You are on page 1of 7

Applied Energy 111 (2013) 351–357

Contents lists available at SciVerse ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Kinetics of faecal biomass hydrothermal carbonisation for hydrochar


production
E. Danso-Boateng a,⇑, R.G. Holdich a, G. Shama a, A.D. Wheatley b, M. Sohail b, S.J. Martin c
a
Department of Chemical Engineering, Loughborough University, Loughborough, LE11 3TU, UK
b
School of Civil and Building Engineering, Loughborough University, Loughborough, LE11 3TU, UK
c
Department of Materials, Loughborough University, Loughborough LE11 3TU, UK

h i g h l i g h t s

 Solids decomposition and hydrochar production modelled by first order kinetics.


 Arrhenius rate constant characterisation of sewage sludge carbonisation.
 The activation energy of synthetic faeces was higher than that of sewage sludge.
 Reaction temperature had more effect on solids decomposition than reaction time.
 Feedstock moisture content affected hydrochar yield and extent of carbonisation.

a r t i c l e i n f o a b s t r a c t

Article history: Decomposition kinetics of primary sewage sludge (PSS) and synthetic faeces (SF), of various moisture
Received 25 January 2013 contents, were investigated over different reaction times and temperatures using a hydrothermal batch
Received in revised form 29 April 2013 reactor. Solid decomposition of PSS and SF was first-order with activation energies of 70 and 78 kJ/mol,
Accepted 30 April 2013
and pre-exponential factors of 4.0  106 and 1.5  107 min1, respectively. Solid decomposition was sig-
Available online 3 June 2013
nificantly affected by reaction temperature more so than reaction time. Higher temperature resulted in
higher solids conversion to hydrochar. Equilibrium solid hydrochar yields (relative to the original dry
Keywords:
mass used) were 74%, 66%, 61% and 60% for PSS at 140, 170, 190 and 200 °C respectively, and 85%,
Hydrothermal carbonisation
Decomposition kinetics
49%, 48% and 47% for SF at 140, 160, 180 and 200 °C respectively. Energy contents of the hydrochars from
Sewage sludge PSS carbonised at 140–200 °C for 4 h ranged from 21.5 to 23.1 MJ/kg, and increased following carbonisa-
Synthetic faeces tion. Moisture content was found to affect the Hydrothermal Carbonisation (HTC) process; feedstocks
Activation energy with higher initial moisture content resulted in lower hydrochar yield and the extent of carbonisation
was more evident in feedstock with lower moisture content. The results of this study provide information
useful for the design and optimisation of HTC systems for waste treatment.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction Foundation, under its ‘Reinvent the Toilet Challenge’, to develop


a self-contained toilet system for developed and developing coun-
Poor sanitation is a global challenge especially in developing tries that is able to process human sewage and which is not reliant
countries with high rates of population growth. It is estimated that on external sources of energy and connection to a sewerage sys-
between about 0.6–1 billion tons of human faeces are generated tem. Factors driving this include increasing population growth,
each year [1,2]. In developing countries, over 90% of sewage is dis- regulatory constraints on disposal options as well as maintaining
charged untreated [3,4]. Faecal contamination of water sources a low reduction of carbon footprint and breaking the cycle of infec-
causes almost 4 billion cases of diarrhoea globally each year, tion by microbial pathogens.
resulting in the deaths of some 2.2 million children aged under five Hydrothermal carbonisation (HTC), a thermochemical process
[5]. The motivation for the work reported here was to investigate a represents one possible means of processing human wastes within
process for rendering human faeces safe whilst at the same time the constraints mentioned above. HTC has been shown to be effec-
recovering energy from it. This work is related to that currently tive for converting biomass with higher moisture content into
being undertaken and funded by the Bill and Melinda Gates carbonaceous solids (biological char) commonly referred to as
‘hydrochar’, aqueous and gaseous products in a pressured and
⇑ Corresponding author. Tel.: +44 (0)1509 222528; fax: +44 (0)1509 223923. oxygen-free reactor at moderate temperature [6]. Biochemical
E-mail address: e.danso-boateng@lboro.ac.uk (E. Danso-Boateng).
conversion methods, namely anaerobic and aerobic digestions,

