You are on page 1of 9

Biomass Conv. Bioref.

DOI 10.1007/s13399-014-0137-3

ORIGINAL ARTICLE

Hydrothermal carbonization of various lignocellulosic biomass


Joan G. Lynam & M. Toufiq Reza & Wei Yan &
Victor R. Vásquez & Charles J. Coronella

Received: 18 March 2014 / Revised: 19 June 2014 / Accepted: 23 June 2014


# Springer-Verlag Berlin Heidelberg 2014

Abstract As a non-food, renewable energy resource, ligno- Keywords Wet torrefaction . Agricultural residues .
cellulosic biomass can be used for reducing greenhouse gas Subcritical water . Forest products . Pretreatment . Hydrochar
emissions. Hydrothermal carbonization (HTC) is a process to
prepare the lignocellulosic biomass for subsequent thermo-
chemical conversion. Biomass is reacted in hot compressed 1 Introduction
water at temperatures between 200 and 275 °C and at pres-
sures sufficient to maintain liquid water. HTC was performed Biomass has the potential to reduce the world’s dependence
on corn stover, rice hulls, Tahoe mix (Jeffrey pine and white on fossil fuels. These are by their nature limited and often
fir), switch grass, and Loblolly pine. These biomass sources difficult to access without risking environmental damage. The
encompass a variety of types, including a primary agricultural consumption of fossil fuels increases the amount of CO2 in the
residue, a secondary agricultural residue, a forest product atmosphere, contributing to global warming. As a renewable
removed to reduce wildfires, a primary energy crop, and a energy source, biomass uses as much CO2 in its growth as it
primary forest crop. HTC reaction temperature was found to releases as a fuel. It is considered to be nearly carbon neutral,
be the process variable with the greatest effect. For the five with minimal contributions to global warming. In the USA
biomass studied, hydrochar mass yields were found to de- alone, over 97 million dry tonne of biomass per year is
crease with increasing reaction temperature, while higher estimated available from forested areas, while 162 million
heating values increased with increasing reaction temperature. dry tons of agricultural residue and waste per year could be
Differences in these results among biomass sources were supplied by agricultural lands [1]. Although ethanol can be
found to correlate with their original hemicellulose, lignin, made from corn and other renewable food crops, many believe
aqueous soluble, and ash contents. These components ap- it is unethical to use food sources for the production of liquid
peared to be observable in scanning electron microscope transportation fuels. Lignocellulosic biomass is not only a
images of hydrochar. non-food resource, but can even be a waste product, as in
the case of rice hulls or corn stover.
Electronic supplementary material The online version of this article After mechanical size reduction, lignocellulosic biomass
(doi:10.1007/s13399-014-0137-3) contains supplementary material, can be used for direct firing or cofiring with coal for power
which is available to authorized users. production [2]. Liquid biofuels can be obtained from ligno-
J. G. Lynam : V. R. Vásquez : C. J. Coronella (*) cellulosic biomass by either biochemical or thermochemical
Chemical & Materials Engineering Department, University of conversion routes [3]. Unfortunately, diverse biomass feed-
Nevada, 1664 N. Virginia St. MS 170, Reno, NV 89557, USA stocks exhibit diverse handling characteristics, complicating
e-mail: coronella@unr.edu
their general use as fuel sources. This challenge is further
M. T. Reza compounded by the expensive logistics of seasonal availabil-
APECS Group, Leibniz Institute for Agricultural Engineering, ity in the case of agricultural wastes or wide distribution in the
Max-Eyth-Alee-100, 14469 Potsdam, Germany case of forestry. There is a need to develop pretreatment
processes that produce a uniform solid fuel from lignocellu-
W. Yan
Gas Technology Institute, Energy Conversion Division, Des Plaines, losic biomass with increased higher heating value (HHV),
IL 60018, USA increased hydrophobicity for better storage, and
Biomass Conv. Bioref.

