You are on page 1of 11

Computers and Geotechnics 118 (2020) 103341

Contents lists available at ScienceDirect

Computers and Geotechnics


journal homepage: www.elsevier.com/locate/compgeo

Research Paper

A degradation–consolidation model for the stabilization behavior of T


landfilled municipal solid waste
Y.M. Chena,b, W.J. Xua,b, , D.S. Linga,b, L.T. Zhana,b, W. Gaoa

a
MOE Key Laboratory of Soft Soils and Geoenvironmental Engineering, Zhejiang University, China
b
Center for Hypergravity Experimental and Interdisciplinary Research, Zhejiang University, China

ARTICLE INFO ABSTRACT

Keywords: The stabilization behavior of landfilled municipal solid waste (MSW) plays an important role in the design,
Landfilled municipal solid waste operation, and after care of landfills. The stabilization is dependent on the biological degradation and con-
Stabilization solidation processes of landfilled MSW. A degradation–consolidation model for MSW was developed, which took
Biochemical-hydro-mechanical coupling biological degradation, skeleton deformation, two-phase flow, and solutes transport into consideration. In this
Numerical modelling
model, the mass transformations of solid to liquid and then to gas were involved. The compression, hydraulic
and solute transport material properties were related to the solid mass loss and the pore volume change.
Indicators for the stabilization behavior of landfilled MSW were proposed. These indicators were calculated
using the degradation–consolidation model to compare the stabilization behavior of landfilled high food waste
content (HFWC) and low food waste content (LFWC) MSW. The results showed that the stabilization process of
HFWC MSW landfills could be divided into rapid degradation, slow degradation, and post-stabilization stages
based on the development of stabilization indicators. The normalized ratio of cellulose to lignin could be im-
plemented to assess the stabilization stage of landfill more easily in practice.

1. Introduction should be based on the assessment of the leachate amount and the so-
lute concentration. And the operation of a LFG collection system is
In modern geotechnical engineering, some special materials, such as dependent on the LFG generation rate [3]. For the land reuse of a
municipal solid waste (MSW), coal, and natural gas hydrate, have a landfill, the settlement rate must be stable enough.
solid skeleton that can be converted into liquid or gas phases by bio- Landfilled MSW normally contains various physical components,
logical and chemical processes, resulting in mass loss and changes of such as food waste, paper, wood, textiles, plastic, metal, and glass. The
the grain size. Such soils are defined as degradable soils [1]. Taking the main degradable chemical compositions of the solid skeleton include
MSW as an example, the degradable substances are decomposed with cellulose, sugars, proteins, and lipids [1,4]. The anaerobic degradation
biochemical reactions, when liquid, landfill gases (LFG), and solutes are can be divided into four stages: hydrolysis, acidogenesis, acetogenesis,
generated. The dissipation of the generated liquid and LFG causes the and methanogenesis. Multiple-stage anaerobic degradation models for
decrease of pore pressure, and the solid skeleton is softened as a result MSW that can refine the anaerobic degradation of MSW more accu-
of the mass loss of the solid phase [2]. The decrease of the pore pressure rately were proposed [5,6]. However, these models are difficult to
and the softening of the solid skeleton induce the compression of the analyze the degradation and the consolidation of coupled process due
volume, which is defined as the consolidation of the MSW. The bio- to their complexity and various of parameters. A simplified two-stage
chemical reaction rate, as well as the degradation rate of the solid model for MSW is commonly used, which mainly considers the impacts
skeleton and the generation rates of the liquid and LFG, are attenuated of factors, including water, ammonia, hydrogen sulfide, and volatile
over time, and the consolidation is carried out with the dissipation of fatty acids on the degradation rate of the chemical composition [1,7].
the pore water and the gas pressure. Hence, the degradable soils and The water content has significant impact on the degradation process
their subsequent impacts on their surroundings would gradually stabi- [8], and large amount of intra-particle water will be released by the
lize, which is called the stabilization of landfilled MSW. The stabiliza- MSW with high food waste content. However, the generation of intra-
tion behavior of landfilled MSW plays an important role in the design, particle water is neglected by most anaerobic degradation models that
operation and after care of landfills. The design of the barrier system are based on MSW with a low food waste content.


Corresponding author.
E-mail address: wenjiexu@zju.edu.cn (W.J. Xu).

https://doi.org/10.1016/j.compgeo.2019.103341
Received 26 May 2019; Received in revised form 13 October 2019; Accepted 8 November 2019
0266-352X/ © 2019 Elsevier Ltd. All rights reserved.
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

Nomenclature Sw water saturation [–]


Ds stiffness matrix [–]
Si mass concentration of the chemical composition in the qw advection velocity of the water flow [m s−1]
degradation model [kg m−3] qg advection velocity of the gas flow [m s−1]
Cj,i mass ratio coefficient of the degradation model [–] ki intrinsic permeability [m2]
αi dimensionless transfer coefficient for Si [–] kr,w relative permeability of the water phase [–]
Rj reaction rate for the jth reaction [–] kr,g relative permeability of the gas phase [–]
kHj hydrolysis constant factor [d−1] μw viscosity of the water phase [Pa s]
fw water content-dependent coefficient [–] μg viscosity of the gas phase [Pa s]
fIh VFA inhibition factor for the hydrolysis stage [–] Dl hydraulic diffusion coefficient [m2 s−1]
θw water content of MSW [–] D0 free diffusion coefficient [m2 s−1]
θmin minimum values of θw for the hydrolysis stage [–] ε strain tensor [–]
θmax maximum values of θw for the hydrolysis stage [–] Sσ′ flexibility matrix of the effective stress induced strain [–]
Kh hydrolysis inhibition constant [kg m−3] SMsd coefficient of the degradable solid mass loss induced strain
nh hydrolysis inhibition index for fIh [–] [–]
Y yield coefficient [–] t time [s]
kmmax maximum growth rate constant of MB [d−1] σ′0 pre-consolidation pressure [Pa]
Ks half saturation constant of MB [kg m−3] C′C0 primary compression indicator for the fresh MSW [–]
fIm function for VFA inhibition of MB growth [kg m−3] C′C∞ primary compression indicator for the fully decomposed
λw intra-particle water factor [–] MSW [–]
kd MB decay rate constant [d−1] εs∞ ultimate secondary compression strain of the fresh MSW
fs solid mass change of MSW [kg m−3] [–]
fw liquid mass change of MSW [kg m−3] n0 referent porosity n0 [–]
fg gas mass change of MSW [kg m−3] k0 referent intrinsic permeability with the referent porosity
fS7,8 dissolved VFA and MB mass concentration change [kg n0 [m−2]
m−3] θe effective water content [–]
pg gas pressure [Pa] θwr residual water content [–]
u displacement [m] θws maximum water content [–]
pc capillary pressure [Pa] αv, nv parameters of the van Genuchten model [–]
σ′ effective stress tensor [Pa] τ tortuosity [–]
α biot coefficient [–] z depth of the landfilled MSW layer [m]
χ bishop coefficient [–] Uhyd hydrolysis degree [–]
ρ density of the porous medium [kg m−3] Uwater water generation degree [–]
ρs density of the solid [kg m−3] UCH4 methane generation degree [–]
ρw density of the water [kg m−3] RC/L ratio between cellulose and lignin
ρg density of the gas [kg m−3] β normalized RC/L [–]

