You are on page 1of 14

Environmental Science and Pollution Research

https://doi.org/10.1007/s11356-019-04613-4

RESEARCH ARTICLE

Catalytic CO2 gasification of rubber seed shell-derived hydrochar:


reactivity and kinetic studies
Pooya Lahijani 1 & Maedeh Mohammadi 2 & Abdul Rahman Mohamed 1

Received: 9 November 2018 / Accepted: 18 February 2019


# Springer-Verlag GmbH Germany, part of Springer Nature 2019

Abstract
In this study, hydrothermal carbonization (HTC) of a biomass was used as a means to improve the physicochemical properties of
rubber seed shell (RSS) and enhance its reactivity in the char-CO2 gasification reaction, known as the Boudouard reaction (C +
CO2 ↔ 2CO). Hydrochar samples were developed by hydrothermal treatment of RSS, without separating the solid residue from
the liquid product, at 433, 473, 513, and 553 K under autogenous pressure. The CO2 gasification reactivity of the developed
hydrochars was then investigated at different heating rates (5, 10, 20, and 30 K/min) by the non-isothermal thermogravimetric
method. The hydrochars revealed higher reactivity and improved gasification characteristics compared to the untreated biomass,
while the hydrochar which was filtered from the liquid slurry showed lower reactivity compared to the untreated biomass. This
was due to the chemical and structural evolutions of the biomass during hydrothermal treatment as indicated by various analyses.
The gasification reactivity of the hydrochar was substantially enhanced by introduction of a catalyst (NaNO3) during HTC.
Kinetic analysis of the char-CO2 gasification reaction was carried out by applying Flynn-Wall-Ozawa (FWO), Kissinger-
Akahira-Sunose (KAS), and Starink isoconversional methods, and thermodynamic parameters were also determined. The
activation energy of the Na-loaded RSS hydrochar in CO2 gasification (120–154 kJ/mol) was considerably lower than that of
the untreated biomass (153–172 kJ/mol). Thermodynamic studies also confirmed the promoting effect of hydrothermal treatment
and catalyst impregnation on enhancement of reactivity of the virgin biomass and reduction of gasification temperature.

Keywords Hydrothermal treatment . Hydrochar . CO2 gasification reactivity . Boudouard reaction . Kinetic studies

Introduction production of renewable energy and products from low-cost


local wastes and biomass residues.
Waste valorization with the goal of protecting the environment Malaysia is among the top-rated producers of natural rub-
and production of energy and fuel is emerging as a strong ber and ranks third in the world after Thailand and Indonesia
trend around the globe. Being aware of the importance of (Reshad et al. 2018). The rubber tree (Hevea brasiliensis),
sustainable development in waste prevention and reuse, many from which latex as source of natural rubber is harvested, is
countries have started to develop long-term plans for widely cultivated in Malaysia and around 1.3 million ha of
land is under rubber plantation (Hassan et al. 2014). Rubber
Responsible editor: Philippe Garrigues
seed is a by-product of the rubber tree whose kernel is rich in
oil, though the oil is non-edible and is potent for biodiesel
Electronic supplementary material The online version of this article
production (Onoji et al. 2016, 2017). During the oil extraction
(https://doi.org/10.1007/s11356-019-04613-4) contains supplementary
material, which is available to authorized users. process, a huge amount of rubber seed shell (RSS), compris-
ing around 40–48 wt% of the seed, is generated which poses
* Abdul Rahman Mohamed waste disposal issues to rubber seed oil millers, whereas this
chrahman@usm.my agricultural by-product can have promising applications as a
sorbent in the adsorption process (Hameed and Daud 2008;
1
Low Carbon Economy (LCE) Research Group, School of Chemical Borhan et al. 2016) and as a base catalyst in chemical reactions
Engineering, Universiti Sains Malaysia, 14300 Nibong Tebal, Pulau
Pinang, Malaysia
(Onoji et al. 2017). Yet, another potential application of this
2
waste residue could be for char production to be used as a
Faculty of Chemical Engineering, Babol Noushirvani University of
Technology, Babol 47148, Iran
versatile carbon precursor in the Boudouard reaction.
Environ Sci Pollut Res

CO 2 gasification of char, commonly known as the the reactivity of HTC char in combustion (Stirling et al.
Boudouard reaction, is a straightforward reaction to produce 2018; Lang et al. 2019), pyrolysis (Lin et al. 2016), and
the fuel gas, carbon monoxide (CO), from carbon dioxide and gasification processes (Ulbrich et al. 2017).
char (C (s) + CO2 (g) ↔ 2CO (g)). The reaction provides a Our intention in utilization of HTC, however, was not to
viable route for conversion of waste biomass in the form of develop a coal-like char. Our interest was in implementation
char to value-added product; meanwhile, CO2 as the primary of hydrothermal treatment as a process to partially destruct the
greenhouse gas is incorporated into a valorization cycle to lignocellulosic structure of biomass (RSS), so that the physi-
reduce to the fuel gas, CO. This reaction has long been the cochemical properties of the final product could be improved.
focus of research, especially in the area of coal gasification, In this study, HTC was implemented to develop hydrochar
and recently it has gained renewed interest in biomass gasifi- samples at different temperatures. The CO2 gasification reac-
cation. The Boudouard reaction is one of the equilibria which tivity of the hydrochars was then investigated by non-
occurs during the gasification of carbonaceous materials. isothermal thermogravimetry, and the results were compared
However, compared to the other reactions taking place during to the reactivity of char derived from untreated biomass. The
gasification such as evaporation and devolatilization, the rate hydrothermal impregnation of NaNO3 on RSS was also inves-
of this reaction is very slow and thus it is considered as the rate tigated, and the reactivity of the resulting metalized hydrochar
controlling step in the gasification process (Lahijani et al. in the CO2 gasification was evaluated. The kinetic and ther-
2015). To obviate this deficiency, several ongoing attempts modynamic studies were also carried out to determine the
have been made to enhance the reaction rate and improve activation energy and thermodynamic parameters of
the attainable carbon conversion. hydrochar CO2 gasification.
One of the attractive options to speed up the reaction rate
is the use of a catalyst. So far, several efforts have been
devoted to enhance the reactivity of biomass-derived char Experimental method
in the presence of alkali (Na, K) (Huang et al. 2009;
Bouraoui et al. 2016), alkaline earth (Ca, Mg) (Huang Preparation of hydrochars and CO2 gasification
et al. 2009; Perander et al. 2015), and transition (Fe) experiments
(Lahijani et al. 2012) metals. Incorporation of metal catalyst
into the char framework would improve the carbon surface Rubber seed shell (RSS) was collected from a rubber planta-
chemistry and thus encourages the heterogeneous char-CO2 tion area in Serdang, Perak, Malaysia. The shells were washed
reaction to obtain higher reactivities. It is also well under- and dried at 105 °C for 24 h. The RSS was then crushed and
stood that the char reactivity during gasification depends sieved to a particle size of 400 to 800 mm for HTC tests.
critically on the physicochemical properties of char and HTC experiments were conducted in an autoclave reactor.
gasification condition such as reaction temperature and In a typical run, approximately 10 g RSS was mixed with
pressure, heating rate, and so on (Lahijani et al. 2015). distilled water at 4:1 water/biomass ratio (w/w) in a 150-ml
The hydrothermal process wherein the feedstock is treated Teflon reaction vessel. The reactor was then placed in the
in a high-temperature, high-pressure aqueous environment stainless steel autoclave equipped with a stirring mechanism
is one of the promising approaches to obtain a char with and sealed. Prior to heating up the reactor, residual air was
improved physicochemical properties. At hydrothermal removed by flushing the reactor with nitrogen. The reactor
condition, the lignocellulosic structure of the biomass un- was then heated at a rate of 7 K/min to the desired temperature
dergoes a complex network of decomposition and degrada- (433, 473, 513, and 553 K) and kept isothermal for 30 min
tion reactions and the biomass is transformed to a solid while stirred at 400 rpm; the pressure buildup inside the reac-
product, commonly referred to as hydrochar (Ulbrich et al. tor was absolutely caused by the temperature augmentation.
2017). Hydrothermal treatment, also known as wet After completion of the reaction and once the reactor cooled
torrefaction and hydrothermal carbonization (HTC), has down to room temperature, it was dismantled and the contents
been widely studied to develop char samples with coal- (including the liquid fraction and solid residue) were collected
like properties. In fact, HTC is considered as a process and dried in an oven at 80 °C overnight. The obtained
which simulates natural coalification where HTC char can hydrochars were then ground to a particle size of 150 μm
be regarded as a coal substitute in combustion and gasifica- and stored for CO2 gasification experiments. The samples
tion processes (Fuertes et al. 2010; Hoekman et al. 2011). were labeled as HTC 433, HTC 473, HTC 513, and HTC
The focus of most HTC studies carried out so far has been on 553 hydrochars according to the corresponding HTC
the characterization of products and investigation of the temperatures.
influence of HTC condition on the product distribution In order to investigate the effect of alkali metal on enhance-
and mass balance (Hoekman et al. 2011; Ulbrich et al. ment of hydrochar reactivity during the CO2 gasification re-
2017; Stirling et al. 2018). Very few works investigated action, metalized hydrochar was prepared by hydrothermal
Environ Sci Pollut Res