homogeneous handling characteristics. Increased HHV means 2 Experimental


that less biomass can be transported to generate the same
amount of power. Hydrothermal carbonization (HTC), also 2.1 Materials
known as wet torrefaction, or dry torrefaction can be used for
this purpose [4]. The objective of this study is to characterize Biomass feedstocks acquired for testing include Loblolly pine
the hydrochar produced by HTC, looking at different types of (Alabama, USA), pelletized corn stover (Pennsylvania, USA),
biomass, and to investigate the effect of process conditions on Tahoe mix (a mixture of Jeffrey pine and white fir from the
hydrochar properties. Tahoe forest), switch grass (Nevada, USA), and rice hulls
An alternative, dry torrefaction (mild pyrolysis) is typically (California, USA). HHVs are 19.5, 15.6, 20.3, 15.3, and
performed in an inert atmosphere at 0.1 MPa between 225 and 20.3 MJ/kg for loblolly pine, corn stover, Tahoe mix, switch
300 °C using reaction times which vary between 30 min and grass, and rice hulls, respectively.
several hours [4, 5]. Dry torrefaction produces a lower HHV
solid product compared to HTC when performed at similar 2.2 Hydrothermal carbonization
temperatures and for similar reaction times [4]. Another dis-
advantage of dry torrefaction is the energy needed to dry the HTC of lignocellulosic biomass was performed in a 100 mL
feedstock prior to reaction, considered to outweigh the energy Parr bench-top reactor (Moline, IL) at temperatures ranging
released by dry torrefaction if the feedstock has more than from 200 to 280 °C. The temperature of the reactor was
30 % water [6]. Since HTC’s solid product is more hydropho- controlled using a Single Display Proportional-Integral-
bic, less energy is required to dry it than to dry raw biomass Derivative (PID) Controller (Winona, MN). The reactor pres-
[7]. However, advantages of dry torrefaction include low sure was autogenic and was indicated by a pressure gauge
pressure operation and the fact that it requires no water ranging from 1.4 to 6.9 MPa. In other words, the reactor
recycling system [5]. pressure was approximately that of the vapor pressure of water
Hydrothermal carbonization (HTC) is also known as hy- under the same conditions. For each run, a mixture of biomass
drothermal pretreatment or wet torrefaction. HTC is usually feedstock and water was loaded into the reaction vessel. The
performed at temperatures higher than 180 °C, at pressures water to biomass mass ratio was either 5:1 or 10:1. Prior to the
high enough to ensure liquid water, and under an inert atmo- reaction, the biomass was stirred manually to ensure complete
sphere, which prevents wet oxidation of lignin [8]. The HTC wetting. Nitrogen was passed through the reactor for 10 min
reaction time has been reported to be 1 min to several hours, prior to each run to purge oxygen from the reactor. Heating of
although most of the reactions that remove mass from the the reactor required 15–30 min, after which the reaction was
biomass seem to occur within the first 20 min after bringing maintained at the desired temperature for either 5 or 20 min.
the reactor to temperature [9–12]. Additional reactions in the Next, the reactor was cooled at room temperature rapidly by
liquid phase occur subsequently [8]. Additives, such as acids immersion into an ice bath. The gaseous product was released
or bases, can affect the formed products [9]. Adding acid to without further evaluation. The liquid and solid products were
the reactants can increase the HHV of the solid product, as separated by vacuum filtration using a 0.6 μm Whatman filter.
well as reduce the mass yield [13]. Water, acting as both All reactions were performed at least twice, and the average
reactant and solvent, shows different physical and chemical reported. Mass yield measurements had uncertainties of ±5 %.
properties depending on the operating conditions [14]. At
temperatures between 227 and 327 °C, water may act as both 2.3 Heat of combustion
a base and an acid because its ionic product is maximized. In
addition, water’s dielectric constant is decreased at these tem- Heats of combustion of solid samples were measured in a Parr
peratures so that it acts more like a non-polar solvent [15]. 1241 adiabatic oxygen bomb calorimeter (Moline, IL). Solid
During HTC, hemicelluloses and cellulose are hydrolyzed to samples (0.5–1.0 g) were dried at 105 °C for 24 h prior to
oligomers and monomers, while lignin is mostly unaffected at analysis. The uncertainty of measurements is ±0.5 MJ/kg.
temperatures below 275 °C [9, 16–20]. The solid product,
hydrochar, which is also known as biochar or biocoal, has 2.4 Fiber analysis
reduced equilibrium moisture content, which favors long-term
storage [7]. The hydrochar is quite friable and is readily made The van Soest method of NDF-ADF-ADL (neutral detergent
into durable pellets to facilitate handling [21]. fiber, acid detergent fiber, and acid detergent lignin) dissolu-
The present work investigates the importance of tempera- tion was used to determine the percentage of aqueous solu-
ture, time, water to biomass ratio, and biomass particle size on bles, hemicellulose, cellulose, and lignin components in bio-
one representative biomass. The effect of HTC on mass yield mass samples [22]. In untreated biomass, hemicellulose is the
and increase in fuel value for four other biomass is also fraction present in NDF but not in ADF, while cellulose is the
studied. fraction present in ADF but not in ADL. Mass removed by
Biomass Conv. Bioref.

