Sunteți pe pagina 1din 11

Research Article

Received: 28 May 2012

Revised: 8 July 2012

Accepted: 9 July 2012

Published online in Wiley Online Library: 28 August 2012

(wileyonlinelibrary.com) DOI 10.1002/jctb.3909

A thermodynamic equilibrium consideration


of the effect of sodium ion in acetoclastic
methanogenesis
Sung T. Oha and Alastair D. Martinb
Abstract
BACKGROUND: Sodium ion, a common constituent of food, is empirically observed to inhibit microbial activity (reduce biokinetic rate) in anaerobic digestion. However such reductions may arise from a number of sources, including loss of enzyme
activity and loss of thermodynamic driving force. In this work the theoretical thermodynamic approach was used to describe
the inhibitory effects of sodium ion in acetoclastic and hydrogen utilizing methanogenesis.
RESULTS: This work takes a modeling approach to investigate thermodynamic limitation by free energy, enthalpy and
entropy analysis. Simple free energy analysis provides no evidence for thermodynamic limitation arising from even very high
concentrations of sodium ion. However, entropic analysis suggests that increasing concentrations of sodium ion result in the loss
of spontaneity. The loss of spontaneity arising from the presence of sodium ions is not directly related to the rise in equilibrium
pH, which may also occur in the presence of sodium ions. The results also indicate that hydrogenophilic methanogens are less
likely to be affected by the presence of sodium ion than their acetoclastic equivalents.
CONCLUSION: From the thermodynamic perspective, the model advises that when the supply of thermal energy is sufficient,
sodium inhibition under anaerobic conditions can be virtually completely overcome. The thermodynamic model also provides
a design tool with which the stability of methanogenesis in response to feedstock mixtures containing sodium ion and a carbon
source such as HAc can be investigated.
c 2012 Society of Chemical Industry

Keywords: thermodynamics; anaerobic digestion; sodium inhibition; methanogenesis

NOTATION
ai
fi
F
G
H
I
K
MWi
Mi
mi
N
P
R
R

Activity of component i
Fugacity of component i
Faradays constant
Change of Gibbs free energy in reaction (kJ)
Change of enthalpy in reaction (kJ)
Ionic strength, ML3 (i.e. mole L1 )
Equilibrium constant
Molecular weight of species i
Molarity of species i (moles of solute per litre of solution)
Molality of species i (moles of solute per kg of water)
Total number of gas mole in gas phase
sat
saturated water pressure, atm
Pressure; PW
Universal gas constant (italic letter)
The initial sodium ion fraction (R=NaAc/(HAc+NaAc))
in the solutes
Re(W) Number of water moles in equilibrium state
S
Change of entropy in reaction
T
Temperature, K
Mole fraction for species i in liquid phase
xi
Mole fraction for species i in gas phase
yi
Ion charge for ion species i
zi

834

Greek letters
i Activity coefficient of component i
i Fugacity coefficient of component i
i Chemical potential of component i
vi Stoichiometric coefficient of component i

J Chem Technol Biotechnol 2013; 88: 834844

Gas molar volume, L3 M1 (i.e. liter per mole)

Superscript

H
redox
sat
Vapour, Liquid
+,
Subscript
i, j, and k
W

Standard property
Infinite dilution
Henrys law
Oxidation-reduction
Saturated property
Vapour or Liquid phase
Ion charge of + or
Components or half reactions
Solvent (water)

INTRODUCTION
Many industrial agricultural and food processing waste streams
contain significant concentrations of dissolved salts. Wastewaters

Correspondence to: Sung T. Oh, Department of urban engineering, London


South Bank University, London, SE1 0AA, UK. E-mail: ohs@lsbu.ac.uk

a Department of urban engineering, London South Bank University, London, UK


b School of Chemical Engineering & Analytical Sciences, University of Manchester,
Manchester, UK

www.soci.org

c 2012 Society of Chemical Industry




Effect of sodium ion in acetoclastic methanogenesis

www.soci.org

J Chem Technol Biotechnol 2013; 88: 834844

for the empirically observed inhibition such as exhaustion of the


Gibbs free energy and entropic effects such as loss of spontaneity.

MODEL DEVELOPMENT
Anaerobic micro-organisms in the digestion process use chemical
energy released by catabolism for growth and maintenance and
in the process produce end-products such as shorter chain fatty
acids, carbon dioxide and methane. For the catabolic reactions
to provide the chemical energy necessary, energy must flow from
a higher to a lower potential. The energy released is employed
to drive the anabolic reactions of the micro-organisms. The
bioenergy economy has been adopted to describe a two-phase
batch anaerobic digestion system with conservation of total
energy and minimization of biochemical convertible energy. The
partial degradation of a substrate by one microbial species has
been assumed to be directly linked to the removal of intermediate
products by other species. As the anaerobic micro-organisms
acclimatise to a given environment they are able to make a
consortium in which their individual metabolisms become highly
interlinked. This microbial consortium permits the system to
closely approach the thermodynamic equilibrium state (G = 0).9
In this work a network of redox equilibria has been established
to describe the metabolic reactions in an acclimatised anaerobic
consortium of acetoclastic and hydrogen utilising methanogens
linked via the exchange of electrons and protons. This approach is
consistent with the interspecies electron and proton hopping theory of Marcus.10 The phase equilibria of the end-products (CO2 , CH4
and H2 O) have been included to investigate the loss or uptake of
thermal (heat) energy by vaporisation/solubility of end-products as
well as the equilibria of acid dissociation (including aqueous CO2 ).
The thermodynamic equilibrium was simulated in an isothermal
and isobaric (298.15 K and 1 atm) batch digester. The initial substrates considered were a mixture of sodium acetate (CH3 COONa)
with acetic acid (CH3 COOH). The anaerobic digestion was assumed
to contain a fully acclimatized methanogenic consortium. A
two-phase (solution and gas) digester was assumed. The simple
stoichiometric model of methanogenesis considers only three
initial materials: liquid acetic acid, aqueous sodium acetate
and liquid water. In addition it also considers the quantitative
expressions that relate them to the 18 products and intermediates.
The thermodynamic framework comprises two parts. The first
includes equilibrium expressions that relate the relative activities
and or fugacities of all 21 species in the two phases to each other.
The second consists of the equations of state that quantify the relationships between activity in the liquid phase and fugacity in the
gas phase to the observable concentrations and partial pressures
via the respective activity and fugacity coefficients. The completed model consists of a set of highly non-linear simultaneous
equations which was solved using an algorithm based on the NewtonRaphson technique.11 13 The following sections describe
the development of the individual components of the model.
Stoichiometric model
Stoichiometric expressions were developed in the conventional
manner to quantify the possible relationships between the
initial materials and the various products and intermediates. 15
expressions describe the material relationships of the species
between the aqueous and gas phases, the ionization of the
electrolytes and the redox reactions. These relationships between
these initial species and the 18 products and intermediates are
set out in Table 1.