impregnation. For this purpose, sodium salt was introduced gasification characteristic index S was determined using the
into the reaction medium, along with the biomass, during following equation (Li et al. 2017):
HTC. Aqueous solution of sodium nitrate (NaNO3, 5% mass
DTGmax :DTGmean
basis) was prepared and used as a reaction solvent. The exper- S¼ ð2Þ
iment was exactly the same as the described HTC method T 2i :T f
(non-catalytic) except that the NaNO3 solution was used in- where DTGmean is the mean value for gasification rate.
stead of distilled water. This sample was designated as Na-
HTC.
Kinetics and thermodynamic studies
To determine the elemental composition of the untreated
biomass and hydrochars prepared at different HTC conditions,
Non-isothermal thermogravimetric analysis (dynamic TGA
ultimate analysis was carried out using an elemental analyzer
process) is one of the most well-known and simplest methods
(Perkin-Elmer 2400 Series II CHNS/O). The ash, volatiles,
to analyze the kinetic and thermodynamic behaviors of the
and fixed carbon contents of the samples were measured fol-
biomass during thermal treatments. The rate of heterogeneous
lowing the standard ASTM method (D 5142–04). Fourier
solid state reaction to volatile products can be generally rep-
transform infrared (FTIR) analysis was used to investigate
resented by the Arrhenius equation (Barbanera et al. 2018):
the surface functional groups of the developed hydrochars
 
using a Thermo Scientific Nicolet iS10 spectrometer with dX EX
KBr pellet. To study the morphology of the prepared ¼ k ðT Þf ðX Þ ¼ Aexp − f ðX Þ ð3Þ
dt RT
hydrochars and also to analyze the chemical composition of
metal-loaded hydrochar, SEM-EDS mapping was conducted where dX/dt is the reaction rate, k represents the reaction rate
using a Hitachi S-3400N scanning electron microscope. constant, f (X) is the reaction model, A (s−1) is the pre-
Non-isothermal CO2 gasification experiments were carried exponential factor, R (kJ/mol.K) denotes the universal gas
out in a thermogravimetric analyzer (TGA, SDTQ-600). In constant, EX (kJ/mol) is the activation energy at a specific
each experiment, approximately 10 mg of hydrochar was conversion, and T (K) is the reaction temperature. Under
loaded in a ceramic pan and heated under pure N2 (100 ml/ non-isothermal condition with known heating rate, the relation
min) from room temperature to 1173 K at a heating rate of of temperature to time can be expressed as:
20 K/min; this was to avoid pyrolysis during gasification. The
T ¼ T 0 þ βt ð4Þ
sample was then cooled down under N2 to 900 K for subse-
quent CO2 gasification test. Next, the sample was heated up where T0 is the temperature at the onset of heating and β
under CO2 (100 ml/min) at heating rates of 5, 10, 20, and (K/min) is the heating rate. Using Eq. (4), the conversion rate
30 K/min to 1300 K to assure full conversion was achieved. (Eq. (3)) can be transformed to a temperature-dependent de-
In order to show the effect of HTC treatment on the CO2 rivative as follows:
gasification reactivity of hydrochars, the untreated biomass  
was used as reference sample. For this, pristine RSS was load- dX A EX
¼ exp − f ðX Þ ð5Þ
ed in TGA and underwent similar CO2 gasification tests as dT β RT
described. The weight loss of samples was continuously re-
corded as a function of time and used for calculations of char Rearranging Eq. (5) to separate the variables X and T and
conversion (X) according to the following equation: then integration yields:
 
m0 −mt X dX T A EX
X ¼  100% ð1Þ ∫0 ¼ ∫T 0 exp − dT ð6Þ
m0 −mash df ðX Þ β RT
T0 A 
Since ∫0 β exp −RT
EX
dT ¼ 0, Eq. (6) is simplified as:
where m0, mt, and mash represent the initial mass of the sample,
instantaneous mass of the sample at time t, and mass of ash  
T A EX
after completion of gasification, respectively. GðX Þ ¼ ∫0 exp − dT ð7Þ
β RT
In order to quantitatively analyze the gasification reactivity
of different samples, some gasification performance indices For determination of most relevant kinetic parameters (A
can be derived from TG-DTG curves. These gasification char- and EX), various methods including isoconversional and
acteristic parameters include the initiating gasification temper- model-fitting methods have been proposed. The
ature (Ti), finishing gasification temperature (Tf), peak temper- isoconversional methods, which are in fact model-free, as-
ature at which the gasification rate is maximum (Tm), maxi- sume that the function f(X) is not influenced by the variation
mum gasification rate (DTGmax), and gasification time (tg). of the heating rate. These methods rely on experimental mea-
Using these performance indices, the comprehensive surement of temperature corresponding to a fixed value of
Environ Sci Pollut Res
 