0306-2619/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.apenergy.2013.04.090
352 E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357

require long residence times, and other thermochemical conver- Faecal sludge ! Hydrochar þ Liquid þ GasðCO2 ; aromaticsÞ ð1Þ
sion technologies, such as dry (fast) pyrolysis, gasification and
The reaction rate depends on temperature and the amount
incineration, require feedstock drying [6,7]. The HTC process does
(mass) of reacting solids present. The operating temperature will
not require drying and the process temperatures employed result
determine the overall degree of conversion to solid product.
in the destruction of pathogens. HTC is self-contained and gener-
The mass of initial reacting solids dependence on the reaction
ates little or no GHG gas emissions [8]. HTC hydrochar typically
rate can be expressed as a first-order differential rate equation.
has H/C and O/C ratios similar to that of coal [9], and higher caloric
(heating) value comparable to coal so can be used as fuel. dmt
r ¼  ¼ kmt ð2Þ
Despite the clear technological potential of HTC, relatively little dt
information is available on the kinetics of biomass hydrothermal
which separating and integrating produces
carbonisation. A small number of studies have reported dry pyro-
lysis rates using thermogravimetric analysis (TGA) [10–15], at ex- mt
 ln ¼ kt ð3Þ
tremely high decomposition temperatures (up to 900 °C) which is mo
above that desired for HTC, where the operating pressure is nor-
where r is the reaction rate, mt is the mass of solids (hydrochar) at
mally fixed by the need to prevent vaporisation of the liquid (i.e.
reaction time (t), mo is the initial mass of solids in the reactor at
boiling). Some of these studies have postulated the rate of reaction
t = 0, (on a dry weight basis), k is the reaction rate constant (min1),
of cellulose pyrolysis to be first-order [10,11,16]. However, there
and t is the reaction time (min).
appears to be no general agreement relating to the pyrolysis kinet-
In terms of reactant conversion,
ics of biomass decomposition from these TGA analyses.
Studies into HTC of animal and municipal wastewater solids mt  m1
X¼ ð4Þ
have previously been reported, but with objectives different to mo  m1
those of this study. Berge et al. [17] studied the carbonisation of
where X is the conversion of solids in the faecal waste to hydrochar,
municipal wastewater solids (anaerobic digested sludge, food
m1 is the equilibrium mass of solids (or hydrochar) over very long
waste, waste paper, and mixed municipal solid waste), and the re-
times or time t = 1.
sults obtained showed that carbonisation improved heat values
(HHVs) of the organic solids which correlated well with the carbon dX
¼ kX ð5Þ
content. Lu et al. [18] studied the effect of hydrothermal treatment dt
on the characteristics and combustion performance of municipal
which has solutions of the form
solid wastes (MSWs). Their results showed that such wastes could
be converted to uniform powders with low moisture content, high XðtÞ ¼ expðktÞ ð6Þ
bulk density, and improved heat values. Cao et al. [19] investigated
The operating temperature will determine the overall degree of
changes to the chemical composition of swine manure under dif-
conversion to solid product; that is at equilibrium.
ferent carbonisation conditions, and found that the hydrochars
From the experiments, the mass mt of solids present at reaction
produced had an increased content of aromatic carbons which
time t, is obtained after oven drying, and solid yield was calculated
accompanying a decrease in carbohydrates and proteins/peptides
by the following equation, as
compared. Prawisudha et al. [20] studied the production of coal
alternative fuel from MSW using hydrothermal treatment, and mt
Y¼  100% ð7Þ
the results showed that the heating value of the MSW after hydro- m0
thermal treatment was slightly higher than the raw MSW, and
A first-order model for the conversion of solids (hydrochar) was
both were almost equal to that of low-grade sub-bituminous coal.
estimated by fitting Eq. (6) to the data obtained.
Wang et al. [21] analysed liquid products from hydrothermal treat-
The temperature dependence of the reaction rate is typically
ment of cellulose, and found ketones, five ring oxygen-containing
correlated as a reaction coefficient by the Arrhenius equation.
compounds, small organic acids and phenol derivatives as the prin-
 