neutral detergent is considered to be soluble in water (aqueous about which variables or interaction of variables produce the
solubles). Klason lignin content was found by subtracting ash largest effect with the least number of experiments.
content from the ADL. Ash was determined by (NREL/TP-
510-42622) [23]. Fiber analysis measurements made with this
method are generally accurate to ±5 % [13]. Application of
this fiber analysis technique to HTC-treated lignocellulosic 3 Results and discussion
biomass is problematic, since the solid products of HTC likely
contains fractions with solubility properties distinct from 3.1 Effects of reaction temperature, holding time, water
those present in untreated biomass, as discussed further in to biomass ratio, and particle size
Sections 3.2 and 3.3.
Important dependent variables for HTC include mass yield,
2.5 Scanning electron microscopy energy densification, and energy yield. Mass yield is defined
as the mass ratio of dried hydrochar to dried reactant biomass.
An FE-SEM Hitachi scanning electron microscope (SEM) Energy densification is the ratio of the higher heating value
model S-4700 was used to visualize pretreated loblolly pine. (HHV) of the dried hydrochar to that of the original dried
Loblolly pine samples (0.5-mm mesh size) either raw or biomass and are both reported on an ash-free basis. The
pretreated at 200, 230, and 260 °C were used for SEM images. energy yield is defined as the mass yield times the energy
The samples were maintained on special studs and platinum densification and describes how the total fuel value of the
coated with polaran coater tar 5,000, under an argon atmo- hydrochar relates to the fuel value of the original biomass.
sphere for a coating thickness of approximately 1,000 Å. The Reaction temperature (T), HTC holding time (t), water to
samples were dried in a 105 °C oven for 24 h prior to analysis. biomass ratio (R), and biomass particle size (S) are considered
A voltage of 15–20 kV with a magnification of 250–1,000 to be the important process variables for HTC [25]. Loblolly
times was used for the images pine was selected as a representative lignocellulosic biomass
to investigate these parameters. In any reaction, temperature
and time influence the products generated. Initial concentra-
2.6 Proximate analysis tion of the biomass reactant (the inverse of the water to
biomass ratio) has been reported to be a significant parameter
Hydrochar samples were analyzed with a Perkin Elmer TGA- for HTC [26]. The mesh size of the loblolly pine was also
7 Thermogravimetric Analyzer (Waltham, MA) to determine varied to investigate the possible mass transfer limitations.
the contents of moisture, volatile matter, and fixed carbon Table 1 indicates a trend toward lower mass yield with in-
[24]. Thermogravimetric analysis was carried out under a creasing holding time. Mass yield was only 42.0 % at the
nitrogen purge at a constant rate of 50 mL min−1 to prevent higher reaction temperature with lower biomass concentration
oxidation of samples. Samples were first heated up to 105 °C (higher water to biomass ratio). It is possible that reduced
at the rate of 10 °C min−1, maintained at 105 °C for 10 min, particle size also reduces mass yield, which would be expect-
then rapidly increased to 900 °C at the rate of 50 °C min−1, and ed if any mass transfer limitations occur. However, the most
held at that temperature for 60 min. Mass evaporating at important product changes occur with temperature change. A
105 °C was considered to be moisture. Mass evaporating at reaction temperature increase from 200 to 260 °C causes a
temperatures between 105 and 900 °C was referred to as large decrease in mass yield and a large increase in energy
volatiles. All mass remaining after heating to 900 °C consisted densification (Table 1). For example, mass yield for the runs at
of fixed carbon and ash [24]. 200 °C and 5 min (Table 1) is near 86.1 %, while mass yields
for runs at 260 °C and 5 min are near 55 %, which is a
2.7 Experimental design substantial decrease with increasing temperature. Energy yield
also decreases with increasing reaction temperature; it goes
The following variables were initially investigated using lob- from near 95 % to close to 70 % for the same runs. Since
lolly pine as a representative biomass: HTC temperature (T), energy yield is the product of mass yield and energy densifi-
time of HTC reaction at temperature (t), mass ratio of water to cation, this decrease is due to the large decrease in mass yield.
loblolly pine (R), and mesh size of biomass (S). Each variable Energy densifications for all the 200 °C runs were near 1.1,
was set at either a low or high level, as indicated in Table 1. while the 260 °C runs had energy densifications near 1.4. This
Experimental order was randomized with duplicates of each means that the higher reaction temperature runs (260 °C)
condition and the average reported. A half factorial design was produced hydrochar with fuel values 40 % higher than that
used to provide level IV resolution, such that no main effect of raw loblolly pine, while the lower reaction temperature runs
was aliased with any other main effect or two-way interaction. (200 °C) gave hydrochar with fuel values 10 % higher than the
This type of experimental design gives valuable information raw loblolly pine. A statistical analysis (ANOVA) of the
Biomass Conv. Bioref.

Table 1 Effect of HTC reaction temperature (T), holding time (t), water biomass ratio (R), and particle size (S) on mass yield, energy densification, and
energy yield for Loblolly pine

T (°C) t (min) R (g water/g pine) S(mm) Mass yield (%) Energy densification Energy yield (%)

200 5 5 0.16 86.1 1.11 95.5


200 5 10 0.24 87.2 1.07 93.5
200 20 5 0.24 85.7 1.15 98.4
200 20 10 0.16 79.0 1.12 88.1
260 5 5 0.24 57.0 1.44 81.8
260 5 10 0.16 53.6 1.24 66.3
260 20 5 0.16 50.0 1.42 71.0
260 20 10 0.24 42.0 1.48 61.9

