Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Biomass Conversion and Biorefinery

https://doi.org/10.1007/s13399-019-00488-0

REVIEW ARTICLE

Understanding the influence of biomass particle size and reaction


medium on the formation pathways of hydrochar
Dominik Wüst 1,2 2 2 3
& Catalina Rodriguez Correa & Dennis Jung & Michael Zimmermann & Andrea Kruse & Luca Fiori
2 1

Received: 15 May 2019 / Revised: 22 July 2019 / Accepted: 24 July 2019


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

Abstract
The chemical-physical processes controlling hydrothermal carbonization (HTC) are still not completely understood. This paper
focuses on two aspects: the influence on the hydrochar formation of the particle size of the feedstock and the presence of solved
compounds in the feedwater. To address these, brewer’s spent grains were crushed to < 1 mm and separated in three fractions. In
addition, residual process water (rPW) from 5-hydroxymethylfurfural (HMF) production instead of bi-distilled H2O was added in
a series of experiments for recycling. The results show a transfer limitation of hydrolysis products through pores for the particle
size fractions > 250 μm proved by HPLC analysis of liquid byproducts, particularly when rPW, containing readily condensable/
polymerizable intermediates, is added. This has a positive effect on the yield and carbon content of the hydrochars caused mainly
by an increase in its secondary char fraction. The reaction pathways involved are discussed in detail.

Keywords Hydrothermal carbonization . Waste prevention . Reaction pathways . Hydrolysis . Hydrochar

Abbreviations PW Process water


HTC Hydrothermal carbonization Cfix Fixed carbon
HC Hydrochar VM Volatile matter
HMF 5-hydroxymethylfurfural LA Levulinic acid
Ceff Carbon retention efficiency FA Formic acid
rPW Residual process water AA Acetic acid
FRU Fructose LaA Lactic acid
BSG Brewer’s spent grains GA Glycolic acid

Highlights
• Detailed understanding of the reaction pathways originated from
oligosaccharides during the hydrothermal carbonization of soft
lignocellulosic biomass
• Control of mass transfer during hydrolysis of oligosaccharides from
small biomass particle sizes through the fast condensation and
polymerization of reactive dehydration products
• Enhancement of yields and carbon retention efficiencies during
hydrothermal carbonization by recirculation of residual process water
from HMF synthesis instead of using water
Electronic supplementary material The online version of this article
(https://doi.org/10.1007/s13399-019-00488-0 ) contains supplementary
material, which is available to authorized users.

* Dominik Wüst 2
Department of Conversion Technology of Biobased Resources,
Dominik.Wust@unitn.it University of Hohenheim, Garbenstrasse 9,
70599 Stuttgart, Germany
3
1
Department of Civil, Environmental and Mechanical Engineering, Institute of Catalysis Research and Technology, Karlsruhe Institute of
University of Trento, via Messiano 77, 38123 Trento, Italy Technology, Hermann-von-Helmholtz-Platz 1,
76344 Leopoldshafen-Eggenstein, Germany
Biomass Conv. Bioref.

MF Methylfurfural with C6 saccharides at similar conditions [7–11]. In other


FU Furfural words: once the hydrolysis yields its intermediates
DOC Dissolved organic carbon (monosacharides), the formation of HC occurs rather fast.
NDF Neutral detergent fiber, The rate limitation arising from the hydrolysis step can be
ADF Acid detergent fiber reduced by increasing the reaction temperature, which as a
ADL Acid detergent lignin drawback shifts the product distribution towards unwanted
PG Process gas transfer of carbon to the liquid and/or gaseous phase. In these
SUC Sucrose terms, a major challenge for HTC applications is to find an
GLU Glucose appropriate solution that combines efficient hydrolysis with
FRU Fructose HC formation. The most important precursor of HC is 5-
GIAD Glyceraldehyde hydroxymethylfurfural (HMF). Various pretreatment technol-
GAD Glycolaldehyde ogies for lignocellulose to produce HMF exist [12].
PA Pyruvic acid Mechanical crushing, among those, is a technology which
DHA Dihydroxyacetone can easily be applied prior to HTC but shall be kept to a
XYL Xylose minimum due to major energy consumption [13, 14]. The
ARA Arabinose effect of the particle size is therefore a parameter which needs
BDL Below detection limit to be further addressed. Another alternative to increase the HC
FAD Formaldehyde formation and carbon retention efficiency (Ceff) is the reuse of
BTO 1,2,4-benzenetriol the residual process water (rPW) obtained during the HMF
EtOH Ethanol production. When using fructose in aqueous acid-catalyzed
DHH 2,5-dioxo-6-hydroxy-hexanal solution, high HMF yields (up to 55 mol%) were reported.
ERY Erythrose The rPW that remains downstream still contains high concen-
FUA Furfuryl-alcohol trations of HMF and fructose (FRU) [15]. Since it is well
MeOH Methanol known that the conversion of C6 saccharides such as FRU to
AAD Acetaldehyde HMF and/or levulinic acid (LA) in aqueous systems suffers
MAN Mannose from the formation of up to 20 mol% yield of insoluble com-
OA Oxalic acid pounds known as humins, a recycling of its rPW within HTC
PAD Pyruvaldehyde can be a possible path to improve the Ceff of both processes.
ProA Propionic acid Concerning a commercial HTC plant, such a recycling ap-
ProeA Propenoic acid proach would reduce both the operating costs and the environ-
mental impact [16–18]. Indeed, rPW contains the key inter-
mediates involving an advantageous set of acid catalysts as
1 Introduction well as a favorable pH for HC production [19, 20]. The pur-
pose of this study is to investigate how biomass particle size
Hydrothermal carbonization (HTC) received increasing atten- and the “quality” of the reaction medium (bi-distilled water
tion in the past years due to its ability to convert wet biomasses versus rPW) affects HTC in terms of product distribution,
into valuable carbon materials. Kruse and Dahmen recently products characteristics, and Ceff.
reviewed the large amount of scientific publications in this Consequently, from the experiments conducted for the
field [1]. Those publications mostly evaluated the perfor- present work, it is expected that major fractions of HMF in
mance of a feedstock material under variation of process con- the rPW will react before the biomass is completely hydro-
ditions, mainly reaction time and temperature. The most dom- lyzed. When considering the use of rPW instead of bi-distilled
inant observation was the strong impact of temperature on water, the effect of the readily available HMF on the biomass
hydrochar (HC) yield and its properties. In this context, it is decomposition must be considered. In fact, a fast conversion
often observed that temperatures below 210 °C do not alter the of saccharides from biomass into HMF is of high interest to
properties of cellulose and lignocellulose remarkably due to allow a fast interaction between the newly created HMF and
the thermal stability of cellulose and lignin under these con- the HMF in the rPW and thus an efficient formation of sec-
ditions [2, 3]. Regarding cellulose, hydrolysis is the rate- ondary char as part of HC. Furthermore, there is limited re-
limiting step. It is known that the reaction kinetics is a search on the recirculation of process water from hydrother-
pseudo-first order, which indicates that the hydrolysis only mal conversion. Some examples included the use of process
takes place at the end of the cellulose chain [4, 5]. This is water from HTC of different biogenic residues such as loblol-
caused by hydrogen bonds within crystalline cellulose, which ly pine, poplar wood chips, miscanthus, paper as well as grape
do not break up in subcritical water [6]. This aspect becomes and orange pomace in addition to poultry litter and particular-
significant as HTC of cellulose is extremely slow compared ly not catalyzed by a Brønsted acid [18, 19, 21–24]. All
Biomass Conv. Bioref.