c 2012 Society of Chemical Industry




wileyonlinelibrary.com/jctb

835

arising from the rayon spinning and cellophane casting processes,


the production of sauerkraut, tomato and shellfish canning and
dairying, for example, all contain significant concentrations of
sodium salts.1 The sodium salts present in these waste streams
can be a significant source of inhibition during their anaerobic
treatment. Despite the negative effects of sodium salts on the
metabolic activity of methanogenic bacteria, the engineer must
still consider the anaerobic microbial treatment of these wastes
as an alternative to the highly sodium sensitive aerobic treatment
methods.2 Under anaerobic conditions low concentrations of
sodium ion, less than approximately 400 mg L1 are observed to
have beneficial effects, however at higher concentrations sodium
ion is observed to inhibit mesophilic methanogens.3,4 This latter
observation may be readily explained as the higher the sodium
ion concentration in a digester, the harder microorganisms have
to work to transport water into their cells to degrade organics due
to the raised osmotic pressure of the solution. McCarty4 reported
sodium ion concentrations in the range of 100200 mg L1 to be
beneficial for the growth of mesophilic anaerobes. Kugelman and
Chin5 also reported that the optimal sodium ion concentration
for mesophilic acetoclastic methanogens in waste treatment
processes was 230 mg L1 . Patel and Roth6 reported that the
optimal growth conditions with respect to sodium ion, for
mesophilic hydrogen utilizing methanogens, was 350 mg L1 .
However, McCarty4 reported that sodium ion concentrations
ranging from 3500 to 5500 mg L1 to be moderately inhibitory and
above 8000 mg L1 to be strongly inhibitory to methanogens at
mesophilic temperatures. Rinzema et al.2 also reported that for an
anaerobic granular biomass at mesophilic temperatures, sodium
ion concentrations of 5000, 10 000, and 14 000 mg L1 caused 10,
50 and 100% inhibition of methanogens, respectively, at neutral
pH, while Ngian and Pearcea7 observed methane formation to fail
when the concentration of sodium ion was at 4600 mg L1 in a
batch type anaerobic digester. This may be attributed to inhibition
by sodium ion although other explanations are also possible.
Bashir and Matin8 observed the accumulation of total volatile
fatty acids in a continuous digester from 300 to 3500 mg L1
in association with a sodium ion concentration of 6000 mg L1
in whey degradation. Grasso et al.3 also investigated the effect
of sodium ion on anaerobic digestion and found that in the
concentration range from 100 to 200 mg L1 it was stimulatory, in
the range 3500 to 5000 mg L1 it was weakly inhibitory and above
8000 mg/l it was strongly inhibitory, in close agreement with the
work of McCarty.4 While much of the work describes inhibition
qualitatively it suggests a pattern of enhancement between 0 and
approximately 400 mg L1 of sodium ion followed by a decline
through bands of weak and strong inhibition to the eventual
elimination of activity when the sodium ion concentration exceeds
14 000 mg L1 .
Inhibition by high concentrations sodium ion has been the
subject of extensive empirical observations and qualitative
deductions, however, there has been no widely published work
which attempts to explain the experimental observations in terms
of the relative proportions of the solutes and the fundamental
thermodynamics of the anaerobic digestion process. In this work
an entirely theoretical approach has been adopted in which the
fundamental laws of thermodynamics are applied to a simplified
stoichiometric model of methanogenesis in anaerobic digestion.
The addition of sodium ion to the stoichiometric model and thermodynamic framework facilitates the quantitative investigation
of its influence on the energy fluxes and position of equilibrium.
The results will enable the establishment of fundamental causes

www.soci.org

Table 1.

ST Oh, AD Martin

Example stoichiometric model for methanogenesis

Phase transitions vapor/water

Electrolyte relations

Redox half reactions

H2 O (g) = H2 O (l)
H2 (g) = H2 (aq)
CO2 (g) = CO2 (aq)

H2 O (l) = H+ + OH

H2 (aq) = 2H+ + 2e

CO2 (aq) + H2 O = H2 CO3


CO2 (aq) + H2 O = HCO3 + H+
CO2 (aq) + H2 O = CO3 2 + 2H+
NaHCO3 (aq) = Na+ + HCO3
NaCO3 (aq) = Na+ + CO3 2
Na2 CO3 (aq) = 2Na+ + CO3 2

CH4 (g) = CH4 (aq)

CH4 (aq) + 2H2 O = CO2 (aq) + 8H+ + 8e


C2 H4 O2 (aq) + 2H2 O = 2CO2 (aq) + 8H+ + 8e

C2 H4 O2 (aq) = C2 H3 O2 + H+

Equilibrium relationships
Equilibrium models superimposed on the stoichiometric foundations are based on the fundamental definition of thermodynamic
equilibrium, which is defined as the equalisation of chemical
potentials between reactants and products:

(1)
vi i = 0
where vi is stoichiometric coefficient of component i and the
chemical potential i is expressed as the standard chemical
potential i and activity ai of component i under ideal conditions:

i = i + RT ln ai

The definition of K in Equation (3) can be applied to each


equilibrium in the anaerobic digestion process. The set of equilibrium relationships is categorized into three types: vapor/liquid,
electrolyte ionisation and redox.
Vapor/liquid equilibrium
The vapor/liquid equilibrium is defined by the chemical potentials
for the component i in the two phases:
=

Liquid
i
Vapour

Vapour

Vapour

= i

+ RT ln fi

Water
Carbon dioxide

Acetic acid

Stoichiometric reactions

log(Ka )

H2 O(l) = OH + H+
CO2 (aq) + H2 O = H2 CO3
CO2 (aq) + H2 O = HCO3 + H+
CO2 (aq) + H2 O = CO3 2 + 2H+
NaHCO3 (aq) = Na+ + HCO3
NaCO3 (aq) = Na+ + CO3
Na2 CO3 (aq) = 2Na+ + CO3 2
C2 H4 O2 (aq) = C2 H3 O2 + H+