conversion at different heating rates. The isoconversional KBT m
ΔGX ¼ EX þ RT m ln ð13Þ
methods are considered as safe alternatives for calculation of hA
accurate activation energy, without knowing the kinetic ΔH X −ΔGX
models of the reaction mechanism (Marinović-Cincović ΔS X ¼ ð14Þ
Tm
et al. 2013). In this study, activation energy was calculated
using three isoconversional methods, i.e., Flynn-Wall-Ozawa where KB is the Boltzmann constant (1.381 × 10−23 J/K) and h
(FWO), Kissinger-Akahira-Sunose (KAS), and Starink represents the Planck constant (6.626 × 10−34 Js).
methods.
The FWO method which is based on Doyle’s approxima-
tion is the most commonly used isoconversional method. It is Results and discussion
expressed as follows (Yuan et al. 2017):
    CO2 gasification of hydrochars and reactivity
AE X EX characteristics
lnβ ¼ ln −5:3305−1:052 ð8Þ
RGðX Þ RT
The developed hydrochar samples were gasified with CO2
At each degree of conversion, activation energy can be
under a non-isothermal condition in TGA and their gasifica-
calculated from the slope of the straight line obtained by the
tion and reactivity characteristics were studies. Figure 1 shows
plot of ln β versus 1/T.
the conversion (dotted curves) and conversion rate (DTG,
The KAS equation which is considered more accurate com-
dashed curves) against gasification temperature for the un-
pared to the FWO method is stated as below (Hardi et al.
treated biomass and hydrochars at several heating rates from
2018):
5 to 30 K/min. For all samples, with increase of heating rate
   
β AR EX from 5 to 30 K/min, the conversion and conversion rate
ln 2 ¼ ln − ð9Þ
T E X g ðX Þ RT (DTG) curves shifted towards the higher-temperature region
due to heat transfer limitation. It was also observed that for a
The plot of ln(β/T2) versus (1/T) for a certain value of X given temperature, with increase of the heating rate, the real-
will develop a straight line from which the activation energy is ized conversion increased; alternatively, to reach a specified
calculated. conversion, higher temperature was required at higher heating
The Starink method was developed to increase the accura- rates.
cy of the KAS method in the following form (Hardi et al. Generally, the evolution of the gasification reaction can be
2018): described as a three-stage process, where in the first stage the
  sample is heated up, in the second stage gasification takes
β EX
ln 1:92 ¼ C s −1:0008 ð10Þ place and proceeds to the complete gasification in the third
T RT
stage. Considering that gasification mainly occurs in the sec-
At each degree of conversion, the slope of the plot of ln(β/ ond phase, this stage is the focus of kinetic studies. Table 1
T1.92) versus (1/T) is used for the calculation of the activation summarizes the characteristic parameters of CO2 gasification
energy. of the untreated biomass and hydrochars.
Although the values of the activation energy can be accu- Results show that for all samples with increase of heating
rately provided by the isoconversional methods, these models rate, the temperatures Tm and Tf obviously increased; the
do not yield the pre-exponential factor. For determination of DTGmax and reactivity index S also increased while the gasi-
the pre-exponential factor, Kissinger’s equation can be used fication time shortened. Since the comprehensive gasification
(Dhyani et al. 2017): performance index reflects the gasification reactivity, a higher
  value indicates a more satisfactory gasification performance
EX (Barbanera et al. 2018). Thus, it can be concluded that with the
βE X exp
RT m increase of heating rate, the gasification reactivity improved.
AX ¼ ð11Þ
RT m2 Lower values of Tm also indicate the higher reactivity (Li et al.
2017). Comparing the peak temperature and reactivity index
where Tm is the temperature at which the reaction rate is max- of all samples inferred that the HTC 513 hydrochar with Tm in
imum. After determination of the activation energy, the ther- the range of 1099 to 1247 K and S index of 1.983 × 10−12 to
modynamic parameters including enthalpy (ΔH), entropy 17.074 × 10−12 1/min2 K3 was the most reactive sample at all
(ΔS), and Gibbs free energy (ΔG) can be calculated using tested heating rates. While the untreated biomass with Tm in
the following equations (Barbanera et al. 2018): the range of 1131 to 1285 K and S index of 1.429 × 10−12 to
8.969 × 10−12 1/min2 K3 was the least reactive sample. This
ΔH X ¼ EX −RT ð12Þ
implies the positive effect of HTC treatment on enhancement
Environ Sci Pollut Res

Fig. 1 Conversion and DTG


curves of the untreated biomass
and hydrochars

of hydrochar gasification reactivity. As results indicate, all benefit provides the possibility of operating the gasifier at a
hydrochars revealed lower peak temperature compared to lower temperature which is an advantage in terms of energy
the untreated biomass gasified at the same condition. This and cost savings.

Table 1 Characteristic
gasification parameters of the HTC β Ti Tm (K) Tf (K) DTGmax S × 1012 tg
untreated biomass and hydrochars condition (K/min) (K) (1/min) (1/min2 K3) (min)
at different heating rates
Untreated 5 950.2 1130.7 1148.2 0.079 1.420 43.0
10 949.5 1176.9 1190.6 0.146 3.880 31.7
20 948.3 1228.1 1253.8 0.195 7.906 15.8
30 950.1 1284.7 1304.1 0.209 8.969 12.1
HTC 433 5 950.1 1115.7 1131.7 0.083 1.492 36.7
10 952.2 1168.0 1179.5 0.130 4.059 20.8
20 951.4 1218.4 1234.1 0.196 8.281 14.1
30 951.3 1257.2 1291.0 0.223 10.333 11.9
HTC 473 5 948.5 1110.4 1127.7 0.077 1.704 36.4
10 958.9 1151.0 1160.3 0.135 3.868 22.5
20 950.5 1198.5 1236.7 0.205 7.926 14.3
30 950.4 1248.3 1289.1 0.281 16.832 11.8
HTC 513 5 950.1 1098.7 1112.3 0.078 1.983 35.2
10 950.3 1123.8 1152.8 0.131 4.160 20.1
20 955.7 1192.9 1229.0 0.249 12.998 13.9
30 951.2 1246.7 1275.0 0.285 17.074 10.0
HTC 553 5 950.1 1114.9 1130.3 0.082 1.449 36.9
10 951.4 1153.4 1165.7 0.142 4.037 21.7
20 951.3 1207.9 1231.8 0.234 11.995 14.2
30 950.5 1253.7 1285.2 0.283 15.789 10.4
Environ Sci Pollut Res