results shows that at 95 % confidence, only reaction temper- temperature for all biomass. Such a decrease would be ex-
ature significantly affects the mass yield and energy densifi- pected, since the rate of mass loss during HTC follows an
cation. Other variables (holding time, water to biomass ratio, Arrhenius form [20].
and biomass feedstock size) were shown to have statistically A closer look at the makeup of the biomass can help to
insignificant effects on the measured variables. The statistical understand these results. Lignocellulosic biomass is consid-
analysis also shows that the interactions among the four ered here to consist of five fractions: hemicelluloses, cellulose,
variables were not significant. Therefore, manipulation of lignin, aqueous solubles, and ash. Table 2a shows fiber anal-
these variables can be done independently within the variable ysis data for each of these raw biomass sources. Biomass with
range of the experiments. greater content of hemicellulose and aqueous solubles results
The literature shows temperature to be a significant factor in lower mass yield of hydrochar after HTC, particularly at
in HTC, with other process variables less consequential. The 260 °C. For example, switch grass and corn stover have more
influence of holding time and water to biomass ratio varies hemicellulose and aqueous solubles compared to the other
with biomass feedstock. Gao et al. (2013) found times less biomass tested and exhibited a lower mass yield at 260 °C.
than 4 h to give similar hydrochar properties when water The softwood biomass, Tahoe mix and Loblolly pine, have
hyacinth was the feedstock, and Heilmann et al. (2010) found lower amounts of hemicellulose and aqueous solubles, and
time was not a significant variable when HTC was applied to exhibit higher mass yield at 260 °C. These results suggest that
microalgae [27, 28]. However, when applied to distiller’s hemicellulose and aqueous solubles react easily in the hydro-
grain as a feedstock, Heilmann et al. (2011) reported that time thermal carbonization (HTC) process [4]. Other published
was a significant factor, with increased carbonization when work supports this conclusion [31–34]. Reza et al. (2013)
time was increased from 0.5 to 2 h [29]. For microalgae and found the rate constant for hemicellulose hydrolysis to be
distiller’s grain, higher water to biomass ratio (19:1) gave 4.5 times that of cellulose at 260 °C for HTC and claimed
lower mass yield than a lower one (3:1) [28, 29]. Rogalinski that the aqueous solubles reaction rate is nearly instantaneous
et al. (2008), when using rye straw, found particle size and [20].
biomass concentration to have no effect on mass yield [30]. Ash tends to be inert in the HTC process, particularly at
Clearly, biomass type plays an important role in determining lower temperatures. For this reason, biomass with higher ash
the effects of process variables. Optimization for each feed- content, such as corn stover and rice hulls, shows higher mass
stock needs to be considered independently. yields for a 200 °C HTC reaction temperature. However,
when HTC causes biomass removal from the lignocellulosic
3.2 Change in biomass and temperature effects on mass yield, matrix, the ash incorporated in the matrix may be released
energy densification, and energy yield simultaneously [35]. This would explain a decreased mass
yield for corn stover and switch grass at higher HTC reaction
Due to the significant effect of reaction temperature, hydro- temperatures. Rice hulls, with their peculiar silicon coating
thermal carbonization (HTC) of various types of biomass was and extraordinarily high ash content, retain ash during HTC,
performed at temperatures of 200, 230, and 260 °C, using thus maintaining a higher mass yield than the other grass-like
water to biomass ratio of 5:1 and a holding time of 5 min. biomass, despite their relatively high aqueous solubles content
Figure 1 shows the mass yield, energy densification, and [36].
energy yield results for rice hulls, corn stover, Tahoe mix, It is possible to determine a fiber analysis on the HTC
switch grass, and Loblolly pine. Mass yield ranged from 30 to biochar, but the results must be interpreted with care, as the
94 % and definitely decreases with increasing HTC reaction components in the char may have quite different solubility and
Biomass Conv. Bioref.

Fig. 1 Effect of temperature on a


a) mass yield (%), b) energy 95
densification, and c) energy yield
(%) for rice hulls, pelletized corn 85
stover, Tahoe mix (a mixture of 75
Jeffrey pine and white fir from the Mass rice hulls
65
Tahoe forest), switch grass, and Yield corn stover
Loblolly pine (%) 55 Tahoe mix
45 switch grass
35 Loblolly pine
25
180 200 220 240 260 280
HTC Temperature (°C)
b 1.45
1.40
1.35
1.30
Energy 1.25 rice hulls
Densi- corn stover
1.20
ficaon Tahoe mix
1.15
1.10 switch grass
1.05 Loblolly Pine
1.00
180 200 220 240 260 280
HTC Temperature (°C)

c
95

85

Energy 75 rice hulls


Yield 65 corn stover
(%) Tahoe mix
55
switch grass
45
Loblolly pine
35
180 200 220 240 260 280
HTC Temperature (°C)

Table 2 Fiber analysis of (a) raw biomasses and (b) Loblolly pine raw and HTC pretreated at 200 °C and 260 °C

Hemicellulose (%) Cellulose (%) Lignin (%) Aqueous solubles (%) Ash (%)

(a) Biomass
Rice hulls 14.9 39.8 11.3 12.9 21.1
Corn stover 26.3 29.7 11.3 26.3 6.5
Tahoe mix 18.2 50.7 23.8 6.1 1.3
Switch Grass 28.1 37.1 5.6 21.8 7.4
Loblolly pine 11.5 55.4 30.0 2.7 0.4
(b) Loblolly pine
Raw 11.5 2.7 0.4
200 °C HTC 1.1 15.2 0.5
260 °C HTC 0.7 18.1 0.8
Biomass Conv. Bioref.