14.00
4.38 102
6.37
16.70
3.68 102
1.27
4.21 102
4.76

log(Ka ) at 298.15 K is calculated by Pourbaixs method.

non-volatile due to the very low Henrys coefficients and therefore


only carbon dioxide, methane, hydrogen and water can exist in
gas phase. Henrys law coefficients relevant to anaerobic digesters
are shown in Oh and Martin.12
Electrolyte ionisation equilibrium
In the solution phase, the chemical potential i of the component

i is the same as the summation of ion potentials (+


j and k ) of its
constituent ions (j and k):

vi i = vj +
j + vk k

(7)

Substituting Equation (2) into Equation (7) yields the following


relationship describing the electrolyte equilibrium, where equilibrium constant corresponds to dissociation constants Ka in
anaerobic digesters:

(5)

Substituting Equations (2) and (5) into Equation (4) yields the
following relationship describing the vapour/liquid equilibrium,
where the constant K H corresponds to the Henrys law coefficient:


Vapour
Liquid
i
)
(i
fi
H
KiW (atm/mole) = exp
=
(6)
RT
ai fi

836

H
The subscript i-W indicates that the Henrys law coefficient KiW
is for solute i in solvent (water). It is assumed that acetic acid is

wileyonlinelibrary.com/jctb

Electrolyte
equilibrium
constants

(4)

is described by
where vapor phase chemical potential i
fugacity fi based on an equation of state (EoS) model:
i

Electrolyte equilibrium constants in anaerobic digesters

(2)

where R is universal gas constant and T is absolute temperature.


By substituting Equation (2) into Equation (1) the fundamental
definition of the equilibrium constant, K is obtained:


vi i

v
i
ln K =
=
ln ai i
(3)
RT

Vapour
i

Table 2.


Ka = exp

vi i vj j + vk k
RT


=

vj vk
(a+
j ) (ak )
v

ai i

(8)

Table 2 shows the stoichiometric reactions and their equilibrium


constants and their relationships are governed by Equation (8).
Redox equilibrium
The reduction and oxidation (redox) reactions are the key steps
in anaerobic digestion and are conventionally described in the

c 2012 Society of Chemical Industry




J Chem Technol Biotechnol 2013; 88: 834844

Effect of sodium ion in acetoclastic methanogenesis

Table 3.

www.soci.org

Mass balance for methanogenesis

Mass balances
N(18.02yW + 2yH2 + 44yCO2 + 16yCH4 )+

mH+ + 17mOH + 2mH2 + 16mCH4 + 44mCO2


Re(W) +62mH CO + 61m

HCO3 + 60mCO3 2 + 59.05mAc


2 3
55.494
+60.05mHAc + 23mNa+ + 84mNaHCO3 + 83mNaCO3 + 106mNa2 CO3
+ 18.02Re(W) = 1000

mCH4 + mCO2 + mH2 CO3 + mHCO3


Re(W)

N(yCH4 + yCO2 ) + 55.494 +mCO3 2 + 2mAc + 2mHAc


mNaHCO3 + mNaCO3 + mNa2 CO3
= CO2 + 2HAc + 2NaAc


mH+ + mOH + 2mH2 + 4mCH4
Re(W) +2m
N(2yW + 2yH2 + 4yCH4 ) + 55.494
H2 CO3 + mHCO3 + 3mAc + 4mHAc
+mNaHCO3
+2Re(W) = 2H2 O + 4HAc + 3NaAc

mOH + 2mCO2 + 3mH2 CO3 + 3mHCO3


Re(W)

N(yW + 2yCO2 ) + 55.494 +3mCO3 2 + 2mAc + 2mHAc


+3mNaHCO3 + 3mNaCO3 + 3mNa2 CO3
+Re(W) = H2 O + 2CO2 + 2HAc + 2NaAc
Re(W)
55.494 (mNa+ + mNaHCO3 + mNaCO3 + 2mNa2 CO3 ) = Na = MNaAc
yW + yH2 + yCO2 + yCH4 = 1
mOH + mHCO

+ 2mCO

+ mAc + mNaCO

Overall mass balance

Carbon Balance

Hydrogen Balance

Oxygen Balance

Sodium Balance
Vapour Balance

= mH+ + mNa+

Ion Charge Balance

mi is expressed as molality of species i, Re(W) is the number of moles for residual water, N is total number of moles in vapour phase. H2 O, CO2 , NaAc
and HAc (acetic acid) are the initial number of moles.

form of coupled half reactions, linked through the exchange of


electrons. The half reactions and redox equilibrium relationships
relevant to acetoclastic methanogenesis are developed in Oh and
Martin12 and are used here without modification.
Relationship of activity and fugacity with concentration and
partial pressure
Aqueous phase
Activity ai and observable concentration mi in any real solution
may be related by an activity coefficient i :14
ai = i mi

(9)

For the liquid phase there are three types of model for
description of the activity coefficients that can be considered: Ideal
model; DebyeHuckel
model and Pitzer model. These models are

equivalent to the conventional equations of state used to quantify


the analogous relationships in the gas phase and have limited
ranges of applicability related to the ionic strength of the aqueous
phase. These limitations together with their respective models
for activity coefficients are discussed in Oh and Martin.12 Their
parameter values are re-used in this work.

Overall equations and mass balance


The equilibrium model requires a strict mass balance condition.
Component balances were constituted from the stoichiometric
model for the vapour phase, ion charge, and elements carbon,
hydrogen, oxygen and sodium, while the overall mass balance is
given by Equation (15):


Re(W) 
N
yi MWi +
mi MWi + 1000 = 1000(g) (11)
55.494
i

i=H2 O

where MWi is the molecular weight of species i. N is total mole of


vapour phase in the system. Re(W) is the number of water moles in
solution, where the total mass of system is conserved to 1000 (g).
The vapour balance is defined such that the summation of mole
fractions in gas phase yi is unity. The solution phase is similarly
defined such that the summation of ion charge zi for ion species i
is zero at equilibrium (electro neutrality):

yi = 1
(12)


zi mi = zero

(13)

Gas phase
Fugacity fi and observable partial pressure yi P may be related in
a manner analogous to activity and concentration in the liquid
phase via the fugacity coefficient i .
fi = i yi P

(10)

J Chem Technol Biotechnol 2013; 88: 834844

Solution algorithm
The completed model describing the methanogenic equilibrium
comprises a set of 21 non-linear simultaneous equations. The

c 2012 Society of Chemical Industry




wileyonlinelibrary.com/jctb

837

where fugacity coefficient i is derived from a suitable equation


of state (EoS). Oh and Martin12 used the Nakamura et al. EoS15
to describe the mixture of non-polar and polar gases found in
anaerobic digesters. This work uses the same EoS, thus ensuring
consistency between the results.