Chemical and structural evolutions of the biomass in agreement with findings of other researchers (Hoekman
during HTC treatment et al. 2011; Ulbrich et al. 2017; Stirling et al. 2018) who
reported that HTC encourages coalification and with increase
Investigation on the CO2 gasification reactivity of hydrochars of process severity coalification progressively increases. It is
showed that all HTC-treated samples were more reactive than worth to re-emphasize here that our intention in utilization of
untreated biomass as discussed in the BCO2 gasification of hydrothermal treatment was not to develop a coal-like fuel
hydrochars and reactivity characteristics^ section. Also, the with enhanced energy content, but to promote the reactivity
reactivity of the hydrochar samples was different depending of the resulting hydrochar during the CO2 gasification pro-
on the HTC temperature. To explain and justify such differ- cess. To this end, we made a slight change in the typical
ences in reaction rate, several analyses were carried out on HTC experiments in that we did not separate the liquid prod-
biomass and hydrochar samples which are discussed here. uct of the hydrothermal process from the resulting solid resi-
The results of the CHNS/O analysis for RSS and HTC- due (hydrochar), as others did. In the preliminary HTC tests,
treated RSS at different temperatures are presented in when we separated the liquid product from the solid residue
Table 2. The changes in atomic composition are attributed to (filtered hydrochar), the reactivity of the resulting hydrochar
the reactions that take place during HTC treatment. The chief was less than that of the non-filtered hydrochar and even the
reactions are hydrolysis, dehydration, decarboxylation, and untreated biomass as depicted in Fig. 2. Thus, after finishing
condensation which result in loss of some carbon, oxygen, the HTC tests, we did not filter out the slurry to isolate the
and hydrogen. The major loss, however, is associated with solid and aqueous products; rather, we dried the whole slurry
oxygen whose removal can occur either by decarboxylation and let the oily aqueous fraction to deposit on the solid prod-
in the form of carbon dioxide or dehydration in the form of uct. With this, losses in organic and inorganic constituents
water (Posmanik et al. 2018). Results showed that upon hy- were negligible, as confirmed by CHNS/O and ash content
drothermal treatment, the carbon content of hydrochars in- (proximate) analyses, respectively. This was because most
creased, while the oxygen content decreased and the hydrogen carbon losses in the hydrothermal treatment are attributed to
content changed slightly. The reduction in the atomic O/C the dissolved organics in the liquid phase and only minor
ratio indicates that the decarboxylation reaction prevailed dur- amounts are lost in gas products (Kruse et al. 2013); in our
ing HTC which resulted in carbon enrichment and oxygen tests, the carbon which was lost in the aqueous phase almost
depletion. Though, dehydration reaction also, to a lesser ex- fully recovered. Thus, the decrease in atomic O/C and H/C
tent, contributed to the oxygen removal as was realized from was not that severe to develop a coal-like product; the HTC-
the FTIR analysis. treated RSS samples still remained in the biomass area as
To analyze the chemical transformations that occur during illustrated in Fig. S1. Retaining the biomass properties was
the HTC treatment of biomass, use of the van Krevelen dia- an advantage from a reactivity point of view, as formation of
gram (Fig. S1, Supplementary Material) is helpful. This dia- structures resembling graphite and anthracite within the bio-
gram is commonly used to examine the degree of coalification mass body will develop a more stable and less reactive char
during the hydrothermal process (Hoekman et al. 2011; structure with increased activation energy (Ulbrich et al. 2017;
Ulbrich et al. 2017). In the diagram, the starting RSS feed- Lane et al. 2018).
stock can be found in the upper right corner, within the bio- The HTC process is also known to accelerate the solubili-
mass area. Upon hydrothermal treatment, the biomass zation of inorganic matters especially due to the acidic nature
underwent some decarboxylation and dehydration reactions, of aqueous products (Fuertes et al. 2010) which may eventu-
yet the resulting hydrochars (HTC 433, HTC 473, HTC 513, ally reduce the reactivity of the hydrochar. However, in our
and HTC 553) are still in the typical biomass area. This is not experiment, since the solid and liquid fractions were not

Table 2 Ultimate and proximate


analyses of raw and HTC-treated Sample Elemental analysis (wt%, db) Proximate analysis (wt%, db) Mole ratio
RSS
C H N Oa S VM FC Ash H/C O/C

Raw RSS 48.36 5.75 0.14 45.45 0.30 77.4 21.9 0.4 1.42 0.71
HTC 433 51.24 6.06 0.24 42.14 0.32 76.1 23.0 0.9 1.41 0.62
HTC 473 51.35 5.40 0.22 42.72 0.31 74.6 24.6 0.8 1.25 0.62
HTC 513 51.57 5.17 0.20 42.77 0.29 75.7 23.3 1.0 1.19 0.62
HTC 553 50.54 5.80 0.01 43.42 0.23 72.8 26.3 0.9 1.34 0.63

db dry basis, VM volatile matter, FC fixed carbon


a
Calculated by difference
Environ Sci Pollut Res

increase of the temperature. This was because of the dehydra-


tion of water at high temperatures and dissociation of hydrox-
yl groups from sugar rings (Yang et al. 2015). The peak locat-
ed at 1730 cm−1 is attributed to the presence of C=O stretching
vibration of acetyl in hemicellulose (Yan et al. 2017). The
intensity of this peak decreased gradually at elevated temper-
atures in favor of fracture of acetyl (Hill et al. 2013; Dai et al.
2017). The peaks at 1244 and 1024 cm−1 were C-O stretching
vibration in either esters or ethers and the alcohol group, re-
spectively (Lua and Yang 2004). A similar trend in the inten-
sity of both peaks was observed where by increasing the pre-
treatment temperature the intensity of the peaks reduced. The
decrease in the intensity of the peaks attributing to the hydrox-
Fig. 2 Comparison of the conversion profiles of the untreated biomass yl and carboxyl groups with increase of HTC temperature
and filtered and non-filtered hydrochars at the heating rate of 20 K/min suggests that decarboxylation and dehydration were the major
reactions that contributed to the loss of organic constituents as
separated, the HTC-treated samples retained the mineral ma- measured by elemental analysis.
ters and had almost the same ash contents as that of the orig- From the results, it could be inferred that chemical trans-
inal feedstock as implied from the proximate analysis present- formation and changes in the atomic C, H, and O composition
ed in Table 2. This was another benefit of our experiment of the RSS biomass during HTC treatment were insignificant.
wherein demineralization of the biomass was hindered. The It is thus reasonable to assume that the observed enhancement
results of proximate analysis also showed that HTC treatment in the CO2 gasification reactivity of hydrochars compared to
of the biomass reduced the concentration of volatile matters the untreated biomass was attributed to the changes in struc-
while the fixed carbon concentration increased. It is most like- tural properties of the biomass during hydrothermal treatment.
ly that most of the volatiles released from the biomass matrix To investigate the surface morphology of the raw biomass and
were in the form of CO2 resulted from the decarboxylation hydrochars and also to observe the possible alterations of the
reaction as discussed. structure during HTC treatment, SEM analysis was carried out
The FTIR analysis of raw and HTC-treated RSS samples, and the results are shown in Fig. 4. The raw RSS revealed a
as exhibited in Fig. 3, showed that the transmittance peaks thick and fiber-like structure with a rough but unbroken tex-
observed for the HTC-treated RSS samples were almost sim- ture. Upon the utilization of hydrothermal treatment and with
ilar to that of raw RSS; however, the intensity of most peaks the increase of the HTC temperature, the biomass underwent a
decreased by increase of the HTC temperature. The wide peak series of decomposition and polymerization reactions which
around 3400 cm−1 assigning to the –OH stretching vibration changed the surface morphology of the biomass. During the
in carboxyl and hydroxyl groups significantly decreased by hydrothermal treatment, with the increase of the temperature,
the hot compressed water changes its physical and chemical
properties significantly and acts as an excellent reaction me-
dium, reactant, and solvent (mimicking organic solvents) for a
diverse range of reactions (Berge et al. 2011; Kruse et al.
2013). The first reaction that takes place during the hydrother-
mal treatment of biomass is hydrolysis through which hemi-
cellulose and cellulose are decomposed to pentose and hexose
sugars and lignin is destructed to its phenolic constituents.
Decomposition of hemicellulose and amorphous cellulose ini-
tiated at around 440 K as could be implied from the DTG
profile of RSS (Fig. S2, Supplementary Material); this result-
ed in the creation of some cracks and fractures on the biomass
surface as can be observed in SEM images. However, with
increase of the HTC temperature to 513 K, the changes in the
biomass structure were more pronounced where the fiber
structure broke down and the rough fractures became wide.
This was probably due to the decomposition of hemicellulose
and most of cellulose, along with partial degradation of lignin
Fig. 3 FTIR spectra of raw and HTC-treated RSS samples which has greater thermal stability. According to the literature
Environ Sci Pollut Res