reactivity characteristics than native biomass. The content of reaction scheme, to fructose, which dehydrates to 5-
hemicellulose is determined by the difference in NDF and hydroxymethylfurfural (5-HMF) [42]. This 5-HMF may pre-
ADF. The presence of compounds in the NDF not in the ADF cipitate into the pores of the solid product and then be
could include compounds other than hemicellulose. However, accounted for as aqueous solubles in the dried biochar [15,
it is safe to conclude that when the difference between the 37]. Compared to cellulose or glucose, 5-HMF has a greater
NDF and the ADF in a treated biochar is less than that in the HHV, and therefore its increased presence may also partially
raw biomass, then the hemicellulose is no longer present in the account for the greater energy densification found in Loblolly
solid. The results of this limited fiber analysis are shown in pine [43].
Table 2b. The experimental conditions of the treatment at Both Loblolly pine and Tahoe mix are softwoods. The
200 °C used a water to biomass ratio of 5:1, a holding time Tahoe mix is a mix of two species and is made from whole
of 5 min, and a particle size of 0.16 mm, and those for the trees, while the Loblolly pine, a commercial timber, is
260 °C run were the same except that particle size was debarked and relatively clean. Both biomass follow similar
0.24 mm. The mass yield for 200 °C was 86.1 % and for trends in HTC, except for the energy densification, shown in
260 °C was 57.0 %, indicating that the higher temperature Fig. 1b, where we see a much greater increase for Loblolly,
removed more of the biomass material from the hydrochar. At compared to the Tahoe mix. An explanation for this observa-
200 and 260 °C, most of the loblolly pine’s hemicellulose is tion is not readily apparent, but there are several possibilities.
apparently extracted, and likely hydrolyzed to monosaccha- Just from the fiber analysis, we find that the Tahoe mix has a
rides, such as xylose, arabinose, galactose, and mannose [34, reactive index value of 0.97, twice that of the loblolly (0.47),
37]. A preliminary analysis of the liquid products by HPLC, as discussed below. The difference in ash composition may
not shown here, showed the presence of various sugars (xy- catalyze or inhibit some reactions. The composition of the
lose, glucose, arabinose, galactose, and mannose), consistent soluble fraction is likely to be quite different between the two,
with other investigations [37–39]. While soluble in water, and the characteristics of the cellulose fraction are likely to be
these sugars are likely to precipitate into the pores of the quite different. Further study of the differences in aqueous
hydrochar, accounting for increased aqueous solubles found solubles, as well as lignin and cellulose, of softwood biomass
in the hydrochars, as shown in Table 2b. could prove useful in determining factors that influence ener-
For all the biomass investigated, energy densification in- gy densification.
creased with HTC reaction temperature, as shown in Fig. 1b. Switch grass and corn stover have more hemicellulose
Hemicellulose and cellulose have a lower HHV compared to compared to the other biomass tested. Hemicellulose is bond-
lignin [40]. When they are removed, a higher proportion of ed to cellulose by lignin “glue”, shielding and preventing it
lignin must remain, causing an increased HHV. In addition, from decomposing [44]. When the hemicellulose removed is a
cellulose that undergoes HTC may form oligomers, which large portion of the biomass structure, the cellulose is then
polymerize into a lignin-like substance [41]. This may explain more accessible for reactions, such as solubilization, hydroly-
why Loblolly pine, which has both the highest cellulose sis, and polymerization [30; 34]. As hemicellulose and some
(55.4 %) and the highest lignin (30.0 %) concentrations, has of the cellulose are solubilized to monomeric forms, a higher
the highest energy densifications (up to 1.43) at all HTC proportion of lignin must remain in the biochar. A portion of
temperatures for the biomass investigated (Fig. 1b). the cellulose may also polymerize under these conditions,
After cellulose hydrolysis to glucose, glucose can isomer- producing higher molecular weight products. This may ac-
ize in the presence of acid, which is produced in the count for the increased HHV and thus the increased energy

Fig. 2 Mass yield versus 100


reactivity index (ratio of reactive 90
to inert biomass components) for
the five biomass types 80
70
60
Mass 200 °C
50
yield (%)
40 230 °C
30
260 °C
20
10
0
0 1 2 3 4 5
Reacvity index
Biomass Conv. Bioref.

densification that switch grass and corn stover show (Fig. 1b). yield. At lower HTC reaction temperatures, energy yields
Rice hulls, which have intermediate aqueous solubles, hemi- were similar for the different biomass types. Thus, a higher
cellulose, cellulose, and lignin values, were found to have HTC reaction temperature may be optimal for softwood bio-
intermediate energy densification. mass, but a lower one may be preferable for grass-type
In light of the results discussed above and presented in biomass.
Table 2 and in Fig. 1, we have attempted to identify an index
of reactivity to predict the mass yield in HTC reactions. The 3.3 Increased aqueous solubles and friability
most successful index identified from this work is one defined
as the ratio of “reactive components” in the raw biomass: Increased pore area likely occurred due to the removal of
cellulose and hemicellulose in the solid product [45]. Since
Reactive Components H 0 þ S0 more surface area is available for precipitation, more sugars
IR ¼ ¼
Inert components L0 þ A0 and other chemicals can adhere to the accessible pores. Mono-
mers of cellulose and hemicellulose and organic acids are
Here, H0, S0, L0, and A0 refer to fiber fractions of hemicel- produced by the HTC reactions [37]. These non-volatiles
lulose, solubles, lignin, and ash, as reported in Table 2. By this can adhere to the surface area in the pretreated biomass unless
index, the biomass tested, in order of reactivity are switch extensive washing is performed. Likely for these reasons,
grass > corn stover > tahoe mix > rice hulls > Loblolly pine, aqueous solubles increased with higher HTC temperature, as
Fig. 2 shows the mass yield as a function of this reactivity shown in Table 2b. The mass fraction of aqueous solubles
index at three temperatures, along with a line fitted to the data compounds increases with extent of reaction, from 2.7 % for
taken at 260 °C. The dependence on this index of reactivity is raw biomass, to 15.3 % for biomass pretreated at 200 °C, to
apparent and becomes obvious with increasing temperature. 18.1 % for biomass pretreated at 260 °C. When the solid is
Energy yield is an important practical consideration for dried at 105 °C, water and volatiles evaporate. Non-volatile
conversion of biomass to hydrochar by HTC. As shown in substances precipitate on the extensive biomass surface area,
Fig. 1c, at 260 °C, softwood biomass gave a higher energy rather than on the container surfaces. These precipitates are
yield hydrochar than the grass-type biomass, except for rice part of the hydrochar, but are quantified as aqueous solubles
hulls, which have a high ash content that increases their mass by fiber analysis, since they are soluble in water. With vigor-
ous washing, these compounds might be more correctly char-
acterized as dissolved chemicals, including sugars. The empty
spaces left in the solid product are also likely the cause of the
higher friability observed in the hydrochar.