cipal components. E
K ¼ A exp  ð8Þ
However, HTC of human faecal waste has not been as exten- RT
sively researched and reported as has lignocellulosic, wastewater
solids and other types of biomass. This paper is the first to report where, A is the pre-exponential or frequency factor (min1), E is the
comparative kinetics of hydrochar production and decomposition activation energy (J mol1), R is the gas constant (8.314 J mol1 -
of primary sewage sludge as well as synthetic faeces. The objec- K1), and T is the temperature (K).
tives of this study, therefore, were to: (i) gain insight into the kinet-
ics of hydrochar production of primary sewage sludge and faecal 2. Methods
simulant by hydrothermal carbonisation, taking into consideration
the influence of reaction temperature and time on the hydrochar 2.1. Primary sewage sludge and faecal simulant
production; (ii) develop a kinetic model based on the results of
the experimental study; and (iii) examine the effect of moisture Primary sewage sludge (PSS), was collected from Wanlip Sew-
content on hydrochar formation and extent of carbonisation. age Treatment Works (Leicestershire, UK) in an enclosed container.
PSS primarily comprises faecal matter removed by settlement. The
1.1. Decomposition kinetics PSS typically contained 4% (wt.) solids (i.e. 96% moisture) as re-
ceived. Table 1 contains the physical and chemical characteristics
Investigating detailed reaction mechanisms for thermochemical associated with the PSS feedstock. The synthetic faeces (SF) sam-
decomposition and hydrochar formation of complex biomass such ples were formulated using the formulation proposed by Wignara-
as human sewage sludge is beyond the scope of this study. Instead jah et al. [22]. This comprised cellulose (37.5%), yeast (37.5%),
a first-order reaction rate and Arrhenius equations were used to peanut oil (20%), KCl (4%), Ca(H2PO4)2 (1%) (all purchased from
model the solids decomposition kinetics for the primary sewage Sigma–Aldrich, UK), and tap water – the amount of which de-
sludge and synthetic faeces samples from hydrothermal pended on the chosen moisture content. The materials were uni-
carbonisation. formly mixed in accurate proportions with water to form a paste
E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357 353

Table 1 with other solid contents of 15 and 5% (wt.) (i.e. 85 and 95% water
Characteristics of initial primary sewage sludge feedstock and hydrochar produced at content respectively) at various reaction times and 200 °C
different temperatures at a retention time of 4 h.
(expected best carbonisation temperature). Photographs of the car-
Parameters Unit Feedstock Hydrochar bonised solids before oven drying were taken with a Canon digital
Dried 140 °C 170 °C 190 °C 200 °C camera (DS126061, Canon Inc., Japan) for visual comparison with
Proximate analyses
standard RGB colour strips. Greyscale values were calculated for
Moisture % 8.73 4.95 4.25 3.89 3.53 the RGB values of each hydrochar. The energy values (HHVs) of
Ash % db 15.97 22.61 24.04 24.32 25.31 the hydrochars were obtained using a bomb calorimeter (CAL 2K,
Volatile matter % db 70.34 66.43 64.92 64.52 63.67 Digital Data Systems, South Africa), to further analyse the effect
Fixed Ca % db 5.33 6.01 6.76 7.27 7.49
of feedstock moisture content on carbonisation extent. All experi-
Calorific value MJ/ 18.01 21.45 22.02 22.65 23.13
(HHV) kg ments were run in triplicate.

Ultimate analyses 2.4. Proximate and ultimate analysis of PSS hydrochar


Carbon (C) % db 46.93 44.43 45.77 45.58 46.17
Hydrogen (H) % db 6.11 5.81 5.99 5.70 5.81
Hydrochars from PSS at 140, 170, 190 and 200 °C were carbon-
Oxygenb (O) % db 50.90 51.96 50.23 50.35 49.39
Nitrogen (N) % db 4.17 2.43 1.96 1.97 1.88 ised for 4 h, and the resulting dried solids were analysed for
Energy yieldc % – 72.75 66.75 62.40 62.15 moisture, ash, volatile matter and fixed carbon using a thermo-
gravimetric analyser (TA Instruments Q5000IR, UK), according to
Liquid filtrate ASTM D7582-10. Carbon (C), hydrogen (H) and nitrogen contents
TOC g/L 2.17 5.06 4.64 5.13 4.83 of dried feedstock and hydrochars were analysed using a CHN Ana-
COD g/L 36.60 47.60 47.83 50.00 48.87
lyser (CE-440 Elemental Analyzer, Exeter Analytical Inc., UK),
BOD g/L 7.01 8.01 8.01 9.01 9.01
according to ASTM D5373-08. Energy content of the hydrochars
a
100 – (moisture + ash + volatile matter). were analysed using a bomb calorimeter. All tests were conducted
b
Calculated as different between 100 and total C/H/N. in triplicate. Tests were only performed on the PSS as it more clo-
c
(Mchar  HHVchar)/(Mfeedstock  HHVfeedstock). db = dry basis.
sely represented the actual faecal matter, whilst the 4 h reaction
time was employed as preliminary tests had revealed this to be
or suspension. These were prepared immediately before conduct-
the best holding time for the carbonisation.
ing the carbonisation tests.