Fig. 4 SEM micrographs of raw


and HTC-treated RSS samples

(Nizamuddin et al. 2017), with increase of hydrothermal tem- CO2 gasification of the metalized hydrochar
perature, the degree of aromatization of the carbon structure
increases and a well-organized char is developed. Comparison During the gasification reaction, the overall reaction rate can
of the SEM micrographs of HTC 513 and HTC 553 indicates be influenced by both non-catalytic and catalytic gasification
that the rough fractures on the HTC 513 surface are irregular rates; the first is associated with physicochemical properties of
and discordant, while those of HTC 553 are somewhat regular char and the latter is affected by the mineral content of char
and a higher degree of ordered structure can be observed. (Lahijani et al. 2015). Considering that the ash content of RSS
Considering that about 15% of cellulose is comprised of an is very low, around 0.4 wt% in the biomass and 1.0 wt% in the
amorphous structure which is very much vulnerable to hydro- hydrochar, the role of inherent catalytic elements of char in
lysis, it is very much likely that with increase of hydrothermal promoting the gasification rate would be insignificant. This is
temperature from 513 to 553 K and degradation of amorphous while the reaction of char-CO2 is slow and highly endothermic
cellulose, the share of crystalline cellulose increased and left and catalytic species should be involved to speed up the gas-
an almost uniform structure. The higher share of ordered car- ification rate. Alkali and alkaline earth metals have been found
bon in HTC 553 and its structural uniformity eventually con- as good catalysts for CO2 gasification of char. Our previous
tributed to the lower CO2 gasification reactivity of this investigation (Lahijani et al. 2013) on the effect of alkali,
hydrochar compared to HTC 513. alkaline earth, and transition metal nitrates on the CO 2
Environ Sci Pollut Res

rates. The peak temperature of this sample was in the range


of 1044 to 1140 K, while this measure was in the range of
1099 to 1247 K for HTC 513. This indicates that by virtue of
the use of catalytic gasification, the reaction temperature can
be lowered; alternatively, the conversion at a specific temper-
ature can be improved compared to the non-catalytic reaction.
As prior discussed, while comparing the reactivity of dif-
ferent samples, two important parameters are peak tempera-
ture and reactivity index. They can be conveniently plotted as
illustrated in Fig. 6. The most reactive sample is the one that
has high reactivity index at low temperatures; this char must
occupy the top-left region of the graph (Fig. 6). On the con-
Fig. 5 Conversion and DTG curves of the Na-loaded hydrochar (Na- trary, the char with least reactivity locates in the bottom-right
HTC 513) area on the graph. At each heating rate, Na-HTC 513 had the
highest reactivity, followed by HTC 513 and the untreated
gasification reactivity of biomass-derived char showed the biomass had the lowest reactivity. The use of metal catalyst
superior performance of NaNO3 over other implemented cat- and hydrothermal treatment contributed to the enhanced reac-
alysts. In a continuing effort to enhance the char-CO2 reaction tivity of the virgin biomass and reduction of gasification
rate, we used NaNO3 to develop Na-loaded hydrochar through temperature.
hydrothermal impregnation. The sample HTC 513 which Promotion of the char reaction rate in the presence of cat-
showed the highest reactivity among the hydrochars was used alyst can be assumed to be attributed to the increase of con-
to prepare Na-loaded hydrochar. The EDS analysis confirmed centration of active sites that play a significant role in the
the presence of Na on the hydrochar surface (Fig. S3, progress of the reaction. To verify this speculation, the char
Supplementary Material). Elemental mapping in this figure surface active sites which include both carbon active sites and
also showed the uniform distribution of Na on the surface of catalytic actives sites were measured by performing the CO2
the hydrochar. chemisorption test at 300 °C in TGA, following the method
Figure 5 depicts the conversion and conversion rate of the proposed in the literature (Xu et al. 2013; Gupta et al. 2018).
Na-HTC 513 hydrochar at various heating rates from 5 to The CO2 chemisorption curves for Na-HTC 513 and HTC 513
30 K/min. The catalyzed hydrochar showed higher reactivity chars (prepared at 900 °C) are shown in Fig. 7. Weak chem-
than the un-catalyzed hydrochar at all implemented heating isorption (Cweak) which refers to the amount of CO2 being
Fig. 6 Plots of the comprehensive
gasification reactivity index and
the peak temperature for raw and
HTC-treated biomass samples at
various heating rates
Environ Sci Pollut Res

Fig. 7 CO2 chemisorption capacity of HTC 315 and Na-HTC 315 chars

desorbed from the char surface during the outgassing process


can be a measure of carbon (organic) active sites. While the
strong chemisorption (Cstrong) that shows the amount of CO2
which is still adsorbed on the char surface can denote the
catalytic active sites, granted by inorganic matter of the char.
As observed in the figure, addition of catalyst increased the
concentration of catalytic active sites, as indicated by higher
Cstrong. It is thus reasonable to conclude that the higher CO2
gasification reactivity of the catalyzed hydrochar was in close
relation with the higher concentration of surface active sites in
this sample as compared to the un-catalyzed hydrochar.