3.4 Scanning electron microscope images

Figure 3 shows SEM images of raw Loblolly pine and Lob-


lolly pine hydrochar from a 200 °C HTC reaction temperature
using a water to biomass ratio of 5:1 and a holding time of
5 min. In general, cellulose can be described as colinear
crystalline bundles of microfibrils in the raw lignocellulosic
biomass, while hemicellulose covers the cellulose, and lignin
is the strata between microfibrils that also connect the hemi-
cellulose and microfibrils to each other [44]. Aqueous solu-
bles can be found as a thin layer over the whole configuration.
Figure 3a is a ×250 magnification of the raw Loblolly pine.

Table 3 Proximate analysis of Loblolly pine raw and HTC pretreated at


200 and 260 °C

Loblolly pine Moisture (%) Volatiles (%) Fixed carbon (%)

Raw 3.6 83.7 12.3


Fig. 3 Scanning electron microscope images of raw Loblolly pine and 200 °C HTC 1.3 86.0 12.2
loblolly pine HTC pretreated at 200 °C. a raw Loblolly pine at 250×, b 260 °C HTC 1.3 72.2 25.7
200 °C hydrochar at 250×
Biomass Conv. Bioref.

The lignocellulosic structure is not clear. The lignocellulosic content decreased and fixed carbon content increased com-
biomass shown likely has starches, proteins, and other soluble pared to 200 °C HTC hydrochar. Oxygenated compounds in
compounds filling the spaces between microfibrils. In the biomass, including polysaccharides, are preferentially degrad-
image of biochar pretreated at 200 °C in Fig. 3b, it is possible ed during HTC, and the HTC char has a higher elemental
that biopolymers such as lignin wrapping microfibril cellulose carbon content than the originating biomass [6, 34, 37]. Dur-
bundles are observable. This idea suggests that at 200 °C all ing HTC, it is known that cellulose fragments undergo poly-
the aqueous solubles are completely removed and that much merization and these carbonaceous polymers are expected to
of the hemicellulose is decomposed, so that lignin is detached have relatively low volatiles content due to cross linking [34;
from cellulose microfibrils. Kobayashi et al. (2009) reported 37]. Thus, the increased fixed carbon in biochar produced at
that below 200 °C fibrous materials are observed, which temperatures that cause cellulose to react is expected.
suggests that the decomposition of woody biomass begins
above 200 °C [46]. Additional SEM images of hydrochar
pretreated at 230 and 260 °C are available in Online Resource
1 (“230 °C and 260 °C Hydrochar SEM images”).
4 Conclusion
3.5 Recycling of liquid product
HTC is a method for producing a hydrophobic, friable solid
fuel for subsequent thermochemical conversion from diverse
Reusing the liquid product of HTC to treat fresh biomass
lignocellulosic biomass feedstocks, without the need for initial
would reduce the quantity of wastewater needing treatment
drying. Energy densification of biomass was found to increase
in an industrial plant. Replicated experiments using recycled
by up to 43 %. Reaction temperature was the main process
liquid product were performed with Loblolly pine, one at
variable significantly affecting mass yield and energy densi-
200 °C and the other at 260 °C. At 200 °C, the mass yield
fication. For an HTC reaction temperature of 260 °C, the solid
and energy densification were relatively constant when the
product hydrochar had increased fixed carbon and decreased
liquid product was reused instead of pure water, but at 260 °C,
volatiles. Fiber analysis suggested that hemicellulose hydro-
there were some distinct changes. The mass yield increased
lyzes readily at temperatures between 200 and 260 °C. When
from 57.0 to 65.6 %, and the energy densification ratio de-
subjected to HTC at 260 °C, grass-like biomass feedstocks of
creased from 1.44 to 1.31. Uddin et al. (2013) reported similar
higher hemicellulose content produced low mass yield, while
results for mass yield and energy densification for hydrochar
softwood biomass of low hemicellulose content produced
when liquid product was recycled in HTC pretreatment [47].
high mass yield, so that softwood biomass hydrochar pro-
When Stemann et al. (2013) performed HTC at 220 °C for 4 h
duced a higher energy yield. A reactivity index is defined as
on poplar wood chips and process water was recycled, they
the ratio of relatively reactive compounds (extractives and
reported increased concentration of organic acids leading to
hemicellulose) to relatively inert compounds (lignin and
polymerization reactions and higher energetic yield for the
ash), which was used to correlate HTC reactivity of various
solid product [48].
biomass feedstocks.
For these shorter time and higher reaction temperature
experiments, HTC chemistry is apparently inhibited by com-
pounds produced at 260 °C that are water soluble, possibly Acknowledgments The authors gratefully acknowledge the support by
Department of Energy (Contract Number: DE-FG36-01GO11082). The
because equilibrium reactions occur in the reaction scheme.
authors acknowledge meaningful conversations with project partners,
Lynam et al. (2011) reported similar reductions in acetic acid including Larry Felix of the Gas Technology Institute, and Craig Einfeldt
formation when acetic acid was added to initial reactants in of Changing World Technologies. Experimental work of Mr. Tapas
hydrothermal carbonization (HTC), which is indicative of Acharjee and Mr. Jason Hastings is gratefully acknowledged.
concentration-shifts due to relative amounts of reactants and
products in equilibrium reactions (i.e., Le Chatelier’s
principle) [13]. References