researchers reported an increase of HC yield by 5–10 wt.% Packard, GC-Oven 5800 Series II). After the reaction time
after the first recycle but the quality of HCs slightly decreased up to 140 min including 20 min of heating time, the autoclaves
in terms of carbon content and higher heating value due to were cooled down rapidly by blowing fresh compressed air on
precipitation of solved salts [18]. them. The amounts of liquid phase (process water (PW)),
For this study, brewer’s spent grains (BSG) as a major by- gaseous phase (process gas), and solid phase (hydrochar) were
product of brewing industry were selected because they were measured by weighing the autoclaves after each separation
proven to be convertible to C-rich chars but the average HC step. All experiments were conducted once, because it was
yields of around 50 wt.% are low compared with other ligno- known from prior experiments that the overall error is around
cellulosic biomasses [25, 26]. This sugar- and protein-rich sub- ±5% (relative) and the reproducibility is even better. The ob-
strate is mainly used in animal nutrition; however, it undergoes tained slurry containing the HC and the PW was separated by
microbial degradation and chemical deterioration especially in centrifugation (Hermle Labortechnik GmbH, Wehingen,
summer due to insufficient acceptance as a commodity, which Germany) and the HC was dried at 60 °C till it reached a
leads to the production of greenhouse gases and other contam- constant weight.
inants [27–30]. These aspects call for an alternative way of
converting BSG into valuable products, such as HC [31]. 2.1.2 Physicochemical analysis
BSG is a complex biomass; hence, it can be assumed that
reducing saccharides and their degradation products react with The ultimate composition of the HCs as well as of the different
amines from the degradation of the protein fraction and form a fractions of BSG was determined with an Euro EA -CHNSO
type of “tertiary char” via Maillard reactions in addition to a Elemental Analyzer (HEKAtech GmbH, Wegberg,
“primary char” formed by a direct solid conversion (intramo- Germany). The determination of ash of BSG and its particle
lecular condensation) of polysaccharide fragments and lignin size fractions as well as their corresponding HCs was per-
and to a “secondary char” formed by aldol condensation/ formed according to ASTM Standard Method Number
addition and polymerization of dissolved intermediates [32]. E1755-01 at 575 (± 25) °C for 180 min. To eliminate the
Therefore, the key question is if there is a positive or per- carbonates and to evaluate only the organic C in HCs, an
haps a negative interaction between intermediates from BSG ash determination at 1000 °C (oven temperature of
degradation with compounds present in the rPW, in view of Elemental Analyser) was also performed. To calculate the
HC formation. The reactions involved are described in detail fixed carbon (Cfix), the volatile matter (VM) was determined
by a complete reaction scheme except for the formation of according to the DIN EN 15148:2009 at 900 (± 10) °C for
tertiary char, whose formation would require additional re- 7 min. Additionally, the total structural composition of the
search efforts to be fully understood. biomass was determined with the Van Soest analysis (ADF,
NDF, ADL) and the total amounts of lipids were determined
by their extraction according to VO (EG) 152/2009 III H
2 Experimental section 2009-01 and 152/2009 III C 2009-01 [33]. For the calculation
of the protein content, the nitrogen content was determined
2.1 Material and methods according to Kjeldahl and multiplied by the factor 6.25, given
for cereal products (standard method ASTM D-52919). The
2.1.1 Experimental setup liquid phase was filtered (0.45-μm polypropylene filters) and
diluted 1:10 with the eluent for sample preparation for high-
The experiments were performed in small stainless steel performance liquid chromatography (HPLC). Subsequently,
(316 L; 1.4404) 13 ml batch-autoclaves filled for 70 v/v%. HPLC was used to determine and quantify the substances in
BSG from the private brewery Hoepfner GmbH were careful- the aqueous phase. The carbohydrates, organic acids, alde-
ly dried (60 °C for 48 h), milled to a particle size smaller than hydes, and alcohols were identified with a column from Bio-
1 mm and then three fractions were separated by sieving: 850– Rad Aminex HPX-87H (300 × 7.8 mm) in a Merck Hitachi
500 μm, 500–250 μm and 250–0 μm. Then, aqueous solu- Primade HPLC system. As eluent bi-distilled and degassed
tions were prepared consisting of 10 wt.% BSG and either bi- water was diluted with sulfuric acid (4 mM). The column
distilled water or the filtered (0.45 μm nylon filters) rPW from was operated at 25 °C with a flow rate of 0.65 ml min−1 for
HMF production (AVA Biochem AG, Muttenz, Switzerland). saccharides, aldehydes, furfurals, and most of organic acids
The solutions were converted via HTC at 230 °C for 0, 30, 60, except lactic and glycolic acids which showed a successful
and 120 min excluding preheating time. This means that 0 min separation at 70 °C with flow rate of 0.65 mL min−1. The
exactly corresponds to the preheating time. Additionally, a detector was a refractive index (RI) Chromaster 5450.
series of experiments were conducted with only rPW without Levulinic acid (LA), formic acid (FA), acetic acid (AA), lactic
BSG. The reactors were placed in an upright position in a acid (LaA), and glycolic acid (GA) were also calibrated via an
metal rack installed in a modified GC-Oven (Hewlett UV detector (1430 Diode Array Detector) at 210 nm. To
Biomass Conv. Bioref.

validate the results for 5-hydroxymethylfurfural (HMF), 2.2.2 Carbon retention efficiency related to brewer’s spent
methylfurfural (MF), and furfural (FU), another HPLC system grains
(Merck Hitachi, Düsseldorf, Germany) equipped with the col-
umn Lichrosper 100 RP-18 was used which was driven with
an eluent composition of water and acetonitrile (90/10) and a
C HC  mHC
flow rate of 1.4 mL min−1 at 20 °C. The total dissolved or- C eff ¼ x 100% ð3Þ
ganic carbon content (DOC) of the filtrate was determined by C BSG  mBSG
liquid chromatography-organic carbon detection LC-OCD Ceff is the carbon retention efficiency related to the BSG
(DIMATEC Analysentechnik GmbH, Essen, Germany). The weighed in, CHC the carbon content of the hydrochar, mHC is
concentrations of ammonium, nitrate, and nitrite in liquids the mass of the hydrochar, CBSG is the carbon content of the
were measured by means of the barcode cuvette tests LCK BSG, and mBSG is the initial mass of BSG (all on a dry basis).
303, 339, and 342 in the spectral photometer DR6000 (HACH
LANGE GmbH, Düsseldorf, Germany). The DOC values 2.2.3 Carbon retention efficiency related to residual process
were used to create a mass-related carbon distribution. The
water from HMF synthesis
mineral composition was analyzed by means of inductively
coupled plasma optical emission spectrometry in an ICP-OES  
cDOCPW xmPW þ mC;PG
Spectrometer (Agilent Technologies, Santa Clara, USA). C eff rPW ¼ 100%− x100% ð4Þ
cDOCrPW x mrPW
2.1.3 Thermogravimetric analysis
Ceff-rPW describes the carbon retention efficiency regarding
the HC formed from the organic compounds dissolved in the
The thermal decomposition of selected HCs was analyzed in a
rPW (cDOCrPW ). For this calculation the mass of C in the PW
thermobalance (STA 449 F3 Jupiter, Netzsch, Germany).
(cDOCPW ) and process gas (PG) (mC,PG) after the HTC is re-
Approximately 6 mg of HC were heated up to 105 °C and left
lated to the mass of C in the rPW (cDOCrPW times mrPW)
at this temperature for 10 min to ensure moisture release.
subtracted from 100% DOC. In other words, the DOC missing
Subsequently, the sample was heated up to 900 °C with a
after HTC and not transferred to PG has reacted to form HC.
heating rate of 30 K min−1. The measurement was conducted
under constant nitrogen flow of 70 mL min−1.
2.2.4 Carbon percentage of detected reaction products
in the process water from hydrothermal carbonization
2.1.4 Scanning electron microscopy
To illustrate how well the carbon balance in the process waters
Scanning electron microscopy (SEM) of dried and pulverized
after HTC of the residual process water from HMF synthesis
HCs was performed with a GeminiSEM 500 from Zeiss (soft-
with and without BSG as well as after the HTC of BSG in
ware: SmartSEM Version 6.01) with a thermal Schottky field
water could be closed, the following equation was established:
emitter cathode. The variable pressure system allowed pres-
sures up to 500 Pa that was applied in the case of non-
conductive samples. For secondary electron (SE) images, the M Carbon
x cReaction Product
SEM was equipped with an inlens SE detector, an Everhart- M Reaction Product
C% ¼ x 100% ð5Þ
Thornley detector, and a variable pressure secondary electrons cDOC
(VPSE) detector.
The carbon percentage of each reaction product (C%) was
2.2 Calculations calculated considering the molar percentage of carbon
(MCarbon) in the reaction product (MReaction Product) depending
2.2.1 Biochemical characterization of brewer’s spent grains on its concentration (cReaction product) related to the concentra-
tion of dissolved organic carbon (cDOC).
For calculation of the values for hemicellulose, cellulose, and
lignin, the following equations were applied:
3 Results and discussion
Hemicellulose ¼ NDF–ADF ð1Þ Firstly, the characterization of BSG and its particle size
Cellulose ¼ ADF–ADL ð2Þ fractions as well as the rPW as a reaction medium alter-
native to water is reported. Secondly, the results obtained
whereas NDF designates the neutral detergent fiber, ADF is from HTC of rPW as well as BSG in water and rPW are
acid detergent fiber and ADL is acid detergent lignin. presented and discussed.
Biomass Conv. Bioref.

3.1 Physicochemical analysis of feedstocks size as can be seen from the thermogravimetric analysis
(Fig. 1). The main degradable region for each particle size
3.1.1 Brewer’s spent grains fraction is located at 200–400 °C but their differences in struc-
tural composition are reflected in the strong decomposition
The contents of structural substances (biochemical characteris- peaks present between 250 and 400 °C. These peaks corre-
tics) as well as chemical characteristics presented in Table 1 are in spond to the decomposition of proteins and polysaccharides
accordance with the literature [26, 34–36]. Santos et al. [37] and such as hemicellulose, cellulose, and starch. However, the
Hardwick [38] reported that the chemical composition of BSG amorphous hemicellulose decomposes at a lower temperature
varies with barley variety, time of harvest, malting, mashing compared with the crystalline cellulose [42]. Since lignin con-
conditions, possible additives added, and brewing technology. sists of phenylpropanoid units interconnected by different
It can be proven by subtracting the ash content determined at chemical linkages which have different binding energies, the
575 °C from the ash content determined at 1000 °C and the decomposition proceeds over the abovementioned tempera-
content of potassium that 0.15% of 4.6% of the ash are carbon- ture range in Fig. 1 [43].
ates. However, this value will be neglected in calculations like for The decomposition of the “small” (0–250 and 250–
Ceff (Figs. 4, 7, and 13) due to its low magnitude. The ash is 500 μm) BSG particle size fractions shows a similar pattern.
mainly composed of Ca2+, Mg2+, P, and S (Table 1), whereas Conversely, the sample with the largest particle size (500–
elements such as Na+, Ba2+, Zn2+, Mn2+, and Sr2+ occur in traces 850 μm) shows faster decomposition rates between 230 and
(≤ 0.004%). Since the ash content of lignocellulosic biomass 400 °C, but slower decomposition rates at temperatures higher
varies from below 1% to more than 20%, BSG can be considered than 400 °C. The reason for the different particle peaks around
as a low ash-containing biomass [21]. Its alkali and alkali earth 300, 350, and between 400 and 500 °C in the decomposition
metal fraction could easily be leached out which has a catalyzing curve (dTG) of 500 to 850 μm may be due to a variable
effect on the formation of the secondary char fraction of HC, structural composition of the different particle size fractions.
affecting the secondary char particle dimension, texture, and sur- Bhatty [44] and Lynch et al. [27] gave hints for this hypothe-
face morphologies [39, 40]. sis. Since the smaller the particle size fraction, the higher its
Besides the mineral components, the saccharides in BSG content of C and H as well as its Cfix but the smaller its O and
play a decisive role as precursors in the HC formation [41]. volatile matter (VM) content (Table 2), it is likely that the
The chromatogram in Fig. S1 (Supplementary materials) macromolecular structure of lignin and xylans from broken
shows four relevant peaks related to oligosaccharides such barley hulls concentrated in it [45].
maltotriose and maltose, sucrose (SUC), glucose (GLU), and Consequently, the physicochemical characteristics of total
fructose (FRU), followed by lactic acid (LaA) and glycolic BSG fractions (≤ 1 mm), presented in Tables 1 and 2, corre-
acid (GA). spond to the major part of the particle size fraction 250–
Three different particle size fractions were used for further 500 μm with 48% followed by the fractions of 0–250 μm
investigation in this study. The characteristics are given in and 500–850 μm with 31% and 19% respectively but also to
Table 2. BSG is also composed of complex saccharides (cel- the minor fraction larger than 850 μm with 2%. The percent-
lulose and hemicellulose), proteins, and lignin. The accessi- ages were calculated based on the mass of BSG weighed in for
bility to these components is strongly related with the particle the sieving procedure.