The carbon, hydrogen, oxygen and sodium balances are defined


by the elemental mass based on the number of moles of each
element between the initial and equilibrium states. The elemental
mass balance of methanogenesis is shown in Table 3.
Table 4 illustrates the set of equilibrium relationships required
to complete the description of the anaerobic, methanogenic
degradation of acetic acid in the presence of sodium ion.

www.soci.org

Table 4.

ST Oh, AD Martin

Overall equilibria equations for methanogenesis

Vapour-liquid (water) equilibria

P Liquid
sat sat
W

yW W P = W PW exp
dp
RT
Psat
W

P
W
H
H
2
yH2 H2 P = aH2 KH W exp
dp
RT
2
Psat
W

P
CO2 W
H

yCO2 CO2 P = aCO2 KCO W exp


dp
RT
2
Psat
W

P
W
CH
H
4
yCH4 CH4 P = aCH4 KCH W exp
dp
RT
4
sat
P

H2 O (g) = H2 O (L)
H2 (g) = H2 (aq)
CO2 (g) = CO2 (aq)
CH4 (g) = CH4 (aq)

Electrolyte equilibria
H2 O = H+ + OH

KW = aH+ aOH
aH CO
KCO2 = a2 3
CO2
aHCO3 aH+
KH2 CO3 =
aCO2
aCO 2 a2H+
3
KHCO3 =
aCO2
aAc aH+
KHAc = a
HAc
aNa+ aHCO3
KNaHCO3 = a
NaHCO3
aNa+ aCO3
KNaCO3 = a
NaCO3
a2 Na+ aCO 2
3
KNa2 CO3 =
aNa2 CO3

CO2 (aq) + H2 O = H2 CO3


CO2 (aq) + H2 O = HCO3 + H+
CO2 (aq) + H2 O = CO3 2 + 2H+
C2 H4 O2 (aq) = C2 H3 O2 + H+
NaHCO3 (aq) = Na+ + HCO3
NaCO3 (aq) = Na+ + CO3
Na2 CO3 (aq) = 2Na+ + CO3 2

Oxidation-reduction equilibria
yH4 2 4H2 yCO2 CO2 P4
redox
= KCH
yCH4 CH4
4
2
yH4 2 4H2 yCO
2 P6
redox
2 CO2
= KHC2
aHAc

CH4 (g) + 2H2 O = CO2 (g) + 4H2 (g)


C2 H4 O2 + 2H2 O = 2CO2 (g) + 4H2 (g)

sat
K H (the Henrys law constant); subscript HAc and Ac are represented as acetic acid and acetate ion respectively. PW
is saturated vapour pressure;
is molar volume; KW (water dissociation constant); K redox (redox potential equilibrium constants).

algorithm based on a NewtonRaphson technique developed by


Oh and Martin12 was used to find the solution.

RESULTS AND DISCUSSION

838

Equilibrium of methanogenesis
Oh and Martin12 studied the overall equilibrium of the
methanogenic, catabolic reactions under isothermal psychrophilic
(298.15 K) and isobaric (1 atm) conditions. The model gave a
prediction of the maximum chemical energy transport overall in the methanogenic process. Catabolic reactions in consorted methanogens (i.e acetoclastic and hydrogen utilizing
methanogens) convert energy-rich compounds to simpler but
still energy-rich compounds. The consortium enables the organisms to use the maximum amount of the available energy for their
growth and maintenance. They showed that the overall catabolic
reactions of methanogens can almost completely decompose
acetic acid (HAc) into end-products (methane and carbon dioxide). The end-products then expand into the vapour phase due to
their high volatility. Their results are consistent with Le Chateliers
principle as applied to a methanogenic process. The concentration
of protons (pH) controls the solubility of carbon dioxide and affects
both pCH4 in the gas phase and acetic acid (HAc) degradation.

wileyonlinelibrary.com/jctb

They showed that the solubility/vaporisation of carbon dioxide


and methane has an influence on the overall enthalpy change in
the system. In particular, the vaporisation of the carbon dioxide
requires much energy, which can be sufficient to make the overall
exothermic system endothermic. Their thermodynamic analysis
highlighted that the generation of the vapour phase from the dissolved species in anaerobic digestion sinks more thermal energy
than evolved during the exothermic methane formation process.
They suggested that either gas collection from the vapour phase
at low pressure or thermophilic anaerobic digestion may result in
even better degradation of acetic acid and methane formation,
but will require the supply of much more external thermal energy
to the system. Practically, it also suggests that the neutral pH
(around pH 78) in a strongly buffered solution may prevent the
vaporisation of carbon dioxide from the digested solution and
could render overall catabolism exothermic.
Methanogenic equilibrium in the presence of sodium ion
In this paper, the overall thermodynamic equilibrium of the
catabolic methanogenic processes in the presence of sodium
ions is modeled. The simulated condition is 1 kg in an isothermal
and isobaric (298.15 K and 1 atm) two-phased batch digester.

c 2012 Society of Chemical Industry




J Chem Technol Biotechnol 2013; 88: 834844

Effect of sodium ion in acetoclastic methanogenesis

www.soci.org

0.03

Residual HAc/ Initial HAc (mol./mol.)

1.0E-05
1.0E-06
1.0E-07
1.0E-08
1.0E-09
1.0E-10
1.0E-11
1.0E-12
1.0E-05

1.0E-04

1.0E-03
1.0E-02
1.0E-01
Initial x (HAc) 0.03

R -0(mol./mol.)
R -0.2(mol./mol.)

R -0.001(mol./mol.)
R -0.5(mol./mol.)

1.0E+00

R -0.01(mol./mol.)
R -0.8(mol./mol.)

R -1(mol./mol.)