3.6 Proximate analysis


1. Perlack RD, Stokes BJ (2011) U.S. Billion-Ton Update: Biomass
Supply for a Bioenergy and Bioproducts Industry. In: ORNL/TM-
Table 3 shows that the proximate analysis of raw Loblolly 2011/224. (ed). U.S. Department of Energy., Oak Ridge National
pine had similar volatiles, fixed carbon, and ash values as Laboratory, Oak Ridge, TN., pp. 227
those for Loblolly pine undergoing HTC at 200 °C. Moisture 2. Zuwala J, Sciazko M (2010) Full-scale co-firing trial tests of sawdust
and bio-waste in pulverized coal-fired 230 t/h steam boiler. Biomass
content is higher for raw Loblolly pine compared to the Bioenergy 34:1165–1174
hydrochars, suggesting that hydrochar is more hydrophobic 3. Bridgwater AV, Double JM (1991) Production costs of liquid fuels
than its feedstock [7]. For 260 °C HTC hydrochar, volatile from biomass. Fuel 70:1209–1224
Biomass Conv. Bioref.

4. Yan W, Acharjee TC, Coronella CJ, Vasquez VR (2009) Thermal 27. Gao Y, Wang XH, Wang J, Li XP, Cheng JJ, Yang HP, Chen HP
pretreatment of lignocellulosic biomass. Environ Prog Sustain (2013) Effect of residence time on chemical and structural properties
Energy 28:435–440 of hydrochar obtained by hydrothermal carbonization of water hya-
5. Sadaka S, Negi S (2009) Improvements of biomass physical and cinth. Energy 58:376–383
thermochemical characteristics via torrefaction process. Environ 28. Heilmann SM, Davis HT, Jader LR, Lefebvre PA, Sadowsky MJ,
Prog Sustain Energy 28:427–434 Schendel FJ, von Keitz MG, Valentas KJ (2010) Hydrothermal
6. Libra JA, Ro KS, Kammann C, Funke A, Berge ND, Neubauer Y, carbonization of microalgae. Biomass Bioenergy 34:875–882
Titirici M-M, Fühner C, Bens O, Kern J, Emmerich K-H (2010) 29. Heilmann SM, Jader LR, Sadowsky MJ, Schendel FJ, von Keitz MG,
Hydrothermal carbonization of biomass residuals: a comparative Valentas KJ (2011) Hydrothermal carbonization of distiller’s grains.
review of the chemistry, processes and applications of wet and dry Biomass Bioenergy 35:2526–2533
pyrolysis. Biogeosciences 2:71–106 30. Rogalinski T, Ingram T, Brunner G (2008) Hydrolysis of lignocellu-
7. Acharjee TC, Coronella CJ, Vasquez VR (2011) Effect of thermal losic biomass in water under elevated temperatures and pressures. J
pretreatment on equilibrium moisture content of lignocellulosic bio- Supercrit Fluids 47:54–63
mass. Bioresour Technol 102:4849–4854 31. Peterson AA, Vogel F, Lachance RP, Froling M, Antal MJ, Tester JW
8. Arvaniti E, Bjerre AB, Schmidt JE (2012) Wet oxidation pretreat- (2008) Thermochemical biofuel production in hydrothermal media:
ment of rape straw for ethanol production. Biomass Bioenergy 39: A review of sub- and supercritical water technologies. Energy
94–105 Environ Sci 1:32–65
9. Funke A, Ziegler F (2010) Hydrothermal carbonization of biomass: a 32. Garrote G, Dominguez H, Parajo JC (1999) Hydrothermal processing
summary and discussion of chemical mechanisms for process engi- of lignocellulosic materials. Holz Als Roh-Werkst 57:191–202
neering. Biofuels Bioprod Bioref 4:160–177 33. Grenman H, Eranen K, Krogell J, Willfor S, Salmi T, Murzin DY
10. Knezevic D, van Swaaij W, Kersten S (2010) Hydrothermal conver- (2011) Kinetics of aqueous extraction of hemicelluloses from spruce
sion of biomass. II. Conversion of wood, pyrolysis oil, and glucose in in an intensified reactor system. Ind Eng Chem Res 50:3818–3828
hot compressed water. Ind Eng Chem Res 49:104–112 34. Reza, M; Uddin, M; Lynam, J; Hoekman, S; Coronella C.(2014):
11. Lu X, Yamauchi K, Phaiboonsilpa N, Saka S (2009) Two-step Hydrothermal carbonization of loblolly pine: reaction chemistry and
hydrolysis of Japanese beech as treated by semi-flow hot- water balance. Biomass Conversion and Biorefinery. 3 (January
compressed water. J Wood Sci 55:367–375 2014): 113–126 Online: http://dx.doi.org/10.1007/s13399-014-
12. Mosier N, Hendrickson R, Ho N, Sedlak M, Ladisch MR (2005) 0115-9
Optimization of pH controlled liquid hot water pretreatment of corn 35. Reza MT, Lynam JG, Uddin MH, Coronella CJ (2013) Hydrothermal
stover. Bioresour Technol 96:1986–1993 carbonization: Fate of inorganics. Biomass Bioenergy 49:86–94