2.2. Kinetic analysis of hydrochar production 2.5. Liquid product analyses

Primary sewage sludge (PSS), as received, was subjected to HTC The total organic carbon (TOC) of the clear liquid filtrate (prod-
using a 250 mL stainless steel batch reactor (BS1506-845B, UK) im- uct) was analysed using a TOC Analyser (DC-190, Rosemount Dohr-
mersed in an oil bath (B7 Phoenix II, Thermo Scientific, UK) con- mann, USA), according to Standard Methods 5310 B – High
taining Shell Thermia oil B. The PSS was periodically stirred to Temperature Combustion Method. This was done to obtain a car-
prevent settling. To ensure that the reaction temperature was ob- bon balance of the products. COD was measured using COD Ana-
tained quickly, the oil bath was heated to the required temperature lyser (Palintest 8000, Palintest Ltd., UK), according to Standard
before the immersion of the reactor with its contents. HTC were Methods 5229 D – Close Reflux Colorimetric Method. BOD was
performed at reaction temperatures of 140, 170, 190, and 200 °C analysed in line with Standard Methods 5210 B – 5-day BOD Test.
separately for 0, 30, 50 min, 1, 1.5, 2 and 4 h. The temperature
range used here was selected so as to minimise energy require- 3. Results and discussion
ment for the HTC system in order that the process would remain
economical for developing countries. After the HTC experiment, 3.1. Solid decomposition and hydrochar formation kinetics
the reactor was cooled to 25 °C and the gaseous products vented.
The hydrochar (solids) were separated (dewatered) from the liquid Figs. 1 and 2 show conversion plots from the experimental data
phase by vacuum filtration (Whatman filter paper, 11 lm) to ob- using the rate laws in Eqs. (4) and (5). The figures depict the vari-
tain a clear liquid product. The dewatered hydrochar was dried ation in the decomposition of solids with time by HTC of synthetic
in an oven at 55 °C for 24 h to remove any residual moisture, and faeces (SF) and primary sewage sludge (PSS). The first-order reac-
the residual hydrochar after carbonisation was determined using tion models for the hydrochar formation were established from
Eq. (7), after subtracting any residual moisture from the residual Eqs. (4)-(6).
solids after drying. The mild temperature was used to avoid further Not all the experimental data gave a good fit to the first-order
loss of volatile compounds; which may occur during TGA analysis model, but overall the agreement seems reasonable and the
[23]. The initial times taken for the reactor content to reach the ac- assumption of first order kinetics is justified. Experiments were
tual reaction temperatures were noted; and termed the ‘‘heat-up conducted in triplicate in order to obtain greater precision and
time’’. These times ranged from 10 to 20 min. the data shown in these figures represents the means of the values
Hydrochar production kinetics of the synthetic faeces samples obtained.
were carried out as above with 25% solid loading (75% moisture Significant solids conversions were obtained in the initial
content) at 140, 160, 180 and 200 °C for 0, 30, 50 min, 1, 1.5, 2, 30 min of the reaction for all feedstocks and reaction temperatures,
4 h; and further 5 and 6 h at 140 and 160 °C, as well as 4.5 h at and the rate slowed down exponentially as the reaction time in-
180 °C. creased (see Figs. 1 and 2). This suggests that significant decompo-
sition of the biomacromolecules started within the initial 30 min of
2.3. Effect of feedstock moisture on hydrochar formation and the reaction. There was no notable conversion of solids during the
carbonisation extent heating-up time. Reaction temperature had a more pronounced
effect on solids degradation and hydrochar yield production for
In order to analyse the influence of feedstock moisture content both synthetic faeces (SF) and primary sewage sludge (PSS) than
on hydrochar formation, carbonisations were performed using SF did the reaction time. For PSS at all temperatures and SF at
354 E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357