Table 1 Physico- and biochemical analysis of brewer’s spent grains on a dry basis

Biochemical characteristics [wt.%]


Crude fiber NDF[a] ADF[b] ADL[c]/ Lignin Hemicellulose[d] Cellulose[d] Crude Fat Crude Protein[e]
17.7 56.3 29.3 6.8 27.0 22.5 2.3 21.4
Chemical characteristics [wt.%]
C H N S O[d] P Ca2+ Mg2+ K+
50.8 7.4 4.1 0.4 32.7 0.51 0.41 0.15 0.05
Atomic ratios [−]
O/C H/C
0.44 1.74
Physical characteristics [wt.%]
Ash VM [f] Cfix [g]
4.6 83.3 12.1
[a]
Neutral detergent fiber. [b]Acid detergent fiber. [c]Acid detergent lignin. [d]Calculated by difference. [e]Calculated by multiplying the total nitrogen value
by a factor of 6.25. [f]Volatile matter. [g]Fixed carbon. Hemicellulose = NDF – ADF. Cellulose = ADF – ADL [33].
Biomass Conv. Bioref.

Table 2 Ultimate and Proximate Analysis of particles of different size According to a simplified process flow sheet created by AVA
resulting from milling (< 1 mm) of softly dried Brewer’s Spent Grains
Biochem AG, the company used nitric acid (HNO3) for pH
Fraction 0–250 μm 250–500 μm 500–850 μm setting (pH = 1.0–1.5) and as a liquid catalyst for dehydration
of monosaccharides to 10–20 wt.% yield of HMF [46].
Elemental composition [wt.%] [wt.%] [wt.%] The detected organic reaction products listed in Table 3
C 49.5 49.1 47.7 correspond to those also detected by the research group of
H 7.0 6.9 6.6 Asghari [47] under acidic conditions (pH < 2).
N 4.2 3.6 2.2 Consequently, the 10 to 100% of DOC likely consists of glyc-
S 0.3 0.2 0.1 eraldehyde (GlAD), glycolaldehyde (GAD), pyruvic acid
[a]
O 34.6 35.4 38.6 (PA), and dihydroxyacetone (DHA). Nevertheless, these com-
Atomic ratios [−] [−] [−] pounds could not be unequivocally identified by the operating
H/C [−] 1.69 1.67 1.65 HPLC system.
O/C [−] 0.48 0.49 0.56 Table 4 shows the concentrations of inorganics which are
Physical characteristics [wt.%] [wt.%] [wt.%] present in rPW due to different reasons. Since ammonium and
Ash 4.4 4.8 4.8 nitrite were not found and it is well known that sugar syrup
VM[b] 81.6 82.6 83.0 does not contain any fats and proteins as precursor for com-
Cfix[c] 14.0 12.6 12.2 pounds containing organic nitrogen, it can be stated that the
[a]
total nitrogen bound in rPW is linked to nitrate and results
Calculated by difference. [b] Volatile Matter. [c] Fixed Carbon
from catalysis. The elements Al3+, Zn2+, Fe2+/3+, Cu2+, and
Mn2+ are assumed to originate from corrosion of components
3.1.2 Residual process water from HMF synthesis of the pilot-scale production plant of AVA Biochem due to the
acidic reaction medium, while Mg2+ and Ca2+ may result from
The rPW, a residual aqueous solution from HMF synthesis the sugar syrup production [48–50].
(un-concentrated sugar syrup) was used as an alternative to
water in this study. It was provided by the specialty chemicals
company AVA Biochem AG [46]. It still contains reactive 3.2 Hydrothermal carbonization of residual process
intermediate products of HTC such as HMF (47.8 C%) and water from HMF synthesis
FU (1.9 C%) as well as its precursor FRU (34.8 C%), which
makes it more attractive as a reaction medium than water [22]. As shown in Table 3, rPW from HMF synthesis is a hetero-
Table 3 presents the identified compounds in rPW and their geneous medium composed of 47.8 C% HMF, 34.8 C% FRU,
detected concentrations as well as their carbon percentages and around 17.0 C% of their degradation products (including
related to 100% DOC. the 10.0% of analytically not identifiable DOC). The presence

Fig. 1 Thermogravimetric 100 0.8


analysis of the separated particle TG 0-250 µm
size fractions of brewer’s spent TG 250-500 µm + rPW
grains (BSG) TG 500-850µm
80
dTG 0-250 µm
0.6
dTG 25-500 µm + rPW
dTG 500-850 µm
-dm/dT [% K-1]
Mass loss [%]

60

0.4

40

0.2
20

0 0.0
200 300 400 500 600 700 800
Temperature [°C]
Biomass Conv. Bioref.

Table 3 Concentrations Furthermore, the findings concerning HC formation agree


of reactions products in Parameter c [mg L−1] C% of DOC
with the concept proposed by Dinjus et al. [32] that non-
residual process water
from HMF production DOC[a] 10,241 degraded leftovers of lignocellulosic biomass, consisting of
identified by high- FRU[b] 8911 34.8 fragments of (hemi)cellulose as well as lignin, undergo a di-
performance liquid chro- FAD[c] 72 0.3 rect solid conversion (intramolecular condensation) and form
matography and their
AA[d] 72 0.2 a (poly)aromatic network. In addition to this HC fraction,
carbon percentages relat-
ed to 100% of dissolved GA[e] 372 1.1 termed primary char, secondary char results from aldol
organic carbon HMF[f] 8566 47.8 condensation/addition and polymerization of degradation
FU [g]
314 1.9 products of monosaccharides and lignin [72–74]. Because of
LA[h] 614 3.1 the similar polyol structure of GLU and xylose (XYL), their
FA[i] 323 0.8 degradation in subcritical water was proven to be similar, al-
Total 90.0 beit their dehydration products are HMF (GLU) and FU
(XYL) [72]. Arabinose (ARA) as a component of hemicellu-
[a]
Dissolved Organic Carbon. [b] Fructose. lose and its reaction pathways are not illustrated due to its low
[c]
Formaldehyde. [ d ] Acetic Acid. occurrence in biomass or rather BSG as well as its low reac-
[ e ]
Glycolic acid. [f] 5-
Hydroxymethylfurfural. [g] Furfural. tivity within short reaction times due to the occurrence of its
[h]
Levulinic acid. [i] Formic acid more stable tautomer (α-pyranose) [75].
To be precise a tertiary or even a quaternary char has actu-
of these compounds is expected to have a favorable influence ally to be described which results from Maillard reactions or
on the overall efficiency of HTC compared with water, be- the incorporation of fatty acids but focusing on it in detail is
cause these organic compounds are the precursors for the for- beyond the scope of this study [71]. As reported by Funke
mation of a carbonaceous material, known as “humins or sec- et al. [51], a process gas (PG) consisting mainly of CO2 with
ondary char” with a sphere-like geometry [11, 19, 51–55]. minor quantities of CO, CH4, and H2 as well as traces of CmHn
Furthermore, HNO3 was reported having positive effects on results from decarboxylation, decarbonylation, and oxidation
aldol condensation/addition and polymerization reactions reactions occurring during HTC.
resulting in a solid product with a higher carbon content. The reaction pathways which are accountable for the con-
The reasons are that the acidic conditions catalyze these reac- version of rPW due to its composition are highlighted in or-
tions and more reactive intermediates with a higher abundance ange color in Fig. 2. The factual identified ones are illustrated
of hydroxyl functional groups are formed by oxidation [20]. as solid lines and those that are unlikely to happen as dashed
Figure 3 and Table S1 (Supplementary Materials) display lines. The standard black lines indicate the most important
the carbon percentages related to DOC of different com- reaction pathways to primary and secondary char.
pounds in the process water after HTC. FRU cannot be de- Figure 3 and Table S1 (Supplementary Materials) shows the
tected anymore after the heating phase, and HMF is only pres- carbon percentage of the detected reaction products related to
ent at the first time point (t = 0 min). For a better understand- the DOC of the process waters after HTC. The DOC was re-
ing of the HC formation in subcritical water in a temperature duced by around 42% from 10,241 to 5,900 mg L−1 (not illus-
range from 170 to 350 °C and to explain the detected compo- trated results) during the reaction time due to carbon transfer to
nents, a scheme illustrating the decomposition pathways of the gaseous phase caused by decarboxylation, disproportion-
(ligno-)cellulosic biomass and the subsequent oligo-, di-, and ation and oxidation reactions but also due to carbon consump-
monosaccharides is presented in Fig. 2 [9, 47, 54, 56–71]. tion by aldol condensation, subsequent polymerization and aro-
matization reactions which resulted in secondary char [8, 76].
Whereas Van Dam and Bekkum [77] and Cottier and
Descotes [78] showed that aqueous and non-aqueous degra-
Table 4 Concentrations dation processes of oligo-, di-, and monosaccharides lead to
of inorganics in process Parameter c [mg L−1]
about 37 products, in this study, 13 organic compounds were
water from acid-
catalyzed conversion of NO3− 2652.7 verified in rPW by two HPLC systems and three methods, and
fructose analyzed by cu- Al3+ 0.1 hereby, up to 90.0 C% of DOC was identified. Since GLU is
vette test in a spectral Mg2+ 0.2 always below its detection limit (Fig. 3 and Table S1 -
photometer and induc-
Ca3+ 0.8 Supplementary materials), the isomerization of FRU to GLU
tively coupled plasma
optical emission spec- Fe2+/3+ 1.8 by the Lobry de Bruyn-Alberda van Ekenstein transformation
trometry in an ICP-OES Mn2+ 0.05 can be excluded in this experimental series. It mostly occurs
spectrometer under mild acidic conditions under which the weak acid anion
Zn2+ 0.2
Cu2+ 0.1 acts as a base but catalysis by dissolved metal compounds is
basically possible [77, 79–81].
Biomass Conv. Bioref.