13. Lynam JG, Coronella CJ, Yan W, Reza MT, Vasquez VR (2011) 36. Lynam JG, Reza MT, Vasquez VR, Coronella CJ (2012) Pretreatment
Acetic acid and lithium chloride effects on hydrothermal carboniza- of rice hulls by ionic liquid dissolution. Bioresour Technol 114:629–636
tion of lignocellulosic biomass. Bioresour Technol 102:6192–6199 37. Hoekman SK, Broch A, Robbins C (2011) Hydrothermal carboniza-
14. Ando H, Sakaki T, Kokusho T, Shibata M, Uemura Y, Hatate Y tion (HTC) of lignocellulosic biomass. Energy Fuels 25:1802–1810
(2000) Decomposition behavior of plant biomass in hot-compressed 38. Matsumura Y, Yanachi S, Yoshida T (2006) Glucose decomposition
water. Ind Eng Chem Res 39:3688–3693 kinetics in water at 25 MPa in the temperature range of 448–673 K.
15. Yu Y, Lou X, Wu HW (2008) Some recent advances in hydrolysis of Ind Eng Chem Res 45:1875–1879
biomass in hot-compressed, water and its comparisons with other 39. Sasaki M, Adschiri T, Arai K (2003) Fractionation of sugarcane
hydrolysis methods. Energy Fuels 22:46–60 bagasse by hydrothermal treatment. Bioresour Technol 86:301–304
16. Bobleter O (1994) Hydrothermal degradation of polymers derived 40. Demirbas A (2005) Estimating of structural composition of wood and
from plants. Prog Polym Sci 19:797–841 non-wood biomass samples. Energy Sources 27:761–767
17. Petersen MO, Larsen J, Thomsen MH (2009) Optimization of hy- 41. Yu Y, Wu HW (2010) Evolution of primary liquid products and
drothermal pretreatment of wheat straw for production of bioethanol evidence of in situ structural changes in cellulose with conversion
at low water consumption without addition of chemicals. Biomass during hydrolysis in hot-compressed water. Ind Eng Chem Res 49:
Bioenergy 33:834–840 3919–3925
18. Sun Y, Cheng JY (2002) Hydrolysis of lignocellulosic materials for 42. Huber GW, Iborra S, Corma A (2006) Synthesis of transportation
ethanol production: a review. Bioresour Technol 83:1–11 fuels from biomass: Chemistry, catalysts, and engineering. Chem Rev
19. Zhang B, Huang HJ, Ramaswamy S (2008) Reaction kinetics of the 106:4044–4098
hydrothermal treatment of lignin. Appl Biochem Biotechnol 147: 43. Verevkin SP, Emel’yanenko VN, Stepurko EN, Ralys RV, Zaitsau DH,
119–131 Stark A (2009) Biomass-derived platform chemicals: Thermodynamic
20. Reza MT, Yan W, Uddin MH, Lynam JG, Hoekman SK, Coronella studies on the conversion of 5-hydroxymethylfurfural into bulk inter-
CJ, Vasquez VR (2013) Reaction kinetics of hydrothermal carboni- mediates. Ind Eng Chem Res 48:10087–10093
zation of loblolly pine. Bioresour Technol 139:161–169 44. Salanti A, Zoia L, Tolppa EL, Orlandi M (2012) Chromatographic
21. Reza MT, Lynam JG, Vasquez VR, Coronella CJ (2012) Pelletization detection of lignin-carbohydrate complexes in annual plants by de-
of biochar from hydrothermally carbonized wood. Environ Prog rivatization in ionic liquid. Biomacromolecules 13:445–454
Sustain Energy 31:225–234 45. Hu B, Yu SH, Wang K, Liu L, Xu XW (2008) Functional carbona-
22. Goering HK, Soest PJV (1970) Forage fiber analyses U.S. ceous materials from hydrothermal carbonization of biomass: an
Agricultural Research Service, pp. 1–3 effective chemical process. Dalton Trans.:5414–5423
23. Ehrman T (1994) Standard Method for Ash in Biomass. Golden, CO 46. Kobayashi N, Okada N, Hatano S, Itaya Y, Mori S (2009) Study on
pp. 20 biomass hydrothermal treatment in a continuous slurry-flow type
24. Garcia R, Pizarro C, Lavin AG, Bueno JL (2013) Biomass proximate reactor. Kag Kog Ronbunshu 35:459–464
analysis using thermogravimetry. Bioresour Technol 139:1–4 47. Uddin MH, Reza MT, Lynam JG, Coronella CJ (2013) Effects of
25. Hicks RC, Turner VK (1999) Fundamental concepts in the design of water recycling in hydrothermal carbonization. Environ. Prog.
experiments. Oxford University Press, New York Sustain. Energy
26. Roman S, Nabais JMV, Laginhas C, Ledesma B, Gonzalez JF (2012) 48. Stemann J, Putschew A, Ziegler F (2013) Hydrothermal carboniza-
Hydrothermal carbonization as an effective way of densifying the tion: Process water characterization and effects of water recirculation.
energy content of biomass. Fuel Process Technol 103:78–83 Bioresour Technol 143:139–146

You might also like