Figure 1. Fractional residual HAc in methanogenesis at 298.15 K and 1 atm for a variety of initial sodium ion fractions (R). Vertical dotted bars labeled
0.03 (mol L1 ) indicate the HAc equivalent of a nominal 10% solids digester feed.

The change in the equilibrium state is considered in response


to a range of initial sodium concentrations under a conserved
system mass of 1 kg. Thus, as the initial sodium proportion
(R=NaAc/(HAc+NaAc)) in the solutes is increased, the initial moles
of solvent water (H2 O) is decreased to conserve the system mass
of 1 kg (18.02H2 O + 60.0HAc + 82.0NaAc = 1000 (g)). When R
= 1.0, initial, solution phase HAc=0.0 (mol L1 ) and when R=0.0,
initial NaAc = 0.0 (mol L1 ). Seven initial mole proportions (R =
0.001, 0.01, 0.1, 0.2, 0.5, 0.8 and 1.0) of solution phase sodium
ion were considered in the conserved mass. A crystallization area
(crystal activity of NaHCO3 (NaHCO3 = 1) or Na2 CO3 (Na2 CO3 = 1))
is shown, delimited by a dashed line in each of the relevant figures.
In this region the presence of a solid phase renders the twophase model developed here invalid hence no results are shown.
The model results show (a) the thermodynamic possibility for
decomposition of HAc, (b) relation between pH and solubilities of
end-products, and (c) enthalpy and entropy driving force changes
at thermodynamic equilibrium.
Biodegradability of HAc
At equilibrium, the substrate, HAc is almost completely decomposed and CO2 and CH4 are produced. Figure 1(a) shows that the
equilibrium HAc concentrations from methanogenesis have very
small values, which decrease further as the initial mole fraction
(x) of HAc is increased. It also shows that as initial sodium proportion (R) in the solutes is increased, the residual HAc/initial HAc
increases and that this effect is stronger at higher x. This result
suggests that the degradation of HAc may be thermodynamically
inhibited by the sodium ion for the whole range of R; despite this,
the low values of residual HAc mean that for practical purposes
the methanogenesis of acetic acid proceeds to completion for all
initial mole fractions (x) of HAc.

J Chem Technol Biotechnol 2013; 88: 834844

c 2012 Society of Chemical Industry




wileyonlinelibrary.com/jctb

839

pH, gas contents and yields


Measurement of gas production is widely used for estimating the
performance of methanogenic processes. As expected the moles

of HAc decomposed correspond to half of the summation of


the moles of carbon dioxide (CO2 ) and methane (CH4 ) produced.
The majority of CO2 and CH4 produced was transferred into
the vapour phase. Figure 2(a) confirms that CH4 and CO2 were
major constituents in the vapour phase, where the pCH4 (partial
pressure of methane) was higher than that of CO2 due to higher
relative solubility of CO2 . The solubility of CO2 is, however, strongly
dependent on pH. The addition of sodium ion to the initial solution
induced high pH at methanogenic equilibrium and consequently
high CO2 solubility. This, in turn, results in low pCO2 and high pCH4
as shown in Fig. 2(a). The deduction is supported by Fig. 2(b),
which shows the distribution of CO2 between the phases as
mol. of vapor CO2 over mol. of CO2 produced. The results show
that as the initial mole fraction x of HAc is increased, the initial
pH decreases and the carbonates are transferred into CO2 gas.
Consequently, at high x, the pCH4 converges with that of pCO2 .
However, Fig. 2(b) also shows that as R is increased, the phase
distribution of CO2 is decreased. This decrease is consistent with
the rising pH shown in Fig. 2(c). Figure 2(c) shows the pH value
at the equilibrium of HAc degradation. In the base case (R=0.0),
the pH value decreases from 4.75 (pKa of HAc) to a constant
value of approximately 4.07 according to increasing initial x of
HAc. Figure 2(c) also shows that as the initial mole fractions
(R) of sodium are increased, the pH values at the equilibrium
of HAc degradation are dramatically increased to approximately
pH 8.4. This value is the equilibrium pH of HCO3 and CO3 2 .
The model result is consistent with Kugelman and Chins5 work,
which observesthat sodium concentration increases pH value and
methane content in the gas phase.
Figure 2(d) shows that the summation of yields (moles of CO2
plus CH4 produced per mole of HAc decomposed) is consistently
2. For the pure methanogenesis (R=0), individual yields CO2 and
CH4 were both close to one. Figure 2(d) also shows that as R was
increased, the yield of CO2 was slightly increased and the CH4
yield was slightly decreased at low x (<0.01). The small deviations
above and below yields of 1 for CO2 and CH4 respectively can be
attributed to the production of small quantities of hydrogen from

www.soci.org

ST Oh, AD Martin

0.03
(a)

0.03

(b)

0.9

0.9
0.8

CH4

(NaHCO3)=1

0.7
Partial pressure

Vapour/Total CO2 (CO2)


(mol./mol.)

0.8

0.6
0.5

(NaHCO3) <1

0.4
0.3

CO2

0.2

0.6
0.5
0.4
0.3

(NaHCO3) =1

0.2

0.1
0
1.0E-05

(NaHCO3) <1

0.7

0.1
1.0E-04

1.0E-03

1.0E-02

1.0E-01

0
1.0E-05

1.0E-00

1.0E-04

1.0E-03 1.0E-02
Initial x (HAc)

Initial x (HAc)
(c)

1.0E-01

1.0E-00

1.0E-03 1.0E-02 1.0E-01


Initial x (HAc) 0.03

1.0E-00

(d)
9

1.00001
(NaHCO3) =1

(NaHCO3) <1

Yield (mole/mole)

1.000005

pH

1
CH4
0.999995

4
1.0E-05

CO2

1.0E-04

1.0E-03 1.0E-02 1.0E-01


Initial x (HAc) 0.03

1.0E-00

0.99999
1.0E-05

R -0(mol./mol.)

R -0.001(mol./mol.)

R -0.2(mol./mol.)

R -0.5(mol./mol.)

1.0E-04

R -0.01(mol./mol.)
R -0.8(mol./mol.)

R -1(mol./mol.)
Figure 2. Partial pressures at 1 mol L1 of HAc (a), total CO2 phase distribution (b), equilibrium pH value (c) and yields of CH4 and CO2 (d) for a variety of
initial sodium mole fractions (R) at 298.15 K and 1 atm. Vertical dotted bars labeled 0.03 (mol L1 ) indicate the HAc equivalent of a nominal 10% solids
digester feed.