Maillard Reactions

Oxidation
Decarboxylation

Pyruvic acid
Amino acid Oxidation Decarboxylation

Oxalic acid Glycolic acid Methanol

Reverse aldol condensation

Hydrolysis
Oxidation
Acetic acid
Cross-Dispropotionation

Oxidation

Oxidation
Self-Dispropotionation
Formaldehyde
Dehydration +

Hydrogenation Propionic acid


Dehydration H2O-addition + Decarboxylation
Protein Rearrangement Ethanol
Lactic acid Dehydration
Dihydroxyacetone Pyruvaldehyde
Hydrolysis Degradation
Decarbonylation Propenoic acid

Isomerization
Reverse aldol
Reverse aldol condensation
Fructose condesation Acetaldehyde
Sucrose Rehydration Cross-Disproportionation
Glyceraldehyde
Rehydration
Formic acid
Dehydration Rehydration

Isomerization HMF Reverse aldol


additon
Levulinic acid
Condensation + Polymerization
Starch (Diels-Alder) Nuclei
Furfuryl Dimerization +
alcohol Degradadation
Reverse aldol 1,2,4-Benzenetriol
addition

Aromatization + Decarboxylation
Hydrolysis Hydrolysis Hydrolysis +
+ Dehydration
Electrocyclic
Rearrangement
Furfural Methylfurfural
Levuglucosan Glucose Rehydration
(Hemi-) Cellulose
Dehydration
Decarboxylation
Intramolecular
Condensation

2,5-dioxo-6-hexanal

Reverse aldol Reverse aldol condensation + Degradation


condensation +
Intramolecular
Condensation + Primary Glycoladehyde
Secondary/
Erythrose
Decarboxylation
Char Tertiary Char

Fig. 2 Decomposition pathways of relevant organic components of biomass under subcritical water conditions in a temperature range from 175 to
350 °C. The formation of gases by decarboxylation, decarbonylation, etc. are not illustrated (adapted from Ref. [9, 47, 54, 56–71])

Consecutive reactions like the dehydration to HMF and its are no clear indications that FRU, GLU, and XYL were in-
rehydration to FA and LA as well as their degradation again volved in condensation and polymerization reactions. Horvat
took place, as other research groups also proved under com- et al. [87] and Patil et al. [54] proposed that an intermediate
parable acidic process conditions [47, 82, 83]. These seem the such as 2,5-dioxo-6-hydroxy-hexanal (DHH) is firstly formed
most promoted reactions catalyzed by remaining protons from via the rehydration (ring opening mechanism) of HMF. This
HNO3 catalysis during HMF synthesis [47, 57, 83]. dimer acts as the basis for the addition of a maximum of five
Considering that furans are potentially polymerizable com- HMF molecules. As a result of this, aldol addition/
pounds, they were not detectable anymore, already after 0 min condensation nuclei are formed that further agglomerate to
of reaction time (Fig. 3 and Table S1 - Supplementary mate- secondary char (see in Fig. 5) [17, 85].
rials) [19, 53, 54]. Consecutive reactions of HMF can gener- Additionally, LA together with FA were produced from the
ally be divided into three groups, according to its functional furan ring cleavage of HMF interestingly in a molar ratio greater
groups, namely, the furan ring, formyl group (aldehyde), and than 1:1 (Figs. 2 and 3), which is in accordance with the investi-
hydroxymethyl group (alcohol). As illustrated in Fig. 2, 1,2,4- gations of Girisuta et al. [66] and Reiche et al. [20] on the decom-
benzenetriol (BTO) is the product formed by furan ring open- position of HMF into LA under acidic conditions. Flanelly et al.
ing and by closing of a cyclohexadiene or triene ring after [88] stated the dehydration of FRU, GLU, and erythrose (ERY) to
electrocyclic rearrangement and dehydration [65, 84, 85]. furfuryl-alcohol (FUA) and FU via reverse aldol condensation and
Since its carbon percentage related to the DOC keeps constant rehydration is the responsible reaction pathways for the excess of
from 60 min of reaction time on (Fig. 3 and Table S1 - FA (Fig. 2). Qi et al. [9] found that FA additionally catalyzes
Supplementary materials) and considering that BTO is less dehydration of FRU to HMF and LA molecules react with the
reactive, a common polymerization with HMF, and thus, an hydroxyl groups of secondary char, and thus, it takes part in the
implication in the formation of secondary char is unlikely [59, growth (nucleation) as building units. However, LA and FA were
86]. However, Ryu et al. [72] studied the effect of such phe- also subjected to decarboxylation and mainly ended in CO2 and
nolic compounds on the formation of HC from saccharides, H2O as well, even embedded in the structure of secondary char
demonstrating a substantial increase in carbon yield. [66, 89]. Such insights agree with the work of van Zandvoort et al.
According to the experimental and modeling studies of [17] and Cheng et al. [90] who found that secondary char origi-
Patil et al. [54], Jung et al. [40], and Dussan et al. [75], there nated from FRU consists of 60% furan rings connected via 20%
Biomass Conv. Bioref.

100%

80%
carbon percentage of DOC [C%]

60% 1,2,4-benzenetriol
Ethanol
Formic Acid
Levulinic Acid
Methylfurfural
5-hydroxymethylfurfural
40%
Acetic Acid
Not Detected

20%

0%
0 30 60 120
reaction time [min]

Fig. 3 Carbon percentages of detected reaction products in the process waters from hydrothermal carbonization of residual process water from HMF
production in [mg L−1] (without biomass)

aliphatic linkers, decorated with oxygen-containing functional comparison of the dTG curves relevant to secondary char
groups, such as carboxyl, carbonyl, and hydroxyl. A decarboxyl- produced from rPW (Fig. 10), to dissolved degradation prod-
ation of these carboxylic functional groups originated from LA ucts of digestate by Funke et al. [92] and to Kraft lignin by
was confirmed by Baccile et al. [89], but according to Li et al. Rasrendra et al. [93], all show a similar thermal decomposition
[91], this reaction was rather slow at a reaction temperature of trend which starts around 250 °C and levels off at around
230 °C. However, it is promoted above this temperature and 800 °C with highest degradation rate around 400 °C. This
under the acidic conditions applied here. A slightly increasing insight is a strong indication of the thermal stability of this
carbon content in the secondary char from 68 to 73%, which kind of secondary char.
fits to the results of Poerschmann et al. [76], and a decreasing Figure 5 shows the indisputable relation between the
DOC in PW from 30 to 23% (Fig. 4), may mean decarboxyl- growth of carbon spheres and reaction time, which in
ation or oxidation reactions rather than embedding of LA in turn is directly related with the increase in the carbon
HC albeit to a minor extent. This is in accordance with van percentage in the HC (Fig. 4) but not with an increase
Zandvoort et al. [17] and Reiche et al. [20]. Nevertheless, the in its yield. This indicates a growth mechanism through
longer the reaction time, the higher the CeffrPW regarding the coalescence [40], visible especially from 30 min of
secondary char formed from the organic compounds dis- reaction time (rPW for 30 min). Finally, the results pre-
solved in the rPW, which yields up to 35% (Fig. 4). sented in this study diverge from those obtained by Qi
Ongoing decarboxylation reactions in addition to dehydra- et al. [9], who proved a slowed growth of carbon
tion reactions are obvious by considering the minor reductions spheres within 6 h of reaction time due to the integra-
in O/C and H/C molar ratios, to a value comparable with hard tion of LA in the char structure and its less reactive
coal, as illustrated in the van Krevelen diagram (Fig. 9). By functional groups.
Biomass Conv. Bioref.