Activation energy and thermodynamic analysis

The activation energy of the char-CO2 gasification reaction


was determined using the three isoconversional methods,
i.e., FWO, KAS, and Starink methods, at conversion levels
of 0.1–0.8. Activation energies at each conversion degree
were obtained by developing the corresponding linear plots.
The changes of activation energies, obtained from the three
methods, with degree of conversion are shown in Fig. 8. The
results obtained from the three methods were almost similar
and comparable among the different methods. The Ex values
computed from FWO were slightly higher than those obtained Fig. 8 Variations of activation energies with evolution of gasification
from the KAS and Starink methods. However, the high regres- reaction obtained from FWO, KAS, and Starink models
sion coefficients of all plots (data not shown) reflect that all the
three methods were suitable to describe the CO2 gasification The activation energy of HTC 513 that showed the highest
behavior of char samples. The three activation energy curves reactivity among the un-catalyzed hydrochars was 135–
displayed a similar descending tendency with increase of the 161 kJ/mol. Nevertheless, deposition of sodium metal on the
conversion degree. This signifies that the greater the extent of sample significantly lowered the activation energy to 120–
char conversion, the lower the activation energy was required 154 kJ/mol. The activation energy of Na-HTC 513 substan-
to initiate and sustain the reaction. This can be attributed to the tially decreased compared to the untreated biomass. Such en-
increase of the portion of inorganic components with progress hancement in the reactivity of the metal-loaded hydrochar and
of the reaction and contribution of these constituents to the reduction of activation energy was resulted from hydrothermal
promotion of the gasification reaction. The untreated biomass treatment of biomass and use of catalytic species to increase
was the least reactive sample and showed the highest activa- the char surface active sites.
tion energy (153–172 kJ/mol); however, when the biomass The activation energies obtained from the Starink
was hydrothermally treated, the activation energy reduced. method were used to calculate the pre-exponential factor
Environ Sci Pollut Res

Table 3 Kinetic and thermodynamic parameters during CO2 gasification of char samples calculated by the Starink method at conversions of 0.1–0.8

Sample Conversion (X) E (kJ/mol) A (s−1) ΔH (kJ/mol) ΔG (kJ/mol) -ΔS (J/mol.K)

Untreated biomass 0.1 172.02 9.64E+04 163.25 366.62 166.27


0.2 168.11 6.79E+04 158.93 365.91 169.25
0.3 157.81 2.26E+04 148.39 366.60 178.43
0.4 152.75 1.31E+04 143.19 366.95 182.98
0.5 146.02 6.37E+03 136.35 367.44 188.98
0.6 148.54 8.36E+03 138.80 367.26 186.82
0.7 152.19 1.23E+04 142.35 366.99 183.70
0.8 153.13 1.37E+04 143.22 366.93 182.93
Average 156.32
HTC 433 0.1 167.72 6.38E+04 158.94 363.55 169.37
0.2 160.86 3.08E+04 151.81 363.90 175.57
0.3 155.81 1.80E+04 146.49 364.18 180.20
0.4 146.41 6.59E+03 137.01 364.71 188.50
0.5 144.75 5.52E+03 135.22 364.81 190.06
0.6 149.40 9.07E+03 139.77 364.54 186.06
0.7 142.60 4.39E+03 132.91 364.94 192.08
0.8 139.94 3.30E+03 130.19 365.10 194.47
Average 150.94
HTC 473 0.1 154.81 6.28E+04 146.20 358.68 175.59
0.2 145.92 8.62E+03 136.95 359.27 183.72
0.3 149.81 1.32E+04 140.58 359.01 180.50
0.4 151.06 1.51E+04 141.71 358.93 179.51
0.5 153.07 1.88E+04 143.58 358.80 177.85
0.6 150.04 1.35E+04 140.47 358.99 180.58
0.7 146.18 8.87E+03 136.52 359.25 184.07
0.8 143.42 6.56E+03 133.70 359.44 186.55
Average 149.29
HTC 513 0.1 161.28 5.85E+04 152.71 354.20 170.02
0.2 148.43 1.40E+04 139.39 355.02 181.95
0.3 151.65 2.01E+04 142.57 354.81 179.09
0.4 143.81 8.40E+03 134.58 355.33 186.28
0.5 134.46 2.96E+03 125.11 356.00 194.83
0.6 133.95 2.80E+03 124.50 356.03 195.38
0.7 135.38 3.28E+03 125.85 355.93 194.15
0.8 135.33 3.26E+03 125.72 355.93 194.26
Average 143.04
HTC 553 0.1 161.11 5.11E+04 152.33 360.13 172.89
0.2 145.87 7.83E+03 136.81 361.10 186.62
0.3 141.83 5.04E+03 132.51 361.38 190.43
0.4 138.43 3.48E+03 129.00 361.62 193.56
0.5 148.34 1.03E+04 138.82 360.94 184.82
0.6 146.99 8.85E+03 137.38 361.03 186.08
0.7 148.49 1.04E+04 138.80 360.93 184.82
0.8 145.28 7.35E+03 135.53 361.14 187.72
Average 147.04
Na-HTC 513 0.1 153.99 4.09E+04 145.62 329.42 169.31
0.2 142.42 2.21E+03 133.88 330.13 180.80
0.3 119.18 1.42E+03 110.47 331.74 203.87
0.4 114.29 7.91E+02 105.45 332.12 208.85
0.5 115.77 9.44E+02 106.85 332.00 207.45
0.6 117.10 1.11E+03 108.09 331.90 206.21
0.7 117.50 1.16E+03 108.42 331.87 205.88
0.8 119.80 1.52E+03 110.62 331.69 203.69
Average 125.10

as well as thermodynamic parameters. The values of the energies and are mainly attributed to the presence of po-
pre-exponential factor were calculated by substitution of rous carbon which makes the reaction faster and easier
EX values into Kissinger’s equation; the results are pre- (Barbanera et al. 2018). The A values between 104 and
sented in Table 3. The pre-exponential factor is a kinetic 109 s−1 indicate a system with less reactivity and show
parameter related to the structure of material (Edreis et al. surface reaction in most cases (Dhyani et al. 2017;
2018). Values less than 104 s−1 show lower activation Barbanera et al. 2018). The lower values of A (7.91 ×
Environ Sci Pollut Res