In addition to traditional unsaturated consolidation, i.e. soil ske- proposed coupled models that were very useful in investigating the
leton deformation, and pore water and gas transport, originating from impact of hydraulic conditions on the settlement development. How-
conventional soils, the stabilization of landfilled MSW includes bio- ever, the degradation process was characterized by the empirical
chemical reactions and solute migration. The processes of consolidation model, which was not coupled with the consolidation process. Feng
and degradation interact with each other. The development of a com- et al. [13] has developed an improved HBM-coupled model, in which a
prehensive model for landfilled MSW should consider three aspects: (1) time-dependent empirical model was used to calculate the deformation
the multi-physics process, including but not limited to solid skeleton during degradation. The volume change caused by effective stress and
deformation, pore water flow, solute transport, pore gas flow, and degree of saturation, and the effect of void ratio change on hydraulic
biochemical degradation; (2) the generation, breakdown, and conver- properties were all considered. In the research work by Hubert et al.
sion of the solid skeleton, liquid, gas, and solute; and (3) the solid mass [14], a full-coupled thermal-hydro-biochemical-mechanical model was
loss induced variations of the engineering properties of MSW, including developed for simulating the long-term behavior of landfilled MSW in a
the strain–stress behavior, intrinsic permeability, and soil–water char- bioreactor landfill. This model adopted the biodegradation model de-
acteristic curve of MSW [9,10]. scribed by McDougall [1], which was more applicable to LFWC MSW.
A few multi-processes-coupled models have been developed to de- In addition, the gas pressure was assumed to be constant in this model,
scribe the stabilization behavior of landfilled MSW. Kindlein et al. [11] hence the accumulated gas pressure in HFWC MSW landfills could not
proposed a thermal-hydro-biological (THB)-coupled model, which be investigated.
considered the macroscopic heat transport, the liquid–gas flow, the In this study, a degradation–consolidation model was established to
multi-composition transport, and the local degradation. However, the describe the stabilization behavior of landfilled MSW. In this model, (1)
variation of hydraulic properties caused by the change of porosity as a solid skeleton deformation, pore water flow, solute transport, pore gas
result of the deformation of a solid skeleton, with a significant effect on flow, and biochemical degradation were linked via selected primary
the liquid-gas flow process, was not considered. McDougall et al. [1] variables; (2) the biodegradation behavior of landfilled MSW was de-
developed an innovative hydro-bio-mechanical (HBM)-coupled model scribed by a two-stage anaerobic model accounting for all degradable
that calculated the secondary settlement of landfills as a time-depen- compositions, and the release of immobile intra-particle water to mo-
dent function. This model could capture the accelerated settlement rate bile inter-particle water; and (3) the changes of the intrinsic perme-
owing to degradation, which was activated by the leachate recircula- ability, water retention curve (WRC), and the diffusion coefficient were
tion. However, the generation of LFG and the influence of pore pressure related to porosity, which varied with both the degradation and con-
on the deformation were not involved in the model. Yu et al. [12] also solidation processes. The numerical solution of this model was

2
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

Table 1
Chemical compositions in mass per unit volume of landfilled HFWC and LFWC MSW (kg/m3).
Category Void Bulk unit weight Rapid degradable Slowly degradable Sugars Proteins Lipids Inert Intra-particle Inter-particle Inorganic
ratio (-) (kN/m3) cellulose cellulose water water fractions

HFWC MSW 2.64 10.73 31 20 66 5.5 5.5 130 442 283 250
LFWC MSW 1.50 11.35 137 176 28 2.3 2.3 167 111 37 607

implemented on an open source finite element method (FEM) code C6H10O5⋅λwH2O + 5H2O → CH3COOH + 8H2 + 4CO2 + λwH2O (1)
OpenGeoSys (OGS) [15–18].
Hydrolysis of slowly degradable cellulose:
2. Mass transformation during degradation C6H10O5 + 5H2O → CH3COOH + 8H2 + 4CO2 (2)

The physical components of fresh MSW in different countries are dif- Hydrolysis of sugars:
ferent. The food waste content is the most significant difference between C6H10O5 + 5H2O → CH3COOH + 8H2 + 4CO2 (3)
developed countries, such as the U.S.A. and European countries, and the
developing countries such as China, India, and Brazil. Thus, the landfilled Hydrolysis of proteins:
MSW can be simply classified into two categories, namely high-food- C46H77O17N12S + 59.26H2O →
waste-content (HFWC) and low-food-waste-content (LFWC) MSW. The 7.88CH3COOH + 30CO2 + 63H2 + 12NH3 + H2S (4)
food waste content in HFWC MSW is higher than 40% [19] which contains
much more rapidly degradable compositions. The water content, including Hydrolysis of lipids:
intra-particle water and inter-particle water, of HFWC MSW is much
C55H104O6 + 78H2O → 13CH3COOH + 29CO2 + 104H2 (5)
higher than that of LFWC MSW. The mass fraction of intra-particle water is
higher than the inter-particle water based on laboratory measurement of Methanogenesis of VFA:
fresh HFWC MSW [20]. The chemical composition concentrations of the
fresh HFWC and LFWC are calculated from the typical physical compo- CH3COOH + 7.43H2 + 0.93CO2 + 0.029NH3 →
nents of the MSWs from China and USA, with units of kg/m3, and listed in 2.79CH4 + 3.8H2O + 0.029C5H7NO2 (6)
Table 1 [19]. The chemical compositions in Table 1 indicate that the most Death of MB:
significant difference between HFWC and LFWC MSWs is the content of
sugars, which is a rapidly degradable material. Due to the high mass C5H7NO2 + 2H2O → 0.83C6H10O5 + NH3 (7)
content of food waste, the rapidly degradable materials in the HFWC MSW The entire degradation process has two stages, which include
take up more than 70% of the total degradable materials, and this per- seven chemical reactions and 13 substances. The biochemical reac-
centage is about 50% in the LFWC MSW. The slowly degradable materials tions are related to the rapidly degradable cellulose with intra-par-
contents in the HFWC and LFWC MSWs are about 20% and 50%, re- ticle water C6H10O5.λwH2O, the slowly degradable cellulose
spectively. C6H10O5, the sugars C6H10O5, the proteins C46H77O17N12S, the lipids
The biochemical degradation drives the multi-field process in C55H104O6, the pore water H2O, the VFA CH3COOH, the MB
landfills. Substance transformations during the biodegradation of MSW C5H7NO2, the dissolved NH3, CO2, H2S, CH4, and H2, which are
are illustrated in Fig. 1. As shown in the figure, like soils, MSW is numbered from 1 to 13, respectively. It should be noted that λw is the
comprised of solid, pore water, and pore gas. The special characteristic intra-particle water factor that described the molar amount ratio
of MSW is that the solid skeleton is degradable, which includes the between the intra-particle water and the rapidly degradable cellu-
intra-particle water and the degradable compositions. During the bio- lose. Thus, according to Eq. (1) the water release amount during the
degradation, the degradable compositions are decomposed and trans- biodegradation can be calculated. The intra-particle water factor λw
formed into liquid and gas phases, including the intra-particle water, value of the HFWC MSW is significantly higher than the value of
the dissolved solutes and pore gas. In this study, volatile fatty acids LFWC MSW as shown in Table 1 [22].
(VFA) and methanogenic biomass (MB) are considered as dissolved According to the principle of mass conservation, the reactant and
solute. Methane (CH4) and carbon dioxide (CO2) are the main compo- the product in the jth biochemical reaction are the same, and this can be
nents of the generated LFG [21]. The biochemical processes can be expressed in the following form [23]:
characterized according to mass conservation theory.
According to the anaerobic two-stage degradation model, the hy- Cj, i·Si=0 (j = 1, 2, , 7; i = 1, 2, , 13) (8)
drolysis and methanogenesis process are involved. Furthermore, in this
where Cj,i is the mass ratio coefficient, and j and i stand for the number
study the intra-particle water release is considered.
of biochemical reactions and the number of compound, respectively. Si
Hydrolysis of rapidly degradable cellulose:
is the mass concentration of each substance.