100 2 70
4 4 4

23 60
30 25 24
80

50
percentage [%]

Ceff_rPW [%]
60
40

30
40
71 72 73
68
20

20
10

0 0

reaction time [min]

C in HC C in PW C in PG Ceff_rPW
Fig. 4 Carbon distribution within the product phases after hydrothermal carbon; HC = hydrochar; PW = process water; PG = process gas;
carbonization of residual process water from HMF production at 230 °C CeffrPW = carbon retention efficiency regarding HC formed from the
for 0, 30, 60 and 120 min. The process gas is assumed as 100% CO2. C = dissolved organic compounds in the residual process water

The oxidation of LA to AA and FA, as Fang et al. [61] concentrations of FA decreased over a reaction time of 120 min
reported, could not definitely be proven, but Reiche et al. [20] is the cross-disproportionation with formaldehyde (FAD) to
confirmed the degradation of LA and FA by a decrease of pH methanol (MeOH), but it cannot be verified due to the low
down to 0 at 220 °C for 6 h. A further explanation why the concentration of FAD as reactant (Table 3 and Fig. 3) [63].

Fig. 5 Scanning electron


microscopy of hydrothermal
carbonized (reaction
temperature = 230 °C; reaction
time = 0–120 min) residual
process water from HMF
production; rPW = residual
process water from synthesis of 5-
hydroxymethylfurfural
Biomass Conv. Bioref.

Another possible reaction pathway which resulted in MeOH is EtOH as a secondary product of this reaction pathway can
the decarboxylation of GA which includes CO2 formation and be verified but with a decreasing concentration till it is below
the self-disproportionation of FAD or Cannizzaro reaction [63, the detection limit at 120 min reaction time (Fig. S2 –
94, 95]. Since MeOH is not detectable and according to Fig. 4 Supplementary materials, peak #7 and Fig. 3). This phenom-
up to 4% of the initial DOC converted to CO2, we conclude that enon is based on the oxidation via acetaldehyde (AAD) to AA
disproportionation reactions of FAD, as well as reverse aldol as well as CO2. Which of the reaction products dominate
condensation of FRU to GlAD, followed by the oxidation to depends on the initial O2 concentration. Since the concentra-
GA and the subsequent decarboxylation intensively took place. tion of O2 is higher than 150 mmol L−1, being already provid-
Therefore, it is assumed that the sum peak #6 in Fig. S2 ed by HNO3− in rPW, the reaction to AAD is expected to
(Supplementary materials) corresponds to only LA probably dominate [64, 103]. However, the carbon percentage of
overlapping a peak of DHA. EtOH is generally low which results in a not verifiable con-
On the contrary, the reaction of FRU to AA verifiably took centration of AAD. Whether the decarboxylation of
place under these reaction conditions (Fig. S2 – pyruvaldehyde (PAD) to AAD and the reverse aldol conden-
Supplementary materials, peak #5) due to reverse aldol con- sation to FA occurred, could not be proven (Figs. 2 and 3).
densation and consecutive isomerization [47, 62, 96]. Therefore, further acid-catalyzed reactions of LaA such as the
However, it has to be considered that AA could also have dehydration to propenoic acid and its decarbonylation to AAD
LA and furans such as HMF and FU as educts [61]. Other were neglected [60, 62, 85, 104]. GlAD in addition to ERY
derivatives of HMF were FU due to the loss of its side group could be formed from GLU in the open form via a reverse
and MF formed by dimerization and further thermal degrada- aldol condensation (Fig. 2) but both are not detectable due to a
tion as well as DHH as a nucleus for polycondensation of FU- lack of isomerization to FRU. The reason why the not identi-
like compounds via their carbonyl group [54, 56, 87, 97–100]. fiable carbon calculated on DOC is reduced by the increase of
So far, oxidation reactions of carboxylic acids and FU- reaction time may be related to the occurrence of decarboxyl-
like compounds were proven under supercritical water con- ation and gas formation as well as the promotion of aldol
ditions or oxygen excess [58, 59, 97]. In this study, the condensation reactions towards the formation of secondary
authors did not apply a high temperature and they exclude char (Fig. 2).
an influence of the gaseous atmosphere inside the HTC
reactors, but the rPW still contains HNO3 which is able 3.3 Hydrothermal carbonization of brewer’s spent
to oxidize hydroxyl and aldehyde groups of FRU and its grains in water
degradation products [20, 101]. As a result, AA as a stable
intermediate product can be detected with a slightly in- As Table 1 illustrates, the BSG consists of 27.0% hemicellu-
creasing concentration by enhancing the reaction time up lose, 22.5% cellulose, and 6.8% lignin as structural substances
to 120 min (Fig. 3 and Table S1 - Supplementary materials) [105]. As commonly known, hemicellulose is composed of
[58]. Likely, a pH of 1.5 and increasing the reaction time pentoses such as xylose (XYL) and ARA with few hexoses
do not cause the oxidation of AA and correspondingly the such as GLU and mannose (MAN) in between and acts as a
release of CO2 as observed by Fang et al. [61]. The phe- connection between the fibers of cellulose and the 3-
nomenon that the pH value increased from 1.5 to 2.0 after dimensional aromatic structure of lignin [75, 106, 107]. The
60 min of reaction time, as illustrated in Tab. S1 recalcitrance of this hetero-matrix and the high crystallinity of
(Supplementary Materials), indicates that the dissociation cellulose pose a major challenge to split it into its components
of HNO3 was pushed back by the strong presence of car- [108]. Additionally, the greater the particle size, the worse the
boxylic acids such as LA, FA, and AA in PW [20, 101]. transfer of H2O in and the transfer of hydrolysis products
The remaining, not analytically identifiable reaction prod- throughout the pores [109, 110]. The reason is that smaller
ucts of the reaction pathways from FRU to LaA via GlAD and particles have a larger specific surface and thus show a larger
its keto-enol tautomers probably reacted further and this could mass transfer resulting in a higher reactivity [111]. As Yan
be the reason for the decreasing value of the ‘not detected’ et al. [112], Kim et al. [2] and Reza et al. [109] reported, within
species (Figs. 2 and 3). In the literature, keto-enol the biomass particles, the bonds between the monomer of
tautomerization and its consecutive reactions, e.g., via LaA hemicellulose could be easily split by releasing XYL (not
by benzylic rearrangement and decarboxylation to EtOH, are calibrated here) and crystalline cellulose began to hydrolyze
described as very fast under acidic conditions [56, 60, 82, 94, to produce anhydrosaccharides such as levuglucosan (not cal-
102]. This is the reason why LaA as an intermediate product is ibrated here) and furthermore its hydrolysis product GLU
not verifiable at a HPLC analysis with a column temperature [113, 114]. The split of the polymers led to the release of
of 70 °C. That means the peak #3 in Fig. S2 (Supplementary acetyl groups connected with a pH decline till the length of
materials) can be identified as GA but with a peak area below the oligomer was low enough to be solved and to be trans-
its detection limit. ferred through the pores of the biomass particles [115]. As
Biomass Conv. Bioref.

soon as there was a lack of acetyl groups the mass transfer hydroxyl group of GA [121]. However, the question raises if
through the pores slowed down [14]. As a consequence, the calcium and magnesium, released from BSG (Table 1) during
dehydration of released XYL from the fraction 500–850 μm hydrothermal treatment, cause the precipitation of OA as ox-
results in a carbon percentage of FU of about 19.0 C% related alates and affect its detectability [122–124]. Peak #11 has a
to the DOC in PW, whereas that for the fraction 250–500 μm large area and is supposed to have a high impact on filling the
is almost halved (10.3 C%) and that for 0–250 μm is already large gap in the DOC balance (designated as “not detected” in
around 0.5 C% after the heating time (Fig. 7). This dehydra- Fig. 6). Surveying the chromatogram of the UV detector of
tion reaction was catalyzed by the increase in carboxylic acids our HPLC system, peak #11 (Fig. S3 – supplementary mate-
such as FA till 30 min of reaction time and the resulting in- rials) has to be a carboxylic acid [125]. According to the bul-
crease of H+ protons [75]. letin of the HPLC column provider BIO-RAD and consider-
FRU was either formed by hydrolysis of leftover SUC from ing the operation conditions of the system, it may be propionic
brewing or it was a product from isomerization of GLU under acid (ProA) in a sum peak with propenoic acid (ProeA) as
the present weak acidic conditions as well as the influence of derivatives of LaA proven by Jin et al. [126]. Furthermore,
metal compounds known as Lewis acid catalysts [77, 79–81, the reaction products of the keto-enol tautomerization of
116]. On the one hand, FRU and GLU were dehydrated to GlAD but also the carbon percentage of proteins and amino
HMF, with a carbon percentage of 1.9 C% in PW of the frac- acids as well as phenols and fatty acids likely contribute to the
tion 0–250 μm but between 5.0 and 6.0 C% in PW of the other lack in the total ‘not detected’ species (Fig. 6 and Table S2 -
fractions within heating time. On the other hand, the reverse Supplementary Materials) [47]. On the one hand, the dehydra-
aldol condensation to AA, for GLU with GAD and ERY as tion of LaA to AcA and its decarboxylation to AAD was so far
intermediate products, and FAD as well as GA and LaA was described at reaction temperatures higher than 340 °C or
initiated within heating time and with an increasing concen- strong acidic conditions [60, 62, 85, 104]. On the other hand,
tration subsequently (Fig. 6 and Table S2 - supplementary it cannot be excluded at a pH of around 4(± 0.5) and with the
materials) [47, 62]. GA may also be a product from deamina- presence of alkali salts, which usually support the splitting of
tion of amines originated from amino acids [117, 118]. As a C–C bonds. AAD showed a further reaction to AA and ace-
consequence, when the carbon percentage of GA increases, tone [85]. As final products of these reaction pathways, gases
that of LaA decreases. As shown in Fig. 6 and Table S2 like CO, CO2, CH4, and H2 were released, which was con-
(Supplementary Materials), the reaction pathway of LaA via firmed by HTC experiments under comparable conditions and
decarboxylation and oxidation to AA can be verified by the subsequent gas chromatography by Kruse et al. [26].
observation of firstly an increase, but after 30 min a decrease, An explanation that the carbon percentages of HMF and
of the carbon percentage of EtOH and simultaneously an in- FU sharply decrease during HTC of particles with sizes great-
crease in the carbon percentage of AA [119]. The increase in er than 250 μm within 30 min may be the fact that HMF gets
carboxylic acids leads to a decrease of the pH to 3.5–4.0 rehydrated to LA and, to a minor extent, also FA [127]. It is
outside the biomass particles and/or within the reaction medi- obvious that they partly react via aldol condensation/addition
um within heating time but then an increase occurs to 4.0–4.5 or Maillard reaction towards secondary or tertiary char.
[105]. The reason is that under these reaction conditions, fur- However, the carbon percentage in HC is decreasing with
ther consecutive reactions occur. Representatives are the oxi- reaction time (Fig. 7). Since the amount of PG is increasing,
dation of LA to AA as well as FA which obviously get likely decarboxylation reactions of carboxylic acids dissolved
decarboxylated and does not react via cross- or embedded in secondary char are promoted at this stage of
disproportionation with FAD available at a high concentra- hydrothermal conversion [20, 89]. This phenomenon has a
tion. The formation of MeOH as second product of the last consequence for the Ceff of HTC of all particle size fractions
described reaction cannot explicitly be verified because its (Fig. 7) which firstly rises to more than 72% after 30 min, but
peak appears in the HPLC chromatograms of Fig. S3 declines again by 2% and more, e.g. for HTC of the particle
(Supplementary materials, peak #10) at the same retention size fraction of 0–250 μm, up to 120 min of reaction time.
time as LA and DHA [118]. The behavior of Ceff for HTC of the particle size fraction of
Further sum peaks are those of PAD, GlAD and ERY (peak 500–850 μm, starting from 67%, increasing to 74% after
#2) as well as LaA and GA (peak #3). Furthermore, there are 30 min and subsequently decreasing to 71% at 120 min of
not specified peaks with #1 and #11. As substances for peak reaction time, constitutes an exception again. The reason is a
#1 either extracted fragments of hemicellulose and/ or perhaps delayed mass transfer through the pores of biomass particles
maltotriose hydrolyzed from leftover, recalcitrant starch or affecting hydrolysis, dehydration, decarboxylation and con-
oxalic acid (OA) are possible [120]. Poerschmann et al. [76] densation reactions [109, 110]. This evolution is clearly visi-
detected OA in the PW obtained from HTC of BSG at 200 °C ble by a comparison of the SEM pictures of the HCs from the
for 24 h by a HPLC system. OA can be formed either by “smallest” and “biggest” BSG particle size fractions in Fig. 8.
reduction of carbonyl groups of PAD or by oxidation of In the SEM picture “0–250 μm + H2O for 30 min,” there are
Biomass Conv. Bioref.