840

the water in the initial solution. This result is consistent with Le


Chateliers principle. However, all of these changes were very small
in proportion to the changes in R. This observation tells us that
the maximum yield of methane is almost independent of x but it
increases slightly through a transition region, where the process
converts from exothermic to endothermic. From a practical point
of view, the injection of sodium ion increases the pH value at equilibrium resulting in greater solubility of CO2 and increased pCH4 .
These observations do not provide any evidence for increased
activity of hydrogen utilizing methanogens at higher pH values.
In fact Fig. 2(d) provides a thermodynamic argument for a very
small reduction in the activity of hydrogenophilic methanogens.

wileyonlinelibrary.com/jctb

Hence, overall the results presented in Figs 1 and 2(d) suggest that
acetoclastic methanogens will be thermodynamically inhibited by
the increased equilibrium pH arising from the addition of sodium
ion to the initial solution, and that hydrogenophilic methanogens
will be inhibited to a much smaller extent.
HAc degradation in relationship to the concentration of carbonates at constant pH was investigated to compare the model with
empirical observations. Figure 3(b) quantifies the strong effect of
the equilibrium pH on the solubility of CO2 . However this change
produces only a relatively weak response in the methanogenic
degradation of HAc. The natural pH at the equilibrium decomposition of HAc is 4.73 (Fig. 2(c)). To examine the effects of lower

c 2012 Society of Chemical Industry




J Chem Technol Biotechnol 2013; 88: 834844

Effect of sodium ion in acetoclastic methanogenesis

www.soci.org

(a) 1.E-05

1
HAc
AcResidual HAc + Ac-

1.E-06

0.9
0.8
Methane Contents

Concentration (mol/l)

1.E-07
1.E-08
1.E-09
1.E-10
1.E-11

0.7
0.6
0.5
0.4

1.E-12

0.2

1.E-13

0.1
0

10

12

14

pH

Methane contents at equilibrium

0.3

Bashir BHand Matin A, 2005


Ngian MF, Pearcea GR, Ngian KF and Lin SH
Ngian MF, Pearcea GR, Ngian KF and Lin SH

0
0

(b) 1.E+01

Carbonates (mol/l)

1.E-01

1.E-03

1.E-04
2

0.3

0.4

0.5

0.6

0.7

Figure 4. Equilibrium model and experiment observation of methane


contents at 25 C and 1 atm for constant initial HAc (1.5 mole kg1 ) and
constant pH 7. The closed black circles show empirical data for whey
wastes, the open white circles show the results for sheep rumen (A) and
triangles show the empirical data for in sheep rumen with 4.54 g kg1
potassium ion (B).

1.E-02

0.2

R (initial sodium mole fraction of solutes)

Total carbonates
CO2 (aq.)+H2CO3
HCO3CO3-2
Na2CO3
NaCO3NaHCO3

1.E+00

0.1

10

12

14

pH

Figure 3. Residual HAc and CO2 of methanogenesis at 25 C and 1 atm


(constant initial HAc (1 mole) and constant pH).

J Chem Technol Biotechnol 2013; 88: 834844

Ionic strength and non-ideality


The ideal equilibrium model assumed no interactions among solvent and/or solute components. This assumption holds only in
the infinitely dilute system at constant P and T. The model was,
however, calculated for dilute initial conditions to highly concentrated initial conditions in order to explore the thermodynamic
consequences arising from unexpectedly high organic loading
rates, OLR (from under 1 kg m3 day1 to over 7 kg m3 day1 ) for
example, which commonly lead to the failure of biodegradation.
Figure 5(a) shows the variation in ionic strength of the residual
liquid phase against initial HAc mole fraction with the initial solute
sodium mole fraction (R) as the parameter. In pure methanogenesis (R=0) the ionic strength rapidly increases and becomes a constant 1 104 mol L1 , due to the solubility of carbon dioxide.12,13
As the sodium acetate fraction (R) is increased, the ionic strength
of the equilibrium solution increases approximately linearly up to
a maximum value of approximately 0.9 mol L1 . This feature is
due to the solubility of sodium carbonate and bicarbonate arising
from the high pH value. From these results significant deviation
from the ideal model is expected in the range of x > 0.001.
The DebyeHuckelPrausnitz
model was introduced to de
scribe higher concentration systems. The model superimposes
theoretical ionion interactions in the solution phase and empirical moleculemolecule interactions in the vapour phase on the
previous ideal equilibrium model. For the DebyeHuckel
ionion

interactions it is assumed that all ions are equi-distance and


the aqueous moleculemolecule and ionmolecule interactions

c 2012 Society of Chemical Industry




wileyonlinelibrary.com/jctb

841

equilibrium pH values protons were added to the initial solution


whereas to examine higher equilibrium pH values part of the HAc
in the initial solution was substituted with NaAc. Figure 3(a) shows
that, in the ranges of pH less than 4.73 and pH greater than 9.4,
as the pH value is increased, HAc degradation increases, while,
for intermediate pH values (4.73 < pH < 9.4), it decreases. In
the low and high pH ranges and based on the artificial addition of protons and ionization of CO2 , the relationship between
pH value and residual HAc can be explained by the decreasing
concentration of protons with higher pH values. The intermediate
regime can be explained by the rapidly rising solubility of CO2 associated with the release of sodium ions from the added NaAc and
the related changes in the concentrations of HCO3 and NaHCO3 .
Through these mechanisms the pH value, or more precisely the
concentration of protons, controls the behaviour of CO2 gas and
hence HAc degradation in methanogenesis. Overall the residual
concentration of HAc and Ac remain low for the whole range
of initial sodium concentrations. The empirical observations2 8
suggest significant inhibition in the presence of sodium ions. Thus
it can be concluded that the empirically observed inhibition was
entirely kinetic in origin.
Figure 4 compares empirical observations of gas composition
with the thermodynamic model. It is clear that there is little
agreement. The only thermodynamic explanation that can be put
on these differences is that as the sodium ion content increases

CO2 is eliminated from carbohydrates in the feed solution to


yield fatty acids with chain lengths greater than two carbons.
This is also consistent with the empirically observed inhibition
of methanogenesis. The thermodynamic model ignores kinetic
factors. Hence the difference between the model results and the
empirical data provides further evidence that increased sodium
ion concentration contributes to kinetic limitation or inhibition
but has only limited effects on the thermodynamics.

www.soci.org

(a)

ST Oh, AD Martin

(b)

0.03

0.03

1.E+01
1.E+00

7
pH

Ionic strength

1.E-01
1.E-02

6
1.E-03

1.E-04
1.E-05
1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E-00

4
1.0E-05

1.0E-04

1.0E-03

Initial x (HAc)
R -0(mol./mol.)
R -0.2(mol./mol.)