18( w 5) 60 176 16
1 0 0 0 0 0 0 0 0
162 + 18 w 162 + 18 w 162 + 18 w 162 + 18 w
0 1 0 0 0 0.55 0.37 0 0 1.08 0 0 0.098
0 0 1 0 0 0.55 0.37 0 0 1.08 0 0 0.096
C=
0 0 0 1 0 0.593 0.394 0 0.12 1.2 1 0 0.126
0 0 0 0 1 1.56 0.91 0 0 0.145 0 0 0.208
0 0 0 0 0 1.14 1 0.0543 0.00812 0.679 0 0.753 0.245
0 0 1.1869 0 0 0.32 0 1 0.15 0 0 0 0

3
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

Fig. 1. Transformation of three-phase substances during biodegradation.

Generally, the changing rate of each compound can be expressed as The mass change of the gas phase can be calculated from the gas
follows: generation rate of CO2, H2S, CH4 and H2, and it has the following form:
7 13
dSi dSi
= i (Cj, i Rj ) (i = 1, 2, , 13) fg =
dt j=1 (9) i = 10
dt (17)

where αi is the dimensionless transfer coefficient for Si, defined as 1/θ The VFA and the MB make up the dissolved solute in the anaerobic
(θ is the volumetric water content) for solutes and equal to 1 for other degradation process:
compound, and Rj is the reaction rate for Eqs. (1)–(7), which has the 7
dSi
following form [4,8,24,25]: f Si = = i (Cj, i Rj ) (i = 7, 8)
dt j=1 (18)
Rj = kHj fW fIh Sj (j = 1, 2, ...,5) (10)

1 k mmax S8 3. Governing equations


R6 = w f S7
Y KS + c Im (11)
Based on continuum mechanics, a degradation–consolidation model
R7 = w k d S8 (12)
of landfilled MSW has been established, considering the biochemical
where kHj is the hydrolysis constant factor, fW is the water content degradation, skeleton deformation, pore water flow, pore gas flow,
dependent coefficient, and fIh is the VFA inhibition factor. fW and fIh had solute transport and their interactions with time. As shown in Fig. 2,
the following formulations [26]: solid mass is lost, gas is generated, and water and solute are released
due to the biochemical degradation, which can be calculated using Eqs.
0 ( < min )
(15)–(18). The solid skeleton of the MSW is deformed due to the solid
fW ( ) = w min
( min max ) mass loss and the change of the effective stress, gas flowed with the
max min
generation of gas, and water and solute are transported with the release
1 ( > max ) (13)
of water and solute. In the model, the capillary pressure (pc), the gas
1 pressure (pg), the displacements (u), and the concentration (Si) of each
fIh (S7) =
1 + (S7 Kh )nh (14) compound are selected as primary variables that controlled the coupled
processes of the degradable soil. Apart from the assumptions applied in
where θw is the water content of the MSW, and θmin and θmax are the
continuum mechanics, the following assumptions are made: (1) the
minimum and maximum values of θ, respectively. Kh is the hydrolysis
solid particles are incompressible; (2) the pore water is continuous in
inhibition constant, nh is the hydrolysis inhibition exponent, and S7 is
the void space and incompressible; (3) the pore gas is simplified as a
the mass concentration of VFA. R6 is the reaction rate of methano-
mixture of gases that was continuous in the void space; (4) the pore
genesis for VFA, R7 is the death rate of MB, S8 is the mass concentration
water and pore gas flows obeys Darcy’s law; (5) the influences of the
of MB, Y is the yield coefficient, km max was the maximum growth rate
solutes on the physical properties (e.g. density, viscosity, permeability)
constant of MB, Ks is the half saturation constant of MB, fIm is the
function for the VFA inhibition of MB growth, and kd is the MB decay
rate constant.
Since the changing rate of each compound can be calculated by Eq.
(9), the mass changing rates of the solid, liquid, gas and dissolved solute
can also be calculated. The solid mass loss, which consists of rapidly
degradable cellulose, slowly degradable cellulose, sugar, proteins and
lipids, can be expressed as:
5
dSi
fs =
i=1
dt (15)
The water change caused by the intra-particle water release and the
consumption of hydrolysis and methanogenesis has the following form:
dS6
fw =
dt (16) Fig. 2. Processes in the degradation–consolidation model of landfilled MSW.