0.4

1,2,4-benzenetriol
Ethanol
carbon percentage of DOC [C%]

Formic Acid
Levulinic Acid
Methylfurfural

0.2 Furfural
5-hydroxymethylfurfural
Lactic Acid
Glycolic Acid
Acetic Acid
Formaldehyde
Sucrose
Glucose
Fructose

0
0 30 60 120 0 30 60 120 0 30 60 120
0-250 µm 250-500 µm 500-850 µm
reaction time [min]

Fig. 6 Carbon percentages of detected reaction products in the process waters from hydrothermal carbonization experiments with brewer’s spent grains
in distilled water in [mg L−1] for each particle size fraction related to the dissolved organic carbon (DOC)

already coagulating carbon spheres visible, whereas in the further benzenediols as well as O-functionalized phenols, phe-
reference picture (500–850 μm + H2O for 30 min) spheres nolic acids (benzoic and cinnamic acid type) and
of different diameters can be seen. The SEM picture “0– cyclopentenone is possible due to solvolysis of lignin, but it
250 μm + H2O for 120 min” partly shows larger diameters occurs to a minor extent below 275 °C [51, 129]. As reported
of spheres in the micrometer range. This insight goes along by Ryu et al. [72] as well as Koch and Pein [73], these phe-
with those of Jung et al. [40] and Sevilla and Fuertes [11] who nolic compounds together with FU-like compounds are in-
investigated the formation, growth mechanism, and structural volved in aldol condensation/addition reactions towards the
properties of carbon spheres from FRU and GLU at compa- formation of secondary char.
rable reaction conditions. Therefore, HCs obtained from HTC of BSG at 230 °C for
As Kruse et al. [85] stated, the formation of the phenolic up to 120 min are generated as a consequence of complete
compound BTO, which is also involved in secondary char transformation of the hemicellulose rather than cellulose,
formation, is depressed in the presence of salts originated from which can be assessed by thermogravimetric analysis (TGA)
biomass because other reaction mechanisms possibly domi- of BSG and its HCs as well as by comparison of Figs. 1 and 11
nated. In the current study, the formation does not occur before [130]. The dTG curve shows that after 30 min of reaction time,
30 min of HTC of the fractions 250–500 and 500–850 μm. a fraction of the cellulose is still present in the HC that lowers
For the fraction 0–250 μm, BTO may have already reacted to the overall carbon content. This means that the carbonization
form carbon spheres; however, the depression of its formation is not fully completed, but strongly progressed after 120 min
due to its inertness can be supposed [128]. The occurrence of [26]. A small peak between 300 and 400 °C for cellulose
Biomass Conv. Bioref.

100 80
6 5 4
14 15 15 12 13
17 17 17
26 78
80 26 29 31
20 20
22 21 21 19
20 20 76
18
Percentage [%]

60
74

Ceff [%]
72
40
69 67 69 67
63 62 64 64 64 65 64
56 70

20
68

0 66

0-250 µm 250-500 µm 500-850 µm


reaction time [min]

C in HC C in PW C in PG Ceff

Fig. 7 Carbon distribution within the product phases after Hydrothermal PW = process water; PG = process gas; 0–250, 250–500, 500–850 μm =
Carbonization of brewer’s spent grains with water at 230 °C for 0, 30, 60 brewer’s spent grains fraction carbonized at 230 °C for 0, 30, 60 and
and 120 min. The process gas is assumed as 100% CO2. HC = hydrochar; 120 min in water; Ceff = carbon retention efficiency

supports this interpretation [1, 92]. As consequence and con- products such as FU-like compounds are able to react with
sidering the contents of hemicellulose in comparison with the amine group of these amino acids via carbonyl groups
cellulose of BSG in Table 1 and the sharply declining carbon [133–136]. However, Wang et al. [137] found out for HTC of
percentages of FU towards HMF in Fig. 6 and Table S2 food waste that enhancing the reaction temperature above
(Supplementary Materials), the resulting secondary char is 180 °C for 60 min of reaction time results in weakened
FU (XYL) based rather than HMF (GLU) based. However, Maillard reactions due to degradation of the more reactive re-
this insight confirms the cross-polycondensation of these ducing saccharides. This affects the carbon percentage in HC as
compounds as well as the self-condensation of FU, even well as the Ceff negatively, as it can be seen in Fig. 7, as well as
though Poerschmann et al. [76] reported its low reactivity. rises the carbon percentage of “not detected” species shown in
According to van Zandvort et al. [17], this fact proves that Fig. 6 and Table S2 (Supplementary Materials). Nevertheless,
secondary char from XYL (FU) has a lower amount of ali- stable aromatic structures by conversion of carbon spheres into
phatic groups but more pseudo-aromatic groups and is more an aromatic network and building blocks that favored N incor-
chemically resistant. poration were formed. Additionally, amino acids got
BSG contains a substantial amount of proteins (21.4 wt.%, decarboxylated to produce amines which on the one hand were
Table 1), which are less stable in subcritical water and get deaminated to produce carboxylic acids such as GA and am-
hydrolyzed to amino acids by breaking the peptide bond monia, and on the other hand underwent cyclization and con-
[117, 131]. Nevertheless, this depolymerization was slower densation to generate pyridine-N in HC [36, 118, 137, 138].
than the hydrolysis of (poly-)saccharides because the peptide According to Qiao et al. [139], the released ammonia catalyzed
bond in the protein is more stable than ß-1.4- and ß-1,6-glyco- polymerization reactions again and thus additionally promoted
sidic linkages in starch and cellulose; however, it was catalyzed the growth of carbon spheres (see in Figs. 8 and 12).
under the present acidic conditions [132]. As a consequence, it In the case of lipids, which are present in a minor amount
is foreseeable that reducing saccharides such as maltose origi- (2.3 wt.%, Table 1), the ester bonds were hydrolyzed into fatty
nated from recalcitrant starch or glucose originated from acids, such as linoleic or palmitic acid [36]. Whether these
(hemi-)cellulose (Fig. 1), and furthermore, dehydration fatty acids participated in HC formation is unclear. Instead,
Biomass Conv. Bioref.

Fig. 8 Scanning electron


microscopy of hydrothermal
carbonized (reaction
temperature = 230 °C; reaction
time = 0–120 min) brewer’s spent
grains of particle size 0–250 and
500–850 μm in water

2.0 BSG <1mm

BSG 500-850µm

BSG 250-500µm

BSG 0-250µm

1.5

1.0

0.5

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Fig. 9 Coalification degree of hydrochars obtained from hydrothermal blue-colored shapes describe the Hydrothermal Carbonization experi-
carbonization of residual process water from 5-hydroxymethylfurfural ments with water (H2O) as reaction medium. The orange-colored shapes
production and brewer’s spent grains as well as its mixture at 230 °C describe the hydrothermal carbonization experiments with the residual
for 0, 30, 60, and 120 min illustrated in the van Krevelen diagram by process water from 5-hydroxymethylfurfural synthesis (rPW) as reaction
plotting H/C atomic ratios versus O/C atomic ratios. Triangles and medium. The orange-colored circles stand for the hydrothermal carboni-
squares refer to brewer’s spent grains’ particle sizes fraction 0–250 and zation of the pure residual process water 5-hydroxymethylfurfural
250–500 μm, respectively. Rhombs refer to the fractions 500–850 μm synthesis
and the cross stands for altogether (< 1 mm), each black colored. The
Biomass Conv. Bioref.

