities and distribution of key microbial groups in landfill, in Sanitary Landfilling:
Process, Technology, and Environmental Impact, Christensen, T. H., Cossu, R., and Stegmann, R., Eds.,
Academic Press, Orlando, FL, 1989, 51.
81. Stegmann, R., New aspects on enhancing biological processes in sanitary landfills, Waste Manage. Res.,
1, 201, 1983.
82. Tchobanoglous, G., Theisen, H., and Eliassen, R., Solid Wastes, McGraw-Hill, New York, 1977, 334.
83. Tibbies, B. J. and Baecker, A. A. W., Effects and fate of phenol in simulated landfill sites, Microb.
Ecol., 17, 201, 1989.
84. Watson-Craik, I. A. and Senior, E., Treatment of phenolic wastewaters by co-disposal with refuse, Water
Res., 13(10), 1293, 1989.
85. Wolfe, R. S., Methanogenesis, in Microbial Biochemistry, International Review of Biochemistry, Vol. 2 1 ,
Quayle, J. R., Ed., University Park Press, Baltimore, MD, 1979.
86. Wujcik, W. J. and Jewell, W. J., Dry anaerobic fermentation, Biotechnol. Bioeng. Symp., (10), 43,
1980.
87. Young, L. Y. and Frazer, A. C., The Fate of lignin and lignin derived compounds in anaerobic envi-
ronments, Geomicrobiology J., 5, 261, 1987.
88. Zehnder, A. J. B., Ecology of methane formation, in Water Pollution Microbiology, Vol. 2, Mitchell,
R., Ed., John Wiley & Sons, New York, 1978, 349.
89. Zehnder, A. J. B. et al., Microbiology of methane bacteria, in Anaerobic Digestion, Hughes, D. E., Ed.,
Elsevier, Amsterdam, 1982, 45.
90. Zeikus, J. G., Microbial populations in digesters, in Anaerobic Digestion, Stafford, D. A., Ed., Applied
Science Publishers, Englewood, NJ, 1980, 61.
91. Zwart, K. B. et al., Anaerobic digestion of a cellulose fraction of domestic refuse by a two-phase rumen
derived process, Biotechnol. Bioeng., 32, 719, 1988.

You might also like