102–4.09 × 104 s−1) and EX (120–154 kJ/mol) obtained for 2013). According to the thermodynamic properties, it
Na-HTC 513 indicate a faster and easier reaction of char could be concluded that hydrothermal treatment of the
with CO2 attributing to the synergetic effect of hydrother- biomass and introduction of a catalyst onto the char skel-
mal treatment and catalyst deposition on the char surface. eton were effective to enhance the reactivity and favor-
The variations of ΔH, ΔG, and ΔS with respect to the con- ability of the reaction.
version are summarized in Table 3. Since the calculated values
of thermodynamic parameters at different heating rates were
very close or even superimposed each other, average values
were reported in the table. According to the results presented Conclusion
in Table 3, enthalpy change was positive in all char-CO2 gas-
ification reactions, indicating the endothermic nature of the In this study, HTC of biomass was used to develop
reaction where the higher the value of ΔH, the more heat hydrochar samples with enhanced CO2 gasification reac-
energy was required. The higher ΔH value for untreated bio- tivities. With some minor modifications in the typical HTC
mass demonstrates that gasification of raw RSS required the process (not separating the liquid product from the solid
introduction of more heat compared to the hydrochars. residue), we could develop hydrochars with higher reactiv-
Moreover, in most cases, a decreasing trend in ΔH value was ity and lower activation energy in CO2 gasification com-
observed as conversion proceeded, which indicates a reduc- pared to the char developed from the untreated biomass.
tion in endothermicity of the reaction with progression of the Physicochemical analyses indicated that the enhanced CO2
reaction. The value of ΔH for the untreated biomass, HTC gasification reactivities of hydrochars were attributed to
513, and Na-HTC 513 was 143–163, 126–153, and 111– the chemical and structural evolutions of the biomass dur-
146 kJ/mol, respectively, which is consistent with the trend ing the hydrothermal treatment. The reactivity of the
of activation energy. hydrochar was further improved by developing the Na-
The ΔG value is an indication of the stability of the loaded RSS hydrochar through hydrothermal impregnation
system where lower values of this parameter imply higher of NaNO3 on the biomass. The activation energy of the
favorability for the reaction. Results show that variation untreated biomass in CO2 gasification was 153–172 kJ/
of ΔG with conversion was insignificant for all samples mol; nevertheless, use of the HTC process and loading of
and no obvious change was observed. However, the ΔG metal on the char substantially lowered the activation en-
values of the hydrochars were lower than that of the un- ergy to 120–154 kJ/mol. Thermodynamic studies also con-
treated biomass, suggesting the increased favorability of firmed that hydrothermal treatment of biomass and use of
char-CO2 gasification after hydrothermal treatment. The catalyst were helpful to enhance the reactivity of the
Na-HTC 315, with the lowest Gibbs free energy (~ resulting hydrochars in the CO2 gasification reaction.
331 kJ/mol), had the highest favorability. The positive
Acknowledgments This research was supported by the Universiti Sains
values of ΔH and ΔG for all char-CO2 reactions show that Malaysia through LRGS Grants 203/PJKIMIA/6720009 and 304/
the reactions require the introduction of heat and are non- PJKIMIA/6050376.K100.
spontaneous processes.
Entropy of a system is a state function that shows the Publisher’s note Springer Nature remains neutral with regard to jurisdic-
degree of disorder. The entropy change reflects how close tional claims in published maps and institutional affiliations.
the system is to its thermodynamic equilibrium. When
low activation entropy values are observed, it means that
the material is close to its thermodynamic equilibrium and
References
is less reactive. On the contrary, higher activation entro-
pies show that the material is far from its thermodynamic Barbanera M, Cotana F, Di Matteo U (2018) Co-combustion performance
equilibrium and is more reactive (Georgieva et al. 2012). and kinetic study of solid digestate with gasification biochar. Renew
The entropy of the Na-HTC 315 hydrochar was the Energy 121:597–605. https://doi.org/10.1016/J.RENENE.2018.01.
076
highest, followed by the HTC 315 hydrochar, suggesting
Berge ND, Ro KS, Mao J, Flora JRV, Chappell MA, Bae S (2011)
that these materials are in a state far from their own ther- Hydrothermal carbonization of municipal waste streams. Environ
modynamic equilibrium and can react faster in shorter Sci Technol 45:5696–5703. https://doi.org/10.1021/es2004528
reaction times. The negative value of ΔS indicates that Borhan A, Abdullah NA, Rashidi NA, Taha MF (2016) Removal of Cu2+
the formation of activated complex is along with the de- and Zn2+ from single metal aqueous solution using rubber-seed shell
based activated carbon. Procedia Eng 148:694–701. https://doi.org/
crease of entropy, that is to say the activated complex is a
10.1016/J.PROENG.2016.06.571
more organized structure in comparison to the initial sub- Bouraoui Z, Dupont C, Gadiou R (2016) CO2 gasification of woody
stance. The reactions with negative ΔS are classified as biomass chars: the influence of K and Si on char reactivity. CR
Bslow^ (Georgieva et al. 2012; Marinović-Cincović et al. Chim 19:457–465. https://doi.org/10.1016/J.CRCI.2015.08.012
Environ Sci Pollut Res