Fig. 1. Variation of conversion of solids with time for hydrothermal carbonisation Fig. 2. Variation of conversion of solids with time for hydrothermal carbonisation
of primary sewage sludge and synthetic faeces: (a) HTC at 140 °C; (b) HTC at 160 °C of primary sewage sludge and synthetic faeces: (a) HTC at 180 °C for SF, and 190 °C
for SF, and 170 °C PSS. for PSS; (b) HTC at 200 °C.

200 °C, increasing the reaction time beyond 1 h did not signifi-
cantly improve solids conversion to hydrochar. Equilibrium rela-
tive masses of 74%, 66%, 61% and 60% wt were obtained for 140,
170, 190, and 200 °C respectively for PSS (shown in Fig. 3).
Fig. 4 shows the Arrhenius plot for the reaction rate constants
from Figs. 1 and 2. The kinetic rate constants obtained from Eq.
(5) with the experimental data gave a good fit to the Arrhenius plot
(Fig. 4). The kinetic parameters (i.e. activation energy and pre-
exponential factor) were estimated from the slope and intercept.
A higher value of activation energy (77.8 kJ/mol, with pre-expo-
nential factor 1.5  107 min1) was obtained for synthetic faeces
(SF) compared with 70.4 kJ/mol, with pre-exponential factor of
4.0  106 min1 for primary sewage sludge (PSS). This denotes that
a higher temperature is required to initiate solids decomposition of
SF to produce hydrochar than that required for PSS. This is notice-
able in the higher solids conversion rate for PSS compared to SF as
Fig. 3. Relative mass remaining in the solid phase at equilibrium after hydrother-
shown in Figs. 1 and 2a. The activation energy for SF was lower mal carbonisation for synthetic faeces and primary sewage sludge.
than the reported activation energy for pyrolysis decomposition
of semisynthetic raw bacterial cellulose (92.7–471.0 kJ/mol) be-
tween 30 and 1000 °C [10], which has a comparable formulation 3.2. Effect of moisture content on hydrochar production
to that of the synthetic faeces used in this study. The activation en-
ergy of PSS was higher than the reported activation energy for Fig. 5 shows a comparison of solid conversion to hydrochar for
pyrolysis of refinery sludge (32.5–45.1 kJ/mol) between 120 and the HTC of synthetic faeces (SF) at different solid concentrations
565 °C [13], but lower than those of olive cake (180.2 kJ/mol) be- (i.e. moisture contents), as well as the extent of carbonisation. All
tween 107 and 947 °C [12], and swine solids and anaerobic lagoon data points are the means of three measurements. Moisture con-
sludge (92–161 kJ/mol) between 207 and 473 °C [14]. tent had a significant effect on hydrochar production and extent
The solid decomposition and hydrochar formation for both syn- of carbonisation. There was significant conversion of solids to
thetic faeces and primary sewage sludge follows first-order (decay) hydrochar in the first 30 min for all feedstocks, as shown in
reaction kinetics given in Eq. (6) and the reaction rates can be cor- Fig. 5a. Within the first 30 min, the solid (hydrochar) yield dropped
related by an Arrhenius plot. Hence, the data presented in this to 60%, 58% and 65% for SF feedstock with 75%, 85% and 95%
study would be suitable for application in a number of different moisture contents respectively (Fig. 5b). With further increase in
calculations to process operated over a range of temperatures. reaction time beyond 50 min, the reaction rates became less
E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357 355

mass of carbonised hydrochar, whilst feedstock with higher initial


moisture content (i.e. lesser solids contents) produced lower
hydrochar yields (Fig. 5b) and the hydrochars from feedstocks of
higher solids content was more carbonised (indicated by the lower
greyscale values, and higher energy values), Fig. 5c and d.