4
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

of pore water are neglected; (6) the mechanical dispersion and the solid where Si is the solute concentration in the pore water phase. Two kinds
adsorption of the solutes are neglected; (7) the isothermal condition is of compounds in the biochemical degradation model are considered as
considered by neglecting the heat generation, consumption, and the dissolved solutes, namely the VFA and the MB. vw is the flow rate of
transfer during the coupled processes; (8) the phase transition due to pore water and Dl is the hydraulic diffusion coefficient, which also
mechanical action is neglected. varies with the degradation-induced change of porosity.
The governing equations (Eqs. (21), (23), (24), and (25)) in this
3.1. Balance equations study take the skeleton deformation, two-phase flow, and solute
transport into consideration. The displacement, capillary pressure, gas
3.1.1. Momentum balance equation pressure, and concentration of dissolved solute are the primary vari-
The momentum balance in terms of the effective stress, which ables of this model.
considered the water saturation, capillary pressure, and gas pressure, is
expressed as [18,27,28]: 3.2. Constitutive equations
( (p g p c ) I) + g = 0 (19)
Constitutive equations are needed to describe the behaviors of the
where is the effective stress tensor, and = Ds: u with Ds is con- mechanical, hydraulic and mass transport processes. The total strain of
sidered the stiffness matrix, which is dependent on the stress state and MSW is related to the effective stress state and the degradation process.
the degradable solid mass. u is the displacement vector, α is the Biot It can be written in the following form:
coefficient, I is the unit vector, g is the gravity vector, χ is the Bishop 5
coefficient, and ρ is the density of the porous medium, which is given = f ( , Msd ), Msd = Si
by: 1 (26)
= (1 n) s + nSw w + n (1 Sw ) g
(20) where Msd is the degradable solid mass. The derivative form of Eq. (26)
where ρs, ρw, and ρg are the densities of the solid, liquid, and gas phases, is as follows:
respectively, n is the porosity of the MSW, and Sw is the liquid satura- =S + S Msd Msd I (27)
tion of the MSW. Eq. (19) is rewritten in time derivative form as fol-
lows: where S is the flexibility matrix of the effective stress-induced strain
and SMsd is the coefficient of the degradable solid mass loss induced
u pg pc strain. According to the definition of the effective stress (Eq. (19)), the
Ds I + I + g=0
t t t t (21) change of the pore gas pressure and the capillary pressure can induce
where, in this study, α = 1 and χ = 0, which means that the capillary the change of the effective stress and then induce deformation. The
pressure’s contribution to the effective stress is neglected. stiffness matrix in Eq. (21) is the inverse of the flexibility matrix S
Based on long-term laboratory experiments on fresh HFWC MSW
3.1.2. Mass conservation equations [9,29], the total compression strain ε is assumed to be insensitive to the
The two-phase flow in a deformable porous medium is considered in history of stress and degradation and that it was only dependent on the
this model. The mass balance equation for both the water and gas final stress level and the degree of biodegradation at the time t (i.e.
phases is expressed as follows [16]: elapsed time). Considering a simplified one-dimensional form, the
stress–strain behavior of MSW can be written in the following form:
n S u
+n S + q =f
t t (22) t
Msd
= CC 0 lg + s ( 0) + (CC CC 0 ) lg 1 ini
where π is the indicator for the water or gas phase, q is the velocity, π 0 0 Msd (28)
and fπ represents the generation or consumption of water and gas. where the application of σ′0 represents the pre-consolidation pressure of
Since the primary variables of the coupled model are pc, pg and u, MSW, and CC0′ and CC∞′ represent the modified primary compression
the mass balance equations of the water and gas phases can be written indicator for fresh MSW and fully decomposed MSW, respectively. εS∞
in the following forms: represents the ultimate secondary compression strain of fresh MSW
Sw p c u n under an applied load of σ′0, and Msd t
and Msdini
are the masses of the
n + Sw + Sw
pc t t t degradable solids of MSW at time t and the initial time, respectively. It
+ ( w k ikr , w (
µw
(p g pc ) + w g)
)=f w
(23)
should be noted that if the degradation process was first-order kinetic,
the compression model was the same as the time-dependent model of
Sw p c g pg u n
Chen et al. [9]. Following the Eq. (26), the flexibility matrix and the
n pc t
+ n (1 Sw ) pg t
+ (1 Sw ) t
+ (1 Sw ) t coefficient of degradation-induced strain can be calculated from Eq.
(27), and these formulas have the following forms:
+ ( g k ikr , g (
µg
pg + g g)
)=f g
(24) t
Msd
CC 0 + (CC CC 0 ) 1
where the source terms of water and gas phase are obtained from Eqs. ini
Msd
S =
(16) and (17) according to the biochemical reaction, ki is the intrinsic ln 10 (29)
permeability tensor, kr,w and kr,g are the relative permeabilities of the
water and gas phases, respectively, and μw and μg are the viscosities of SMsd =
1
(
s 0) + (CC CC 0) lg
the water and gas phases, respectively. The intrinsic permeability and
ini
Msd 0 (30)
the water retention curve change with degradation process due to the
In this study, the intrinsic permeabilities for both the water and gas
solid mass loss. This effect is considered in this study and it will be
phases are assumed to be identical. Laboratory water permeability and
introduced in Section 3.2.
air permeability tests were carried out to study the intrinsic and relative
The dissolved solute transport is involved in this model, which can
permeabilities of Chinese MSW at different depths. The measured in-
be also described by the mass conservation equation in the following
trinsic permeability showed a decrease with increased depth and age. In
form:
this study, a modified intrinsic permeability model is introduced, and
Si
+ v w Si Dl 2S = f Si (i = 7, 8) the variation of the intrinsic permeability is related to the porosity with
t
i
(25) the following formula:

5
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

n which can be expressed as follows:


ki = k 0 exp A 1
n0 (31) D li = n D0i (i = 7, 8) (42)
where k0 is the referent intrinsic permeability with the referent porosity i
where D1 is the diffusion coefficient, D0 is the free diffusion coefficient, i
n0. In this study, k0 = 10−10 m2 and n0 = 0.75. A is a fitted model and τ is the tortuosity. The diffusion coefficient also changes with
parameter for the measurement of in-situ samples from different depths porosity, which is decreased with the depth and age of landfills.
and ages (Fig. 3), which is presented by [10], and the value is taken as
33 in this study.
4. Numerical Solutions
The results of laboratory experiments showed the relative perme-
abilities of the water and gas phases could be well represented by the
The governing equations for the degradation–consolidation model
van-Genuchten–Mualem model as follows [30]:
include Eqs. (21) and (23)–(25). The influence of the biochemical re-
kr,w = e
0.5
[1 (1 e ) ]
1 mv mv 2
(32) actions on the coupled multi-processes in landfilled MSW is reflected
by: (1) the source terms of the solid, water, gas and solutes (i.e. fs, fw, fg
kr , g = (1 e)
0.5
(1 e )
1 mv 2mv
(33) and fS7,8), (2) the change of the porosity and compressibility due to the
biochemical degradation induced mass loss, and (3) the change of the
where krw and krg are the relative permeabilities of the water phase and
engineering properties of landfilled MSW due to the change of porosity,
the gas phase, respectively.
i.e. the intrinsic permeability, WRC, and diffusion coefficient.
In the MSW, the porosity varies with compression and degradation.
The finite element model is solved by applying the Galerkin pro-
During compression, the porosity decreases and the degradation causes
cedure for spatial integration and the Generalized Trapezoidal Method
the loss of the solid mass, which induces an increase of porosity. The
for the time integration of the weak forms of the balance equations in
change of porosity can be calculated from the following equation:
Section 3. After spatial discretization within the isoperimetric for-
t
Msd mulation, the following non-symmetric, non-linear, and coupled system
nt = n0 + nt = n0 + v
s (34) of equations are obtained:

where n is the initial porosity, Δn is the increment of the porosity, Δεv


0 t
Cgg Cgc Cgu pg K gg K gc 0 pg fg
is the increment of the volumetric strain, which is negative for com- 0 Ccc Ccu pc + K cg K cc 0 pc = fw
t
pression in this study, and ρs is the density of the MSW grain, which is 0 0 0 u Kug Kuc Kuu u fu (43)
assumed constant in this study.
The hydraulic behavior is described by the conventional un-
Cs S7,8 + Ks S7,8 = f s7,8
saturated flow in porous mediums. The capillary pressure is defined by t (44)
the sorption equilibrium as follows:
where the solid displacements (u) and the gas and capillary pressures
pc = p g pw (35) (pg, pc) are the primary variables of the multi-phase flow and de-
The van Genuchten model is used to define the relationship between formation-coupled equations. The equation is solved with the full-
the water saturation and the capillary pressure as follows [30]: coupled monolithic schema. The solute transport is partially coupled
with two-phase flow and the mechanics processes. The source term fS7,8
1
1 nv 1 nv in the Right-Hand-Side (RHS) of Eq. (44) is calculated according to Eq.
pc = nv
1
v
e
(36) (18). This degradation–consolidation model is implemented in the open
source FEM calculator OGS. Detail of the numerical formulations can be
where αv and nv represent the model factor and θe represents the ef- found in the series publications of OGS [15–18].
fective water saturation, which has the following form: The solute transport, multi-phase flow, and deformation coupling
w wr are solved in OGS. The biochemical degradation is solved with a cal-
=
e
ws wr (37) culator developed by Zhejiang University (ZJU calculator) [7]. As
shown in Fig. 5, during the calculation, the concentration of each
where θwr is the residual water content and θws is the max water con- compound and the water content are delivered from OGS to the ZJU
tent. calculator. Then the reaction rate and the mass change of each com-
The parameters of the van Genuchten model, i.e. αv, nv and θws, pound, as well as the water/gas generation and the solid mass loss, are
tended to decrease with the increase of the landfill depth and the fill calculated according to Eqs. (15)–(17). After that, the results are
age, while θwr is decreased. This is mainly caused by the lower void
ratio and a larger amount of the small particles in deeper layers. In this 1E-09
study, the change of the WRC of the MSW is related to the change of
porosity. The parameters in the van Genuchten model are all expressed
1E-10
in functions of porosity:
Intrinsic permeability [m2]

v = 2.243n 0.820 (38)


1E-11
ws = 1.087n 0.072 (39)

wr = 0.486n + 0.551 (40) 1E-12

n v = 1.766n + 0.367 (41)


1E-13 Fitted curve
These functions are fitted using the WRC measurement data of the Measured data
in-situ MSW samples from different depths and ages (Fig. 4). This data
can be found in Xu et al. [10]. 1E-14
0.6 0.62 0.64 0.66 0.68 0.7 0.72 0.74 0.76 0.78 0.8
For the mass transport process of dissolved solutes (VFA and MB) Porosity [-]
the sorption effect of MSW is neglected and only the diffusion effect is
considered. The diffusion coefficient matrix in Eq. (26) is calculated Fig. 3. Fitted curve of the empirical relationship between the intrinsic perme-
from the free diffusion coefficient, the porosity, and the tortuosity, ability and the porosity of MSW.

6
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

100 water) with a depth z at age t and the initial time. The degradable solid
W1-1 Measured data(shallow layer)
90 W1-2 Measured data(Middle layer) mass Msd is the total dry mass of five degradable compositions (Si, i = 1,
2, 3, 4, 5). According to this definition, the hydrolysis degrees of the
Volumetric water content (%)

80 W1-3 Measured data(deep layer)


W1-1 Fitted curve(shallow layer) fresh MSW and the fully decomposed MSW are 0 and 100%, respec-
70 W1-2 Fitted curve(middle layer)
W1-3 Fitted curve(deep layer)
tively. This indicator takes the influence of sugars, proteins, and lipids
60
into consideration. In particular, the dry mass fraction of sugars in the
50 HFWC MSW can reach 40%.
40 The second indicator is the water release degree, which is defined as
30 the ratio between the current volumetric water content and the max-
imum volumetric water content in fresh MSW. The maximum volu-
20
metric water content is yielded from the sum of the initial water content
10 and the intra-particle water content. It can be expressed as follows:
0 z
1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04
(S6 (z , t ))dz
Suction (kPa) t 0
Uwater = z
Fig. 4. Comparison of the WRC for the empirical model and the measured data
of the in-situ samples. 0
(S (z, 0) +
6 w (z , 0) S1 (z , 0)
Mwater
MRDC ) dz (46)

where S6 is the volumetric water content, w (z , 0) is the intra-particle


water factor of the fresh MSW, and Mwater and MRDC are the molar mass
of water and rapid degradable cellulose. The water release degree can
be used to assess how much intra-particle water generation potential of
landfilled MSW has been already released.
The methane generation degree, defined as the ratio between the
already generated methane and the initial methane generation poten-
tial, can be used to assess the methane production potential of landfilled
MSW:
z
QCH4 (z , t )dz
t 0
UCH4
= z
PCH4 (z , 0)dz
0 (47)
Fig. 5. Solution diagram of the degradation–consolidation coupled model.

PCH4 (z , 0)
delivered back to OGS. The generated water/gas and the dissolved
S1 (z , 0) S (z , 0) S (z, 0) S (z , 0) S (z , 0)
compound from the ZJU calculator are treated as source terms in OGS. = 3MCH4 + 2 + 3 + 4 + 5
The changes of the material properties, i.e. stiffness, porosity, solid MRDC MSDC Msugar Mprotein Mlipids
water character curve, intrinsic permeability, and diffusion coefficient, (48)
due to the solid mass loss are updated after each computational itera-
where QCH4 (z , t ) is the MSW in the accumulated generated methane at a
tion. depth z until time t, and PCH4 (z , 0) is the methane generation potential
of fresh MSW at a depth z, which can be simply calculated based on the
5. Stabilization behavior of landfilled MSW chemical compositions of fresh MSW, and MCH4, MRDC, MSDC, Msugar,
Mprotein, and Mlipids are the molar masses of each chemical composition.
5.1. Indicators of stabilization behavior for landfilled MSW Because lignin is difficult to decompose, the ratio between cellulose
and lignin (RC/L) was often used to assess the degradation and stabili-
As presented in Section 2 and Fig. 1, with the solid mass loss during zation state of landfilled MSW. The RC/L value varies with the compo-
degradation, the generation of water, gas, and solute takes place, and sition of the MSW. For HFWC and LFWC MSWs, the initial RC/L values
the processes are evaluated with time. These processes change the are in the range of 3–4. A normalized RC/L value (β) was introduced
displacement, pore water pressure, pore gas pressure, and solute con- here to assess the stabilization state of MSW. The normalized RC/L (β) is
centration field in the landfilled MSW. defined as follows:
The degradation–consolidation model of landfilled MSW developed z z
herein can be used to assess the degradation process of landfilled MSW. RC L (z, t )dz (S1 (z, t ) + S2 (z , t )) SL (z )dz
Five indicators from different aspects are introduced, which are the =1 0
=1 0
z z
normalized ratio between cellulose and lignin, the hydrolysis degree, RC L (z, 0)dz (S1 (z, 0) + S2 (z , 0)) SL (z )dz
the intra-particle water release degree, the methane generation degree, 0 0 (49)
and the consolidation degree.
where RC/L(z,0) and RC/L(z,t) are the RC/L values at a depth z for fresh
The hydrolysis degree is defined as the ratio between the already
and aged MSW, respectively. S1 and S2 are the mass concentrations of
hydrolyzed solid mass and the initial degradable solid mass, which can
rapidly and slowly degradable cellulose and SL is the mass concentra-
be expressed as follows:
tion of lignin, which is assumed constant in this study. This normalized
z
indicator (β) in this study is similar to the indicators introduced above.
(Msd (z, 0) - Msd (z, t ))dz
t 0
It is 0 and 1 for fresh MSW and completely degraded MSW, respectively.
Uhyd = z The advantage of β in comparison to the other stabilization indicators is
Msd (z , 0)dz that it can easily be measured with the initial and current RC/L, which
0 (45)
can be investigated in a laboratory with the in-situ-sampled MSW and
where z is the depth of MSW layers in a landfill, and Msd (z , t ) and the fresh MSW.
Msd (z , 0) are the degradable solid masses (without the intra-particle The consolidation degree can be defined as follows:

7
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

t ut
Uset = × 100%
uend (50)
t end
where u is the surface settlement at time t and u is the surface set-
tlement when the MSW is completely degraded and excess pore pres-
sure is fully dissipated. Because the deformation of a landfill is affected
by the hydraulic conditions, the final settlement will be achieved when
the gas pressure in the MSW body is equal to the atmospheric pressure
and the water content is equal to or lower than the field water capacity.
The consolidation degree of landfilled MSW is similar to that of con-
ventional soils. The most important difference between them is the
degradation-induced deformation, which plays a significant role in the
consolidation process.

5.2. Case studies of HFWC and LFWC MSWs Fig. 6. Stabilization for a layer of landfilled MSW.

Two cases were studied to investigate the stabilization behavior of years, the hydrolysis degree of the HFWC MSW is already over 80%
HFWC (case 1) and LFWC (case 2) MSWs with the indicators defined (Fig. 7a). Under the same conditions, the hydrolysis degree of LFWC
above (Eqs. (45)–(50)). The chemical compositions of fresh MSW are MSW is only about 50–60%, and it is still less than 80% after 15 years.
shown in Table 1. All parameters of the degradation–consolidation The results of the water release degree (Fig. 7b) indicate that the
model are listed in Table 2. On the one hand, the biochemical de- total intra-particle water is released in the first year for HFWC MSW.
gradation related parameters for both cases (HFWC or LFWC MSWs) are For the case of LFWC MSW, the intra-particle water release amount is
the same, because the degradation process is only dependent on each much less than the HFWC MSW, and it is not presented here.
degradable chemical composition and the degradation environment. On As shown in Fig. 7c, the methane generation degree demonstrates
the other hand, the mechanical properties, i.e., the compression para- different trends for HFWC and LFWC MSWs. For both types of MSW, the
meters, WRC parameters, and permeability, are dependent on the state methane generation take place after the acid inhibition is relieved. For
of the MSW, such as the stress state, porosity, and solid grain-size dis- LFWC MSW, the methane generation rate is nearly linear on a loga-
tribution, which should be different between HFWC and LFWC MSWs. rithmic time scale. Meanwhile, for HFWC MSW, most methane gen-
However, in order to assess the degradation and stabilization behavior eration potential is released in two years due to its higher ratio of rapid
of HFWC and LFWC MSWs and focus on the impacts of the different degradable composition. The methane generation degree for HFWC
initial contents on the degradation and stabilization behavior, all used MSW reaches 0.8 after three years, then it slows down notably. Another
parameters in the degradation–consolidation model are the same both phenomenon should be noted. For HFWC MSW, the methane genera-
for HFWC and LFWC MSW, except the intra-particle water factor. The tion degree is less than 0.9 after 50 years, but this degree is close to 1.0
HFWC MSW’s intra-particle water factor is significantly higher than for LFWC MSW. This indicates that more than 10% of the methane
that of the LFWC MSW. The parameters were selected based on early generation potential is lost during the degradation of HFWC MSW,
studies [7,31–33]. which can be interpreted as the impact of the bottom drainage
In landfills the MSW was disposed layer by layer. The initial boundary. In the early stage of degradation, a high amount of leachate
thickness of a waste layer was 5–10 m. The waste in the same layer had is generated and drained through the bottom boundary. The flow-out
similar age and degradation degree. Thus, a 10-m-high one-dimensional leachate contained the production of the hydrolysis stage, i.e. VFA,
model was set up. The initial and boundary conditions are shown in which would be further transformed into methane if remains in the
Fig. 6. The initial gas pressure was 0 Pa and the initial water content MSW body. Thus, the generated methane for HFWC MSW is less than its
was 0.283. The top boundary was impervious for liquid but pervious for initial potential.
gas, and the gas pressure was fixed at 0 Pa. The bottom boundary was As shown in Fig. 7d, the results of β show a similar trend as the
imperious for gas but it had free outflow for liquid. There was no extra hydrolysis degree for HFWC and LFWC MSWs, because β and the hy-
loading on the top boundary, and the domain deformed under the self- drolysis degree are both defined according to the mass loss of the de-
weight stress. The calculation duration was 50 years. gradable solid compositions. Sugars, which took up a large proportion
These case studies focus on the stabilization behavior of HFWC and of the HFWC MSW, are neglected in the definition of β. Thus, β reaches
LFWC MSWs through the newly introduced stabilization degree in- 0.6 after two years, which is lower than the hydrolysis degree of HFWC
dicators. The summarized results are shown in Fig. 7. In the first three

Table 2
Parameters of degradation–consolidation model of landfilled MSWs.
Intra-particle water release factor Hydrolysis constant factor Inter pore water content impact parameters for hydrolysis
λw kH1 (day−1) kH2 (day−1) kH3 (day−1) kH4 (day−1) kH5 (day−1) θmin θmax
HFWC MSW LFWC MSW
130 10 3 × 10−3 2 × 10−4 3 × 10−3 3 × 10−3 3 × 10−3 12% 45%

Acid inhibition parameters for hydrolysis Methanogenesis parameters Acid inhibition parameters for methanogenesis Biomass death factor

Kh (kg m−3) nh Y kmmax (day−1) KS Km (kg m−3) nm Kd (day−1)

10 4 0.08 0.05 4 5 3 0.005


Compression parameters Porosity Intrinsic permeability WRC parameters
C′c0 C′c∞ εs∞ (σ′0) σ′0 (kPa) n ki (m2) av nv
0.2075 0.1 0.38 0.5 0.55 10−12 0.88 1.58
Free diffusion coefficient Tortuosity
D0 (m2 s−1) τ
10−5 1

8
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

Fig. 7. Calculated stabilization indicators of HFWC and LFWC MSWs (a. hydrolysis degree, b. water release degree, c. methane generation degree, d. β value, e.
uniformized porosity and f. consolidation degree).