100 0.6
TG 0-250 µm + H2O
The H/C and the O/C atomic ratios of HCs produced from
TG 0-250 µm + rPW the particle size fractions 0–250 and 250–500 μm are reduced
TG rPW 0.5
dTG 0-250 µm till they resemble the character corresponding to bituminous-
80
dTG 0-250 µm + rPW rich brown coals. HCs from the particle size fractions 500–
dTG rPW 0.4
850 μm can be ranked as peat after the heating time, as lignin

-dm/dT [% K-1]
Mass loss [%]

after 30 and 60 min, and as low-rank brown coal after


60 0.3
120 min. Since the particle size fractions 0–250 and 250–
500 μm can be described as almost chemically equivalent
0.2
(Table 2), their HCs marginally differ in H/C and O/C atomic
40
ratios and partly appear in the van Krevelen diagram as over-
0.1
laying points.
Kruse et al. [26] and Ryu et al. [72] obtained comparable
20 0.0
200 400 600 800 results for the HTC of BSG as well as a mixture of model
Temperature [°C] substances including monosaccharides and phenolic com-
100 0.6 pounds. In summary, H/C ratios “…with H2O” and “…with
TG 0-250 µm + H2O rPW” are not significantly different from those of typical fossil
TG 0-250 µm + rPW
TG rPW 0.5 coal structures. The H/C ratios vary between 1.0 and 1.3. The
80
dTG 0-250 µm locations of the HCs produced from BSG in Fig. 9 are close to
dTG 0-250 µm + rPW
dTG rPW 0.4
those of anaerobically digested maize silage, which also rep-
resents a residual biomass from industrial processing (having
-dm/dT [% K-1]
Mass loss [%]

60 0.3
different biochemical characteristics from BSG), once carbon-
ized at 230 °C for 6 h [145]. Typical ranges for low-rank
brown coal from Germany with a VM higher than 52% are
0.2
between 0.7 and 0.9 [146]. The H/C ratios of HCs produced
40
from “…with rPW” decrease to typical condensed coal struc-
0.1
tures with ratios between 0.5 and 0.7. The area of anthracites
with H/C ratios below 0.5 was not achieved.
20 0.0
200 400 600 800
Temperature [°C] 3.4 Hydrothermal carbonization of brewer’s spent
Fig. 10 Thermogravimetric analysis of the hydrochars from the particle grains and residual process water
size fraction 0–250 μm of BSG obtained after a reaction time of 30 min from 5-hydroxymethylfurfural synthesis
(on the top) and 120 min (on the bottom) either by addition of distilled
water or residual process water from HMF production. BSG = brewer’s
spent grains; H2O = water; rPW = residual process water from HMF
In this section, the results relevant to HTC runs with the addi-
production tion of rPW to BSG instead of water are described and
discussed. This rPW consists of organic compounds reactive
towards secondary char formation such as 47.8 C% HMF,
they might be adsorbed on the HC surface and therefore could 3.1 C% LA, and 0.8 C% FA as well as 34.8 C% of their
be easily extracted for refining [140–144]. precursor FRU (Table 3). More importantly, the supply of
As can be seen in the van Krevelen diagram in Fig. 9, H+ protons by remaining HNO3 influences mass transfer as
reflecting also the physicochemical characteristics illustrat- well as HTC reactions [47, 57, 83, 147]. H+ is expected to
ed in Table 2, the particle size fraction 500–850 μm has a promote a catalytic hydrolysis, usually achieved by the forma-
slightly lower H/C and significantly higher O/C atomic tion of AA via reverse aldol condensation or its addition as an
ratio, compared with the other particle size fractions. acid catalyst [3, 47, 62]. More precisely, it has a great catalytic
Dehydration and decarboxylation along with aldol effect on acetal hydrolysis [108]. On the one hand, this can be
condensation/addition and polymerization as well as aro- evaluated by comparing the thermographs of Fig. 10 taken
matization are the dominating reactions during this coali- from HCs produced from the particle size fraction 0–
fication, resulting in the release of H and O in the form of 250 μm for 30 min (on the top) and 120 min (on the bottom)
H2O and CO2 [40, 51, 122]. However, the van Krevelen in water and rPW. Since the peak between 230 and 300 °C in
diagram cannot display how complex the reactions are the dTG disappears, hydrolysis of polysaccharides such as
(aldol condensation/addition, polymerization, aromatization) hemicellulose but also hydrolysis of proteins to amino acids
on the pathway to secondary char. This emphasizes the by breaking peptide bonds took place [137]. Hydrolysis of
relevance of a more detailed liquid analysis to fully un- cellulose needed longer reaction time, which is confirmed by
derstand HC formation. the decreasing peak between 300 and 400 °C [148, 149].
Biomass Conv. Bioref.

0.6
Fig. 11 Dissolved organic carbon
(DOC) and concentration of
identified reaction products in the
HTC liquids of experiments with
brewer’s spent grains in residual 0.5

process water from 5-HMF pro-


duction in [mg L−1] for each par-
ticle size fraction related to the 1,2,4-benzenetriol

dissolved organic carbon (DOC) Ethanol

carbon percentage of DOC [C%]


0.4
Formic Acid
Levulinic Acid
Methylfurfural
0.3 Furfural
5-hydroxymethylfurfural
Lactic Acid
Glycolic Acid
0.2 Acetic Acid
Formaldehyde
Sucrose
Glucose
0.1 Fructose

0
0 30 60 120 0 30 60 120 0 30 60 120
0-250 µm 250-500 µm 500-850 µm
reaction time [min]

On the other hand, the results from an intensified hydro- continuous accumulation of FU-like compounds (nucle-
lysis can be retraced by comparison of carbon percentages ation), crosslinking via reactive O-containing functional
for SUC, FRU, and GLU in Figs. 6 and 11. However, the groups, appearing as oil droplets, and the subsequent poly-
difference between rPW and water as reaction medium, con- merization (particle growth) towards formation of carbon
sidering hydrolysis of (poly-)saccharides (Figs. 6 and 11), spheres (Fig. 12) [150]. Li et al. [151] and Luijkx et al.
carbon distribution (Figs. 7 and 13) as well as the carboni- [84] showed that phenolic compounds were also involved
zation degree (Fig. 9) is small, which was unexpected. It by reacting with FU-like compounds. However, BTO as the
becomes apparent from Fig. 12 that the progress in reactions rehydration product of HMF occurs with an increasing car-
towards HC up to 120 min depends on the particle size. The bon percentage in PW by enhancing the reaction time dur-
reason may be that the transfer of polysaccharide fragments ing HTC of the particle size fraction 500–850 μm (Fig. 11
through the tighter pores of larger BSG particles is limited and Table S3 - Supplementary materials).
due to quick condensation and polymerization of HMF mol- Taking the insights from the HTC of rPW into consider-
ecules on the BSG particles’ surface. Regions of particle ation, this compound seems to react due to its concentration
aggregates as well as films without defined structural fea- and the pH of the reaction environment, because of its chem-
tures, resulting from the degradation of GLU in the presence ical inertness [128]. Moreover, HMF is also rehydrated to LA
of HNO3 (pH = 1) according to Reiche et al. [20], and as and FA in a molar ratio greater than 1:1 till 120 min of reaction
they also appear in the SEM picture from HC “500-850 μm time, which implies an excess of FA again. Then, FA is obvi-
+ rPW for 30 min” in Fig. 12 may be responsible for clog- ously decarboxylated [20, 66, 88, 89]. Further decarboxyl-
ging of pores. The consequence is that the hydrolysis of ation reactions like that of LaA can also be verified not only
polysaccharide fragments and di-saccharides is reduced. by gas formation but also by the presence of EtOH that gets
Therefore, the supply of monosaccharides such as GLU oxidized again. Since we consider peak #11 in the chromato-
and FRU for dehydration to FU-like compounds is limited grams of Fig. S3 (Supplementary materials) as ProA, its de-
and it has to be considered that both are also involved in the hydration also takes place [126].
reverse aldol condensation to AA as well as FAD and GA As it can be seen from the SEM picture of HC “0–
[47, 62, 96]. 250 μm + rPW for 30 min” in Fig. 12, carbon spheres
Nevertheless, due to the addition of rPW, carbon percent- of different diameters accumulate and start fusing, as also
ages for HMF and FU are still around 20.0(± 3.0) and observed by Jung et al. [40] within short reaction times.
11.0(± 5.0) C% respectively in PW after the heating time The growth mechanism looks more advanced in compar-
for HTC of all particle size fractions. This allows for ison with the results obtained from the HTC of pure rPW
Biomass Conv. Bioref.