3.3. Hydrochar characteristics and energy contents

The characteristics of the PSS hydrochar carbonised at 140, 170,


190 and 200 °C for 4 h holding time and that of the dried PSS feed-
stock are shown in Table 1. Fixed carbon resulting from carbonisa-
tion range from 6.0% to 7.5%, compared with that of the initial PSS
feedstock (5.3%), and increased with increasing temperature. These
values are significantly lower than those reported for swine man-
Fig. 4. Arrhenius plot for determination of kinetic parameters for hydrothermal ure solids (11.8%) [14], municipal wastewater solids (6.4–29.7%)
carbonisation of synthetic faeces and primary sewage sludge. [17,18]. As a result of carbonisation, volatile matter in the hydro-
char decreased significantly compared with that in the dried PSS
feedstock. A decrease in volatiles was also observed as the reaction
dramatic. The reaction was faster at lower solid contents (higher temperature was increased from 140 to 200 °C (Table 1).
moisture), especially after the 30 min of the reaction, producing HTC brings about changes to the elemental compositions of the
lesser hydrochar yields: 62–53%, 58–50%, and 56–46% from 50 to raw feedstock that is dependent on temperature (Table 1). An in-
240 min, for feedstock with 75%, 85% and 95% moisture contents crease in carbon content occurs coupled with a decrease in oxygen
respectively. content is seen following carbonisation, except for carbonisation at
As indicated in Fig. 5c, hydrochars obtained from feedstocks of 140 °C for 4 h. Increasing temperature between 170 and 200 °C
75% moisture content were darker in colour (i.e. were more leads to an increase in carbon content as oxygen content decreases,
carbonised) – as revealed by their lower greyscale values, followed with significant change being obtained at 200 °C, suggesting car-
by that of 85% moisture content. Fig. 5d gives the energy values bonisation was most effective at the higher temperature (200 °C).
(HHVs) of the hydrochars, which are higher for hydrochars Details on carbon balance are presented in Section 3.3.1.
produced from 75% moisture feedstock (25–27 MJ/kg), followed Energy contents of the hydrochars produced from PSS carbon-
by that of 85% moisture (21–26 MJ/kg) and finally 95% moisture ised at 140–200 °C for 4 h ranged from 21.5 to 23.1 MJ/kg, and in-
(20–24 MJ/kg), these all increased as the reaction time was creased following carbonisation (Table 1). The HHVs improved as
increased. The results showed that, feedstock with lower initial temperature was increased, and correlated well with the carbon
moisture content (i.e. higher solids contents) produced a greater content of the hydrochars (Fig. 6). These results are similar to a

Fig. 5. Effect of moisture content on solid decomposition during hydrothermal carbonisation of synthetic faeces at 200 °C and at different reaction times on: (a) conversion;
(b) hydrochar yield; (c) and (d) extent of carbonisation.
356 E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357

Fig. 6. Characteristics of PSS hydrochar at different temperatures with 4 h retention


Fig. 7. Characteristics of PSS hydrochar at different temperatures with 4 h retention
time: (a) temperature effect on energy values and carbon contents; (b) relating
time: (a) temperature effect on carbon recovery of the starting carbon (error bars
calorific values with carbon content. a.d. = as determined.
represent standard deviations from triplicate measurements); (b) temperature
effect on energy density and carbon storage factor.