R -0.001(mol./mol.)
R -0.5(mol./mol.)

1.0E-02

1.0E-01

1.0E-00

Initial x (HAc)
R -0.01(mol./mol.)
R -0.8(mol./mol.)

R -1(mol./mol.)

I, R=1

I, R=0.8

I, R=0.2

I, R=0.01

DH, R=1

DH, R=0.8

DH, R=0.2

DH, R=0.01

P, R=1

P, R=0.8

P, R=0.2

P, R=0.01

Base Case

I, solid
boundary

DH, solid
boundary

P, solid
boundary

Figure 5. Ionic strength (a) and non-ideality of pH value (b) in variety of initial sodium mole fractions (R) of solutes at thermodynamic methanogenesis
equilibrium (298.15 K and 1 atm). Vertical dotted bars labeled 0.03 (mol L1 ) indicate the HAc equivalent of a nominal 10% solids digester feed. I stands
for the ideal model, DH for DebyeHuckel and P for PitzerPrausnitz model.

842

are zero. The DebyeHuckel


assumption holds in dilute systems

(ionic strength 0.00514 at constant 1 atm and 298.15 K). For the
empirical moleculemolecule interactions in the vapour phase,
Prausnitzs method15 was employed to describe the behaviour
of the highly polar gas molecules. The maximum ionic strengths
predicted by the ideal model (i.e. 0.9 mol L1 ) lie outside the
valid range for the DebyeHuckelPrausnitz
model. Despite the

conditions obviously lying above the limit of validity of the DebyeHuckelPrausnitz


model, it was studied as it is based upon

a theoretical description of ion interactions which compares with


the empirical descriptions used in the PitzerPrausnitz model.
Thus, the PitzerPrausnitz model was used for high values of
ionic strength (0.005 I 2).14 The empirical Pitzer interactions
were superimposed on the ideal equilibrium model to describe
the ionion, moleculemolecule and ionmolecule interactions.
Figure 5(b) shows the pH as modelled by Ideal, DebyeHuckel
and

Pitzer models. The trends shown by the DebyeHuckel


and Pitzer

models generally correspond to the previous ideal condition,


however, these models predict significantly different conditions
under which a solid phase is present. The results show that the
domain for which a solid phase is predicted increased from the
Ideal model with the DebyeHuckel
predicting a larger domain or

lower solubility of the solid phase than the Pitzer model. It tells us
that the activity coefficient predicted from the DebyeHuckel
is

lower than the Pitzers empirical coefficient arising from the higher
molecularmolecular interaction. Thus, Fig. 5(b) shows that as either R or x is increased, molecularmolecular interaction increases
and the deviation of pH value between ideal and nonideal model
increases. The solid phase was increased while the other chemical
species in solution do not deviate much from Ideal behaviour.
However, despite the deviations in the predicted equilibrium pH
of non-ideal models, the remaining results of non-ideal model are
very similar to the results of the ideal equilibrium model. Under
these circumstances the interactions become insignificant for the

wileyonlinelibrary.com/jctb

anaerobic digestion process and the ideal model can be used over
the whole range of ionic strength.
Enthalpy and entropy analysis
The overall change in enthalpy represents total production or
consumption of energy in an anaerobic digester. When the value
is positive, the process is endothermic, requiring external heat
energy to progress. Figure 6 shows that the enthalpy change
(A) of HAc conversion to CO2 (aq.) and CH4 (aq.) against initial x
of HAc was an exothermic process and almost independent of the
injection of sodium ion, but total enthalpy change (H of HAc
conversion to CO2 (g.) and CH4 (g.)) becomes endothermic due to
the H of vaporisation for CO2 and CH4 . Figure 6 also shows that
as R was increased the process becomes less endothermic and
at high values, R > 0.5, becomes exclusively exothermic. This is
entirely due to the increased solubility of CO2 .
The isothermal enthalpy change at equilibrium represents the
change of entropy. Since the change in Gibbs free energy is zero
at equilibrium, the change in entropy is equal to the change
in enthalpy per absolute temperature (S = H/T). Thus the
change in entropy has the same tendency as the change in
enthalpy. When the value is positive, it means that the overall
process of digesters is spontaneous from initial values to the static
equilibrium condition resulting in an increase in the degree of
disorder in the system. When the value is negative, it means that
the overall process is non-spontaneous. The results in Fig. 6 show
two cases: (H > 0 and S > 0) and (H < 0 and S < 0).
First, the case (H > 0 and S > 0) tells us that the system
requires external heat energy to proceed spontaneously and the
energy is consumed for the vaporisation of end-products (CH4
and CO2 ). Second, the case (H < 0 and S < 0) tells us that the
system produces heat energy, but the system does not progress
forward. In these two cases, the thermodynamic model tells us
that the addition of sodium ion to the initial solute may render

c 2012 Society of Chemical Industry




J Chem Technol Biotechnol 2013; 88: 834844

Effect of sodium ion in acetoclastic methanogenesis

www.soci.org

0.03
1.E+04
1.E+02
Endotherm
1.E+00
1.E-02
-1.E-03
-1.E-01

1.0E-05

1.0E-04

1.0E-03
1.0E-02
1.0E-01
Initial x (HAc) 0.03

R -0(mol./mol.)
R -0.001(mol./mol.)
R -0.2(mol./mol.)
R -0.5(mol./mol.)
R -1(mol./mol.)

Exotherm

-1.E+01
-1.E+03
1.0E+00

R -0.01(mol./mol.)
R -0.8(mol./mol.)