Dai L, He C, Wang Y, Liu Y, Yu Z, Zhou Y, Fan L, Duan D, Ruan R Lane DJ, Truong E, Larizza F, Chiew P, de Nys R, van Eyk PJ (2018)
(2017) Comparative study on microwave and conventional hydro- Effect of hydrothermal carbonization on the combustion and gasifi-
thermal pretreatment of bamboo sawdust: hydrochar properties and cation behavior of agricultural residues and macroalgae:
its pyrolysis behaviors. Energy Convers Manag 146:1–7. https://doi. devolatilization characteristics and char reactivity. Energy Fuel 32:
org/10.1016/J.ENCONMAN.2017.05.007 4149–4159. https://doi.org/10.1021/acs.energyfuels.7b03125
Dhyani V, Kumar J, Bhaskar T (2017) Thermal decomposition kinetics of Lang Q, Zhang B, Liu Z, Chen Z, Xia Y, Li D, Ma J, Gai C (2019) Co-
sorghum straw via thermogravimetric analysis. Bioresour Technol hydrothermal carbonization of corn stalk and swine manure: com-
245:1122–1129. https://doi.org/10.1016/J.BIORTECH.2017.08. bustion behavior of hydrochar by thermogravimetric analysis.
189 Bioresour Technol 271:75–83. https://doi.org/10.1016/J.
Edreis EMA, Li X, Luo G, Sharshir SW, Yao H (2018) Kinetic analyses BIORTECH.2018.09.100
and synergistic effects of CO2 co-gasification of low sulphur petro- Li R, Zhang J, Wang G, Ning X, Wang H, Wang P (2017) Study on CO2
leum coke and biomass wastes. Bioresour Technol 267:54–62. gasification reactivity of biomass char derived from high-
https://doi.org/10.1016/J.BIORTECH.2018.06.089 temperature rapid pyrolysis. Appl Therm Eng 121:1022–1031.
Fuertes AB, Arbestain MC, Sevilla M, Maciá-Agulló JA, Fiol S, López https://doi.org/10.1016/J.APPLTHERMALENG.2017.04.132
R, Smernik RJ, Aitkenhead WP, Arce F, Macías F (2010) Chemical Lin Y, Ma X, Peng X, Yu Z, Fang S, Lin Y, Fan Y (2016) Combustion,
and structural properties of carbonaceous products obtained by py- pyrolysis and char CO2-gasification characteristics of hydrothermal
rolysis and hydrothermal carbonisation of corn stover. Aust J Soil carbonization solid fuel from municipal solid wastes. Fuel 181:905–
Res 48:618. https://doi.org/10.1071/SR10010 915. https://doi.org/10.1016/J.FUEL.2016.05.031
Georgieva V, Zvezdova D, Vlaev L (2012) Non-isothermal kinetics of Lua AC, Yang T (2004) Effects of vacuum pyrolysis conditions on the
thermal degradation of chitosan. Chem Cent J 6:81. https://doi.org/ characteristics of activated carbons derived from pistachio-nut
10.1186/1752-153X-6-81 shells. J Colloid Interface Sci 276:364–372. https://doi.org/10.
Gupta A, Thengane SK, Mahajani S (2018) CO2 gasification of char from 1016/J.JCIS.2004.03.071
lignocellulosic garden waste: experimental and kinetic study. Marinović-Cincović M, Janković B, Jovanović V, Samaržija-Jovanović
Bioresour Technol 263:180–191. https://doi.org/10.1016/J. S, Marković G (2013) Kinetic and thermodynamic analyses of non-
BIORTECH.2018.04.097 isothermal degradation process of acrylonitrile–butadiene and ethyl-
Hameed BH, Daud FBM (2008) Adsorption studies of basic dye on ene–propylene–diene rubbers. Compos Part B 45:321–332. https://
activated carbon derived from agricultural waste: Hevea brasiliensis doi.org/10.1016/j.compositesb.2012.08.006
seed coat. Chem Eng J 139:48–55. https://doi.org/10.1016/J.CEJ. Nizamuddin S, Baloch HA, Griffin GJ, Mubarak NM, Bhutto AW, Abro
2007.07.089 R, Mazari SA, Ali BS (2017) An overview of effect of process
Hardi F, Imai A, Theppitak S, Kirtania K, Furusjö E, Umeki K, parameters on hydrothermal carbonization of biomass. Renew Sust
Yoshikawa K (2018) Gasification of char derived from catalytic Energ Rev 73:1289–1299. https://doi.org/10.1016/J.RSER.2016.12.
hydrothermal liquefaction of pine sawdust under a CO2 atmosphere. 122
Energy Fuel 32:5999–6007. https://doi.org/10.1021/acs. Onoji SE, Iyuke SE, Igbafe AI, Nkazi DB (2016) Rubber seed oil: a
energyfuels.8b00589 potential renewable source of biodiesel for sustainable development
Hassan SNAM, Ishak MAM, Ismail K, Ali SN, Yusop MF (2014) in sub-Saharan Africa. Energy Convers Manag 110:125–134.
Comparison study of rubber seed shell and kernel (Hevea https://doi.org/10.1016/J.ENCONMAN.2015.12.002
brasiliensis) as raw material for bio-oil production. Energy Onoji SE, Iyuke SE, Igbafe AI, Daramola MO (2017) Transesterification
Procedia 52:610–617. https://doi.org/10.1016/J.EGYPRO.2014.07. of rubber seed oil to biodiesel over a calcined waste rubber seed shell
116 catalyst: modeling and optimization of process variables. Energy
Fuel 31:6109–6119. https://doi.org/10.1021/acs.energyfuels.
Hill SJ, Grigsby WJ, Hall PW (2013) Chemical and cellulose crystallite
7b00331
changes in Pinus radiata during torrefaction. Biomass Bioenergy
56:92–98. https://doi.org/10.1016/j.biombioe.2013.04.025 Perander M, DeMartini N, Brink A, Kramb J, Karlström O, Hemming J,
Moilanen A, Konttinen J, Hupa M (2015) Catalytic effect of Ca and
Hoekman SK, Broch A, Robbins C (2011) Hydrothermal carbonization
K on CO2 gasification of spruce wood char. Fuel 150:464–472.
(HTC) of lignocellulosic biomass. Energy Fuel 25:1802–1810.
https://doi.org/10.1016/J.FUEL.2015.02.062
https://doi.org/10.1021/ef101745n
Posmanik R, Martinez CM, Cantero-Tubilla B, Cantero DA, Sills DL,
Huang Y, Yin X, Wang C et al (2009) Effects of metal catalysts on CO2
Cocero MJ, Tester JW (2018) Acid and alkali catalyzed hydrother-
gasification reactivity of biomass char. Biotechnol Adv 27:568–572.
mal liquefaction of dairy manure digestate and food waste. ACS
https://doi.org/10.1016/J.BIOTECHADV.2009.04.013
Sustain Chem Eng 6:2724–2732. https://doi.org/10.1021/
Kruse A, Funke A, Titirici M-M (2013) Hydrothermal conversion of acssuschemeng.7b04359
biomass to fuels and energetic materials. Curr Opin Chem Biol 17:
Reshad AS, Tiwari P, Goud VV (2018) Thermo-chemical conversion of
515–521. https://doi.org/10.1016/J.CBPA.2013.05.004
waste rubber seed shell to produce fuel and value-added chemicals. J
Lahijani P, Zainal ZA, Mohamed AR (2012) Catalytic effect of iron Energy Inst 91:940–950. https://doi.org/10.1016/J.JOEI.2017.09.
species on CO2 gasification reactivity of oil palm shell char. 002
Thermochim Acta 546:24–31. https://doi.org/10.1016/J.TCA. Stirling RJ, Snape CE, Meredith W (2018) The impact of hydrothermal
2012.07.023 carbonisation on the char reactivity of biomass. Fuel Process
Lahijani P, Zainal ZA, Mohamed AR, Mohammadi M (2013) CO2 gas- Technol 177:152–158. https://doi.org/10.1016/J.FUPROC.2018.
ification reactivity of biomass char: catalytic influence of alkali, 04.023
alkaline earth and transition metal salts. Bioresour Technol 144: Ulbrich M, Preßl D, Fendt S, Gaderer M, Spliethoff H (2017) Impact of
288–295. https://doi.org/10.1016/j.biortech.2013.06.059 HTC reaction conditions on the hydrochar properties and CO2 gas-
Lahijani P, Zainal ZA, Mohammadi M, Mohamed AR (2015) Conversion ification properties of spent grains. Fuel Process Technol 167:663–
of the greenhouse gas CO2 to the fuel gas CO via the Boudouard 669. https://doi.org/10.1016/J.FUPROC.2017.08.010
reaction: a review. Renew Sust Energ Rev 41:615–632. https://doi. Xu K, Hu S, Su S, Xu C, Sun L, Shuai C, Jiang L, Xiang J (2013) Study
org/10.1016/J.RSER.2014.08.034 on char surface active sites and their relationship to gasification
Environ Sci Pollut Res

reactivity. Energy Fuel 27:118–125. https://doi.org/10.1021/ through wet torrefaction. Fuel 146:88–94. https://doi.org/10.1016/
ef301455x J.FUEL.2015.01.005
Yan W, Perez S, Sheng K (2017) Upgrading fuel quality of moso bamboo Yuan X, He T, Cao H, Yuan Q (2017) Cattle manure pyrolysis process:
via low temperature thermochemical treatments: dry torrefaction kinetic and thermodynamic analysis with isoconversional methods.
and hydrothermal carbonization. Fuel 196:473–480. https://doi. Renew Energy 107:489–496. https://doi.org/10.1016/J.RENENE.
org/10.1016/J.FUEL.2017.02.015 2017.02.026
Yang W, Shimanouchi T, Iwamura M, Takahashi Y, Mano R, Takashima
K, Tanifuji T, Kimura Y (2015) Elevating the fuel properties of
Humulus lupulus, Plumeria alba and Calophyllum inophyllum L.

You might also like