relationship previously reported by Berge et al. [17] for carbonisa-


tion of wastewater solids. Energy contents of the hydrochars hydrochars during HTC ranged from 0.34 to 0.29, and decreased
following carbonisation of wastewater solids and sludge range as the reaction temperature was increased from 140 to 200 °C
from 14.4 to 27.2 MJ/kg [13,17,18,20]. Energy densification of the respectively (Fig. 7b). Carbon storage factor (CSF) is the fraction
hydrochars occurs as a result of decreases in solid mass caused of carbon which remains unoxidised following biological decom-
by dehydration and decarboxylation reactions (evident by the in- position, which is defined by Barlaz [25] as mass of carbon remain-
creased carbon contents and decrease in oxygen contents). Energy ing in the solid after biological decomposition in a landfill per unit
densification factors associated with the hydrochars range from dry mass of feedstock, and provides the means to compare the
1.19 (for 140 °C carbonisation) to 1.28 (for 200 °C carbonisation), mass of carbon remaining (sequestered) within solid material
and increase significantly as reaction temperature increases ensuing biological decomposition in landfills. Previously reported
(Fig. 7b). Energy densification ratio is defined as ratio of HHV of values of CSF for food waste and mixed MSW in landfills are 0.08
hydrochar to that of initial dried feedstock. Energy yield, which is and 0.22 respectively [25], and 0.34, 0.23 and 0.14 for food, mixed
defined as energy densification ratio times mass yield of the hyd- MSW and AD waste hydrochars respectively [17]. Although CSFs in
rochars provide a means of relating the energy remaining within the hydrochars decrease as the reaction temperatures increase, the
the hydrochars to that of the original sewage sludge feedstock, long-term stability of the carbon in the char is not known.
and ranged from 72% for (140 °C carbonisation) to 62% (for Small proportions of carbon are transferred into either the li-
200 °C carbonisation), and decreased significantly with increase quid or gas phases, as shown in Fig. 7a. Fractions of carbon of the
temperature (Table 1). total initial carbon present in the feedstock transferred to the li-
quid phase range from 22% to 24%. These are calculated from mea-
3.3.1. Carbon balance surements of total organic carbon (TOC) of the liquid products. The
The carbon content of the PSS hydrochars produced ranged transfer of the small proportions of carbon into the liquid phase
from 42% to 45% (Fig. 6a). Carbon content of the hydrochar after may be due to products of the Maillard or Browning reactions
carbonisation at 140 °C is slightly lower than that of the initial which results in the dissolution of low molecular weight carbon
feedstock after carbonisation, suggesting the carbonisation of pri- compounds during the carbonisation reaction. The Maillard reac-
mary sewage sludge may not be effective at 140 °C. Mass balance tion is reported [24] to generate hundreds of heterocyclic com-
analyses show that a significant fraction of carbon was retained pounds that are mostly soluble in water. The percentage of
within the hydrochar following carbonisation of the PSS feedstock carbon transferred to the gas obtained from mass balances range
for 4 h (Fig. 7a). The fraction of carbon in the hydrochars as a from 12% to 22%, and increase as temperature is increased. Re-
percentage based on that of the untreated PSS feedstock decreased ported fractions of carbon to the liquid and gas phases for HTC of
as the temperature was increased (65% for carbonisation at 140 °C wastewater solids range from 20% to 37% and 2% to 11% respec-
to 55% for carbonisation at 200 °C). Carbon sequestered in the tively [17].
E. Danso-Boateng et al. / Applied Energy 111 (2013) 351–357 357