Figure 6. Enthalpy production in a variety of initial sodium fractions (R) at 298.15 K and 1 atm. Vertical dotted bars labeled 0.03 (mol L1 ) indicate the
HAc equivalent of a nominal 10% solids digester feed.

the thermodynamically feasible process non- spontaneous and


that the addition of enthalpy may restore spontaneity. These
actions are equivalent to moving from left to right and upwards
respectively in Fig. 6. Practically this result suggests that where
sodium ion inhibition is evident the application of thermophilic
anaerobic digestion may overcome the lack of spontaneity. In
the high-range of R (>0.5), however, the additional sodium ions
introduce inhibition by loss of spontaneity that may be difficult to
overcome. This result is independent of the pH value in a given
system but is related to the positive entropy change of the process.

CONCLUSION

J Chem Technol Biotechnol 2013; 88: 834844

REFERENCES
1 Callander IJ and Barford JP, Recent advance in anaerobic digestion
technology. Process Biochem 18:2437 (1983).

c 2012 Society of Chemical Industry




wileyonlinelibrary.com/jctb

843

In this paper the theoretical thermodynamic approach was used


to describe the inhibitory effects of sodium ion in acetoclastic and
hydrogen utilising methanogenesis.
The thermodynamic equilibrium was simulated in an isothermal
and isobaric (298.15 K and 1 atm) batch digester. The initial substrates considered were a mixture of sodium acetate (CH3 COONa)
with acetic acid (CH3 COOH). The anaerobic digestion was assumed
to contain a fully acclimatized methanogenic consortium. In the digestion process, acetoclastic hydrogen production was found not
to proceed spontaneously (H  0 and G  0). However, the
overall acetoclastic digestion process, including methanogenesis,
can proceed spontaneously (G < 0). For acetoclastic hydrogen
production to proceed, it must be coupled with hydrogen-utilising
methanogenesis. The non-spontaneous process requires the free
energy released from hydrogen-utilising methanogenesis to oxidise HAc to CO2 and release hydrogen (protons and electrons).
The spontaneous, hydrogen-utilising methanogenesis uses the
hydrogen (protons and electrons) for methane production and
releases the free energy.
Overall the model results show that the addition of sodium ion
slightly reduces the predicted conversion of acetate to products
but significantly improves the methane partial pressure, pCH4 , in

the gas phase. The model results also indicate two contrasting
cases for methanogenesis: (H > 0 and S > 0) and (H < 0
and S < 0). In the former case, the model tells us that, for
very low initial concentrations of sodium ion, methanogenesis
is spontaneous depending on the relationship of solubility of
CO2 . However for the range of R (>0.5) substrate degradation is
predicted to be non-spontaneous despite the Gibbs free energy
change being negative. Thus the empirically observed inhibition
of gas formation brought about by elevated concentrations
of sodium ion may be partially explained by the loss of
thermodynamic spontaneity or the absence of an entropy driving
force. Under certain circumstances spontaneity may be restored
by providing external heat to vaporise the gases, in particular
dissolved carbon dioxide. Practically this may be achieved
in thermophilic digestion. The results of the thermodynamic
modeling strongly indicate that the inhibitory phenomena arising
from the presence of sodium ion in methanogenesis are directly
related to the entropic and enthalpic changes in the process rather
than the deviations from the neutral pH range (6.5<pH<8.0).
From the thermodynamic perspective, the model advises that
when the supply of thermal energy is sufficient, sodium inhibition
under anaerobic conditions can be virtually completely overcome.
The thermodynamic model also provides a design tool with which
the stability of methanogenesis in response to feedstock mixtures
containing sodium ion and a carbon source such as HAc can be
investigated. An operator using a thermodynamic model of his
process can predict, for example, the equilibrium pH and maximum
gas production and quality of individual gas components of the
digester and where this lies outside the prescribed range he may
choose to combine available feed materials in ratios to achieve
the desired equilibrium pH.

www.soci.org
2 Rinzema A, van Lier J and Lettinga G, Sodium inhibition of acetoclastic
methanogens in granular sludge from a UASB reactor. Enzyme
Microb Technol 10:2432 (1988).
3 Grasso D, Strevett K and Pesari H, Impact of sodium and potassium on
environmental systems. J Environ Syst 22:297323 (1993).
4 McCarty PL, Anaerobic waste treatment fundamentals. Part three.
Public Works 95:9194 (1964).
5 Kugelman IJ and Chin KK, Toxicity, synergism, and antagonism in
anaerobic waste treatment processes. Adv Chem Series 105:5590
(1971).
6 Patel GB and Roth LA LAR, Effect of sodium chloride on growth
and methane production of methanogens. Canadian Journal of
Microbiology 23:893897 (1977).
7 Ngian MF, Pearcea GR, Ngian KF and Lin SH, Anaerobic digestion of
sodium hydroxide pretreated pig faeces using sheep rumen liquor
or anaerobic digester mixed liquor as inoculum. Agric Environ
4:139154 (1978).
8 Bashir BH and Matin A, Combined effect of potassium and magnesium
on sodium toxicity in anaerobic treatment processes. ElectrJEnviron,
Agric Food Chem 4:827834 (2005).
9 Jackson BE and Mclnerney MJ, Anaerobic microbial metabolism can
proceed close to thermodynamic limits. Nature 415:454456
(2002).

ST Oh, AD Martin

10 Marcus RA and Norman S, Electron transfers in chemistry and biology.


Biochim Biophys Acta 811:265322 (1985).
11 Press WH, Teukolsky SA, Vetterling WT and Flannery BP, Numerical
Recipes in C: The Art of Scientific Computing. Cambridge University
Press (1998).
12 Oh ST and Martin AD, Thermodynamic equilibrium model in anaerobic
digestion process. Biochem Eng J 34:256266 (2007).
13 Oh ST and Martin AD, Long chain fatty acids degradation in
anaerobic digester: thermodynamic equilibrium consideration.
Process Biochem 45:335345 (2010).
14 Pourbaix M, Atlas of electrochemical equilibria in aqueous solutions,
NACE Cebelcon Publishing. New York (1974).
15 Stumm W and Morgan JJ, Aquatic chemistry, in Dissolved Carbon
Dioxide, John Wiley & Sons, Inc, New York, 150152 (1996).
16 Nakamura R, Breedveld GJF and Prausnitz JM, Thermodynamic
Properties of gas mixtures containing common polar and nonpolar
components. Ind Eng, Chem Process Design Division 15:557564
(1976).

844
wileyonlinelibrary.com/jctb

c 2012 Society of Chemical Industry




J Chem Technol Biotechnol 2013; 88: 834844

S-ar putea să vă placă și