MSW. the degradation solid mass loss and consolidation cause the increase
The porosity also has great impact on the coupled processes in and decrease of the porosity, respectively. However, the results show
landfilled MSW. As mentioned in Section 3 the hydraulic properties, i.e. decrease of the porosity from the beginning, which means the de-
permeability and water retention behavior, are dependent on the por- gradation induced consolidation has greater impact on the change of
osity. In Fig. 7e the normalized porosity of HFWC and LFWC MSW are porosity than solid mass loss. The porosity change is dominated by
presented. The definition of the normalized porosity had the form as: consolidation. Due to the rapid development of the consolidation of
HFWC MSW, the porosity is also decreased to a relative stable level
t nt after two years. Meanwhile, it takes nearly 50 years for LFWC MSW.
nnorm =
n0 (51)
The results of the consolidation degree (Fig. 7f) show that HFWC
where, nt is the porosity at time t, and n0 was the initial porosity. During MSW reaches a relative stable status after three years, which is much

9
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

landfills were investigated. As shown in Fig. 9, the stabilization lifespan


of HFWC MSW landfills can be divided into three stages, that is, (I)
rapid degradation, (II) slow degradation, and (III) post-stabilization.
The cut-off point between rapid and slow degradation is different
according to different stabilization indicators. The water release degree
and the consolidation degree reach an inflection point in less than two
years. Based on the hydrolysis degree and the β value, it takes more
than two years to reach the slow degradation stage. Because the me-
thanogenesis process takes place after the hydrolysis of the solid mass.
The hysteresis of methane generation degree can be observed.
During the rapid degradation stage, 80% of the solid mass is hy-
drolyzed, 100% of the intra-particle water is released, 60% of the LFG is
generated, and 80% of the compression is completed. The rapid de-
gradation stage lasts about 2–3 years, when β increases to 0.6. During
the slow degradation stage, over 90% of the solid mass is hydrolyzed,
90% of the compression is completed, and 80% of the LFG is generated.
This stage lasts about 20–30 years, when β approaches to 0.9. After that,
the HFWC MSW is almost fully stabilized.
Fig. 8. Comparisons between the calculated and measured β values of HFWC For different stabilization stages, the landfilled HFWC MSW has
MSW. significantly different degradation behaviors. During the rapid de-
gradation stage, a large amount of leachate is released and LFG is
generated rapidly. The high leachate level in the landfill body increases
earlier than LFWC MSW. The degradation induced consolidation is the slope failure risk and decreases the LFG collection efficiency,
mainly caused by the decomposition of the solid mass. According to the therefore certain engineering measures should be taken during the
results of the hydrolysis degree (Fig. 7a), more than 80% degradable rapid degradation stage. For the slow degradation stage, the rest of the
solid mass of HFWC MSW is decomposed in the first three years, which organic substances are difficult to degrade. Changing the degradation
is much faster than LFWC MSW. Then, it should be noted that the model environment from anaerobic to aerobic with air ventilation may sig-
has a drainage boundary for leachate on the bottom and LFG on the top. nificantly accelerate the degradation process and shorten the slow de-
The water table, i.e., the pore pressure, plays an important role in the gradation stage. In the post-stabilization stage, the MSW is almost fully
settlement development. The results indicate that when the leachate is degraded, while the LFG generation and the development of the landfill
drained out of the landfilled MSW layer, the settlement, i.e., the con- surface settlement are negligible, making the reuse of the field or the
solidation degree, of the HFWC MSW evolves quickly in the first landfilled MSW possible. However, environmental issues must be fur-
2–3 years. However, for LFWC MSW, due to the lower water content in ther investigated and evaluated before reuse of the field or the mate-
the fresh MSW, the trend of the consolidation degree is similar to the rials.
hydrolysis degree.
β is the normalized ratio of cellulose to lignin (RC/L), which can be 6. Conclusions
measured with standard methods in laboratory. The calculated β values
of the HFWC MSW from the above case studies and several field mea- A degradation–consolidation model for the stabilization behavior of
surements [34,35] are compared and presented in Fig. 8. The results landfilled MSW has been developed. This model considered (1) the
show a good agreement between the data sets. Both in-situ measured multi-physics process, including solid skeleton deformation, pore water
and calculated β indicate that the degradation develops rapidly in the flow, solute transport, pore gas flow, and biochemical degradation, (2)
first few years, and then slows down significantly. the biological degradation induced solid mass transformation of solids,
liquids, gases, and solutes, and (3) the variations of mechanical, hy-
draulic and solute transport properties due to the solid mass loss. For
5.3. Stabilization stages of HFWC MSW landfills
the numerical solution of the model, the deformation, multi-phase flow
and solute transport coupled processes and the biochemical degrada-
Since indicators describing the degradation process of landfilled
tion were solved by OGS and ZJU calculator, respectively. The model
MSW from different aspects have been previously defined, based on the
can be used to investigate the stabilization behaviors of landfilled MSW
results of previous cases, the stabilization stages of HFWC MSW
with different food waste contents.
Indicators for a stabilization assessment were introduced con-
sidering the solid mass loss, intra-particle water release, consolidation,
and LFG generation. The calculated results of the normalized ratio of
cellulose to lignin (β) had a good agreement with the measurements
from in-situ landfilled MSW. It indicated that the newly developed
degradation-consolidation model was capable for analyzing the stabi-
lization state of landfilled MSW. The stabilization behaviors of HFWC
and LFWC MSWs were compared. During the degradation, solid mass
loss and consolidation would cause the increase and decrease of the
porosity, respectively. The results showed monotonous decrease of
porosity, which meant the porosity change was dominated by con-
solidation for both HFWC and LFWC MSWs. The results indicated that,
due to a higher content of rapid degradable compositions, the stabili-
zation process of HFWC MSW developed more rapidly than LFWC
MSW.
Using these indicators, the stabilization process of HFWC MSW
Fig. 9. Stabilization stages of HFWC MSW landfills. landfills can be divided into three stages, which are the rapid