4. Conclusions processes and applications of wet and dry pyrolysis. Biofuels


2011;2(1):89–124.
[7] Titirici M-M, Antonietti M, Thomas A. Back in the black: hydrothermal
Decomposition kinetics for hydrothermal carbonisation of pri- carbonization of plant material as an efficient chemical process to treat the CO2
mary sewage sludge and synthetic faeces were investigated in a problem? New J Chem 2007;31:787–9.
[8] Cantrell K, Ro K, Mahajan D, Anjom M, Hunt PG. Role of thermal conversion in
batch reactor. Solid decomposition followed first-order reaction
livestock waste-to-energy treatments: obstacles and opportunities. Ind Eng
kinetics. Activation energy of primary sewage sludge was lower Chem Res 2007;46:8918–27.
than that for synthetic faeces. Solid decomposition was signifi- [9] Demir CR, Titirici M-M, Antonietti M, Cui G, Maier J, Hu Y-S. Hydrothermal
carbon spheres containing silicon nanoparticles: synthesis and lithium storage
cantly affected by reaction temperature more to a greater extent
performance. Chem Commun 2008;32:3759–61.
than the reaction time. Higher initial moisture content resulted [10] Ledakowicz S, Stolarek P. Kinetics of biomass thermal decomposition. Chem
in lower hydrochar yield and the extent of carbonisation was more Pap 2002;56(6):378–81.
evident in feedstock with lower moisture content. Energy contents [11] Antal Jr MJ, Gronli M. The art, science, and technology of charcoal production.
Ind Eng Chem Res 2003;42:1619–40.
of the hydrochars improved as the reaction temperature and time [12] Biagini E, Guerrini L, Nicholella C. Development of activation energy model for
increased. The experimental data fitted a first order reaction biomass devolatilization. Energy Fuels 2009;23:3300–6.
model, with the kinetic data providing a linear relation for the [13] Harun NY, Afzal MT, Shamsudin N. Reactivity studies of sludge and biomass
combustion. IJE 2009;3(5):413–25.
Arrhenius plot, indicating that the data presented in this study is [14] Ro KS, Cantrell KB, Hunt PG, Ducey TF, Vanotti MB, Szogi AA. Thermochemical
suited to global calculations and determination of kinetic data of conversion of livestock wastes: carbonisation of swine solids. Bioresour
sewage sludge and synthetic faeces HTC at various carbonisation Technol 2009;100:5466–71.
[15] Sadawi A, Jones JM, Williams A, Wojtowicz MA. Kinetics of the thermal
temperatures and operating strategies including for example decomposition of biomass. Energy Fuels 2010;24:1274–82.
continuous hydrothermal carbonisation. [16] Conesa JA, Marcilla A, Caballero JA, Font R. Comments on the validity and
utility of the different methods for kinetic analysis of thermogravimetric data.
J Anal Appl Pyrolysis 2001;58–59:619–35.
Acknowledgements
[17] Berge ND, Ro KS, Mao J, Flora JRV, Chappell MA, Bae S. Hydrothermal
carbonization of municipal waste streams. Environ Sci Technol
This research was part of Gates Foundation: ‘‘Reinventing the 2011;45:5696–703.
[18] Lu L, Namioka T, Yoshikawa K. Effects of hydrothermal treatment on
Toilet’’. The authors are grateful to Claire Millward for her initial
characteristics and combustion behaviours of municipal solid wastes. Appl
work on the hydrothermal carbonisation of synthetic faeces, and Energy 2011;88(11):3659–64.
to Geoffrey Russell for his help with collection of primary sewage [19] Cao X, Ro KS, Chappell M, Li Y, Mao J. Chemical structures of swine-manure
sludge. Reference to trade names or commercial products in this chars produced under different carbonization conditions investigated by
advanced solid-state 13C Nuclear Magnetic Resonance (NMR) spectroscopy.
publication is solely for the purpose of providing specific informa- Energy Fuels 2011;25:388–97.
tion and does not denote recommendation or endorsement. [20] Prawisudha P, Namioka T, Yoshikawa K. Coal alternative fuel production from
municipal solid wastes employing hydrothermal treatment. Appl Energy
2012;90:289–304.
References [21] Wang Z, Lin W, Song W. Liquid product from hydrothermal treatment of
cellulose by direct GC/MS analysis. Appl Energy 2012;97:56–60.
[1] Schouw NL, Danteravanich S, Mosbaek H, Tjell JC. Composition of human [22] Wignarajah K, Litwiller E, Fisher J, Hogan J. Simulated human feces for testing
excreta – a case study from Southern Thailand. Sci Total Environ human waste processing technologies in space systems. In: 36th International
2002;286:155–66. conference on environmental systems, SAE paper No. 2006-01-2180. Norfolk,
[2] Sobsey MD. Excreta and household wastewaters – introduction. Global Water. VA; 2006.
Sanitation and Hygiene; 2006; ENVR 890 section 003, ENVR 296 section 003. [23] Whitely N, Ozao R, Cao Y, Pan W. Multi-utilization of chicken litter as a
[3] Esrey SA. Towards a recycling society. Ecological sanitation – closing the loop biomass source. Part II. Pyrolysis. Energy Fuels 2006;20:2666–71.
to food security. In: Proceedings of the international symposium, 30–31 [24] Heilmann SM, Davis HT, Jader LR, Lefebvre PA, Sadowsky MJ, Schendel FJ, et al.
October, 2000. Bonn, Germany. GTZ, GmbH; 2001. Hydrothermal carbonisation of microalgae. Biomass Bioenergy
[4] Langergraber G, Muellegger E. Ecological Sanitation – a way to solve global 2010;34(6):875–82.
sanitation problems? Environ Int 2005;31:433–44. [25] Barlaz MA. Carbon storage factors during biodegradation of municipal solid
[5] WHO/UNICEF/WSSCC. Global Water Supply and Sanitation Assessment Report. waste components in laboratory-scale landfills. Global Biochem Cycles
WHO/UNICEF/WSSCC; 2000. 1998;12(2):373–80.
[6] Libra JA, Ro KS, Kammann C, Funke A, Berge N, Neubauer Y, et al. Hydrothermal
carbonization of biomass residuals: a comparative review of the chemistry,

You might also like