10
Y.M. Chen, et al. Computers and Geotechnics 118 (2020) 103341

degradation stage, the slow degradation stage, and the post-stabiliza- model for municipal solid waste in leachate recirculation. Waste Manage
tion stage. Understanding of the stabilization behavior of landfilled 2019;98:81–91. https://doi.org/10.1016/j.wasman.2019.08.016.
[14] Hubert J, Liu XF, Collin F. Numerical modeling of the long term behavior of
MSW could help to assess the stabilization state of MSW landfills more Municipal Solid Waste in a bioreactor landfill. Comput Geotech 2016;72:152–70.
easily in practice. [15] Kolditz O, Bauer S, Bilke L, Böttcher N, Delfs JO, Fischer T, et al. OpenGeoSys: an
The compatibility of the introduced model is not just limited to open-source initiative for numerical simulation of thermo-hydro-mechanical/che-
mical (THM/C) processes in porous media. Environ Earth Sci 2012;67(2):589–99.
landfilled MSW problems. It can also be used to analyze other problems https://doi.org/10.1007/s12665-012-1546-x.
such as methane hydration extraction and coal pyrolysis, in which cases [16] Kolditz O, Görke U, Shao HB, Wang W. Thermo-hydro-mechanical-chemical pro-
the solid skeleton is decomposed and the material properties are varied cesses in porous media. Springer Berlin Heidelberg; 2012.
[17] Kolditz O, Shao H, Wang W, Bauer S. Thermo-hydro-mechanical-chemical processes
due to biological or chemical reactions. in fractured porous media: modelling and benchmarking – closed-form solutions.
Springer International Publishing; 2015.
Declaration of Competing Interest [18] Wang W, Kosakowski G, Kolditz O. A parallel finite element scheme for thermo-
hydro-mechanical (THM) coupled problems in porous media. Comput Geosci
2009;35(8):631–1641. https://doi.org/10.1016/j.cageo.2008.07.007.
The authors declare that they have no known competing financial [19] Chen YM, Zhan LT, Gao W. Waste Mechanics and Sustainable Landfilling
interests or personal relationships that could have appeared to influ- Technology: Comparison Between HFWC and LFWC MSWs. In: Zhan L, Chen Y,
ence the work reported in this paper. Bouazza A, editors. Proceedings of the 8th International Congress on Environmental
Geotechnics Volume 1. ICEG 2018. Singapore: Environmental Science and
Engineering. Springer; 2019.
Acknowledgment [20] Yang QF. Laboratory research on Soil-water characteristic curve of municipal solid
waste under bio-mechanical effect Master Thesis China: Zhejiang University; 2016
[21] IPCC. Guidelines for National Greenhouse Gas Inventories. Hayama, Kanagawa,
The authors acknowledge the financial support from the research Japan: Institute for Global Environmental Strategies; 2006.
grant (No. 51988101 and 51508504) provided by the National Natural [22] Chen YM, Guo RY, Li YC, Zhan LT. A model for anaerobic degradation of municipal
Science Foundation of China and the research grant (No. solid waste. In: Proc. of Coupled Phenomena in Environmental Geotechnics Symp.,
Torino, Italy; 2013.
2012CB719806) provided by the National Basic Research Program of [23] Zumdahl SS, DeCoste DJ. Chemical principles. 7th ed. Belmont, USA: Brooks/Cole;
China. 2012.
[24] Vavilin VA, Lokshina LY, Jokela JPY, Rintala JA. Modeling solid waste decom-
position. Bioresour Technol 2004;94(1):69–81. https://doi.org/10.1016/j.biortech.
References 2003.10.034.
[25] Vavilin VA, Fernandez B, Palatsi J, Flotats X. Hydrolysis kinetics in anaerobic de-
[1] McDougall J. A hydro-bio-mechanical model for settlement and other behavior in gradation of particulate organic material: an overview. Waste Manage
landfilled waste. Comput Geotech 2007;34:229–46. 2008;28(6):939–51. https://doi.org/10.1016/j.wasman.2007.03.028.
[2] Tchobanoglous G, Theisen H, Vigil S. Integrated solid waste management: en- [26] Meima J, Naranjo NM, Haarstrick A. Sensitivity analysis and literature review of
gineering principles and management issues. McGraw-Hill Inc; 1993. parameters controlling local biodegradation processes in municipal solid waste
[3] Emkes H, Coulon F, Wagland S. A decision support tool for landfill methane gen- landfills. Waste Manage 2008;28:904–18.
eration and gas collection. Waste Manage 2015;43:307–18. https://doi.org/10. [27] Sanavia L, Pesavento F, Schrefler BA. Finite element analysis of non-isothermal
1016/j.wasman.2015.07.003. multiphase geomaterials with application to strain localization simulation. Comput
[4] El-Fadel M, Findikakis A, Leckie J. Numerical modelling of generation and transport Mech 2005;37(4):331–48. https://doi.org/10.1007/s00466-005-0673-6.
of gas and heat in landfills I. Model formulation. Waste Manage Res [28] Fredlund DG, Rahardjo H. Soil mechanics for unsaturated soils. New York: Wiley;
1996;14:483–504. 1993.
[5] Barlaz MA. Carbon storage during biodegradation of municipal solid waste com- [29] Chen YM, Zhan TLT, Wei HY, Ke H. Aging and compressibility of municipal solid
ponents in laboratory-scale landfills. J Global Biogeochem, Cycles wastes. Waste Manage 2009;29:86–95.
1998;12(2):373–80. [30] van-Genuchten MT. A closed-form equation for predicting the hydraulic con-
[6] He PJ, Feng SW, Shao LM. Municipal solid waste management. Science Press; 2003. ductivity of unsaturated soils. Soil Sci Soc Am J 1980;44(5):892–8.
[7] Chen YM, Guo RY, Li YC, Liu HL, Zhan LTL. A degradation model for high kitchen [31] Chen YM, Zhan LTL, Li YC. Biochemical, Hydraulic and Mechanical Behaviors of
waste content municipal solid waste. Waste Manage 2016;58:376–85. Landfills with High-Kitchen-Waste-Content MSW. In: 7th ICEG, Melbourne,
[8] White JK, Beaven RP. Developments to a landfill processes model following its Australia; 2014.
application to two landfill modelling challenges. Waste Manage [32] Gao W, Chen YM, Zhan LT, Bian XC. Engineering properties for high kitchen waste
2013;33(10):1969–81. content municipal solid waste. J Rock Mech Geotech Eng 2015;7(6):646–58.
[9] Chen YM, Ke H, Fredlund DG, Zhan LT, Xie Y. Secondary compression of municipal https://doi.org/10.1016/j.jrmge.2015.08.007.
solid wastes and a compression model for predicting settlement of municipal solid [33] Xu XB, Zhan TLT, Chen YM, Guo QG. Parameter determination of a compression
waste landfills. J Geotech Geoenviron Eng 2010;136(5):706–17. model for landfilled municipal solid waste: an experimental study. Waste Manage
[10] Xu XB, Zhan TLT, Chen YM, Beaven RP. Intrinsic and relative permeabilities of Res 2015;33(2):199–210.
shredded municipal solid wastes from the Qizishan landfill, China. Canadian [34] Chen YM, Liu XC, Xu WJ, Li YC, Lan JW, Zhan LT, et al. Analysis on stabilization
Geotech J 2014;51(11):1243–52. characteristics and exploitability of landfilled Municipal Solid Waste: Case of a
[11] Kindlein J. Numerical modelling of multiphase flow and transport processes in typical landfill in China. Scientia Sinica Technologica 2018;49(2):99–211. https://
landfills. Waste Manage Res 2006;24(4):376–87. https://doi.org/10.1177/ doi.org/10.1360/N092018-00140.
0734242X06065506. [35] Liu H. Research on biodegradation-consolidation-solute migration coupled beha-
[12] Yu L, Batlle F, Lloret A. A coupled model for prediction of settlement and gas flow in viors of municipal solid waste and landfill stabilization. Zhejiang University, China;
MSW landfills. Int J Numer Anal Meth Geomech 2010;34:1169–90. 2016. Ph.D. Thesis.
[13] Feng SJ, Fu WD, Zhou AN, Lyu F. A coupled hydro-mechanical-biodegradation

11

You might also like