Fig. 12 Scanning electron


microscopy of hydrothermal
carbonized (reaction
temperature = 230 °C; reaction
time = 0–120 min) brewer’s spent
grains with a particle size of 0–
250 and 500–850 μm in residual
process water from synthesis of 5-
hydroxymethylfurfural; rPW =
residual process water from syn-
thesis of 5-hydroxymethylfurfural

(Fig. 5) or BSG in water (Fig. 8) for the same reaction the shapes of carbon spheres appear similar to those of HCs
time. However, according to Reiche et al. [20] in the case produced from XYL (FU) and GLU/FRU (HMF), whereas
of low pH and thus high H+ concentration, the internal they are embedded in an aromatic network [109, 130, 133].
condensation was faster than the formation of carbon This indicates intramolecular condensation and the formation
spheres resulting in a distribution of shapes (Fig. 5), or of an aromatic network starts forming after 30 min of reaction
in the case of pH > 3 and at intermediate H+ concentra- time, which is additionally confirmed by the constant carbon
tion, the polymerization was slowed down to enable the percentages of HC illustrated in Fig. 13. However, a slight
diffusion-controlled formation of a homogenous emulsion decline in carbon percentage of HC after 30 min reaction time
of carbon spheres. In other words, the latter results in is observed, which points out a limited but continuing transfer
carbon spheres of a larger diameter, which may also be of poly- and/ or di-saccharides through the biomass pores. As
influenced by dissolved salts, but with a less extended can be seen in Fig. 11 and Table S3 (Supplementary Materials)
conjugated network. The carbon spheres shown in the the release of GLU from hydrolysis fails to appear: This re-
SEM picture “0–250 μm + rPW for 120 min” have a ducing saccharide seems to react with present amino acids to
diameter around 10 μm which was so far obtained from N-containing heterocycles, which enhances the tertiary char
hydrothermal treatment of FRU at 180 °C under neutral fraction of the final HC. This becomes clear by comparison of
conditions, GLU at 220 °C in diluted HNO3 for longer the “not detected” species in Figs. 6 and 11. After 120 min, the
reaction times as well as a mixture of monosaccharides DOC is reduced to a similar value for water and rPW variants,
(fructose, xylose) with phenolic compounds (phenol, res- which proves that the usage of rPW in the HTC process is a
orcinol and phloroglucinol) [20, 72, 152]. suitable option with a positive impact on the HC formation.
Furthermore, the acid anion concentration reduces the re- Also, the composition of the PW is almost similar, showing
pulsion of formed carbon spheres, and as consequence, the that the rPW has been successfully recycled. As shown in
coalescence proceeds faster [40]. After 120 min of reaction Figs. 7 and 13, the carbon release as PG is even lower. The
time for both particle size fractions 0–250 and 500–850 μm, consequence is that the Ceff for the particle size fraction 0–
Biomass Conv. Bioref.

100 3 3 120
6 4 6
12 8 11
13 14 17
17
21 19 110
17
80 26 32 32
20 19 19 17
15 14

100
Percentage [%]

60

Ceff [%]
90

40
74 75 74
68 68 68 68 69 69 69 80
65 65

20
70

0 60

0-250 µm 250-500 µm 500-850 µm


reaction time [min]

C in HC C in PW C in PG Ceff

Fig. 13 Carbon distribution within the product phases after hydrothermal HC = hydrochar; PW = process water; PG = process gas; 0–250, 250–
carbonization of brewer’s spent grains at 230 °C for 0, 30, 60 and 120 min 500, 500–850 μm = particle sizes of brewer’s spent grains carbonized at
with residual process water from the synthesis of 5- 230 °C for 0, 30, 60 and 120 min in residual process water from HMF
hydroxymethylfurfural. The process gas is assumed as 100% CO2. production; Ceff = carbon retention efficiency

250 μm is enhanced by 31% and for both other fractions by recirculation of the PW from the HTC of miscanthus at
27% up to 120 min of reaction time in comparison with HTC 260 °C for 5 min which contained reactive dehydration prod-
in water. In total, 89 and 77% DOC of rPW is converted to ucts towards HC formation and provided acidic conditions
carbon in HC (Figs. 7 and 13). In addition to this, there is (pH = 2.68).
another positive aspect: while in the studies of Uddin et al.
[21], Stemann et al. [22], Weiner et al. [24] and Kabadayi et al.
[23] the HC yield increased by 5–10 wt.% by PW recircula- 4 Conclusion
tion, in this study, an increase in HC yield of 15–35 wt.%
(values not illustrated) was achieved as from 30 min of reac- This publication covers two aspects, the influence of the par-
tion time, albeit depending on the BSG particle size. ticles size of the feedstock and the addition of rPW from HMF
Additionally, the results from Kambo et al. [18], Weiner production which contains reactive intermediates towards sec-
et al. [24], and those from this study show that the quality in ondary char formation. Therefore, the substitution of water
terms of HC’s carbon percentage (Fig. 13) as well as its car- with rPW may enhance the efficiency of HTC of BSG.
bonization degree improved by PW addition (Fig. 9). HTC of the smallest particle size fraction 0–250 μm of
However, the PWs in the abovementioned publications were BSG in water shows the worst results regarding its carbon
recirculated from HTC of different biomasses. These were percentage in HC as well as Ceff due to a larger surface and
loblolly pine, poplar wood chips, miscanthus, unbleached tis- a greater reactivity. As a consequence, poly- and monosaccha-
sue paper and a mixture of grape pomace, orange pomace and rides which are relevant for Maillard reactions and thus for the
poultry litter, whose PWs mainly consisted of less reactive enhancement of HC yields by the formation of tertiary char are
compounds such as short-chain carboxylic acids (AA, LA, mainly degraded to carboxylic acids. This indicates that HTC
GA, FA). The consequence is that the research groups in- occurs via soluble intermediates and mainly results in carbon
creased the carbon contents but obtained lower mass yields spheres (secondary char). The particle size fraction of BSG
of HC due to intensified hydrolysis as well as reverse aldol (250–500 μm) with the greatest presence shows constant car-
condensation reactions [24]. The reason why Kambo et al. bon percentages in HC as well as a constant Ceff up to 120 min
[18] also achieved an improvement in HC quality was the of reaction time, even though rPW is added. The reason is that
Biomass Conv. Bioref.

the transfer of polysaccharide fragments through BSG particle 3. Lynam JG, Reza MT, Yan W, Vásquez VR, Coronella CJ (2015)
Hydrothermal carbonization of various lignocellulosic biomass.
pores and its hydrolysis to reducing saccharides is limited and
Biomass Conv Bioref 5(2):173–181
thus their supply is sustainable for reacting with amines. The 4. Sharples A (1957) The hydrolysis of cellulose and its relation to
result is a stable structure consisting of carbon spheres (sec- structure. Trans Faraday Soc 53:1003
ondary char) formed by aldol condensation/addition and N- 5. Vos KD, Burris FO, Riley RL (1966) Kinetic study of the hydro-
containing heterocycles (tertiary char) resulting from Maillard lysis of cellulose acetate in the pH range of 2–10. J Appl Polym
Sci 10(5):825–832
reactions embedded in an aromatic network (primary char) 6. Tolonen LK, Zuckerstätter G, Penttilä PA, Milacher W, Habicht
formed by a direct conversion of polysaccharide fragments. W, Serimaa R et al (2011) Structural changes in microcrystalline
Indeed, the variation in the composition of the particle size cellulose in subcritical water treatment. Biomacromolecules 12(7):
fractions seems negligible. Nevertheless, investigations with 2544–2551
7. García-Bordejé E, Pires E, Fraile JM (2017) Parametric study of
additional experimental series have to be continued because
the hydrothermal carbonization of cellulose and effect of acidic
Maillard reactions as well as their influence on yields and conditions. Carbon 123:421–432
characteristics of HCs do not seem completely clarified (ter- 8. Baugh KD, McCarty PL (1988) Thermochemical pretreatment of
tiary char). In particular, tests with reaction time higher than lignocellulose to enhance methane fermentation: I.
120 min could allow clarify this issue. In the case of fats, Monosaccharide and furfurals hydrothermal decomposition and
product formation rates. Biotechnol Bioeng 31(1):50–61
which are considered as less reactive under subcritical water 9. Qi Y, Song B, Qi Y (2016) The roles of formic acid and levulinic
conditions, investigations on their interactions in Maillard re- acid on the formation and growth of carbonaceous spheres by
actions become important. hydrothermal carbonization. RSC Adv 6(104):102428–102435
As a whole, the recirculation of PW is successful when 10. Sumerskii IV, Krutov SM, Zarubin MY (2010) Humin-like sub-
stances formed under the conditions of industrial hydrolysis of
dissolved saccharides and FU-like compounds are available wood. Russ J Appl Chem 83(2):320–327
since the beginning of the upcoming HTC run. There is a 11. Sevilla M, Fuertes AB (2009) Chemical and structural properties
positive impact on the conversion of the larger particle sizes of carbonaceous products obtained by hydrothermal carbonization
concerning their structural substances, starch, and extractives of saccharides. Chem Eur J 15(16):4195–4203
(proteins, fats), to a stable aromatic structure containing N- 12. Steinbach D, Kruse A, Sauer J (2017) Pretreatment technologies
of lignocellulosic biomass in water in view of furfural and 5-
heterocycles with an acceptable yield. The more DOC is re- hydroxymethylfurfural production- a review. Biomass Conv
duced from the recirculated PW and transferred to HC, the Bioref 7(2):247–274
more efficient the HTC runs and the less problematic the 13. Luo S, Zhou Y, Yi C, Luo Y, Fu J (2014) Influence of the feed
PW regarding its chemical oxygen demand for further moisture, rotor speed, and blades gap on the performances of a
biomass pulverization technology. TheScientificWorldJournal
treatments. 2014:435816
14. Cocero MJ, Cabeza Á, Abad N, Adamovic T, Vaquerizo L,
Acknowledgments The authors would like to thank the Martínez CM, Pazo-Cepeda MV (2018) Understanding biomass
students Johannes Winkler and Markus Götz from the University of fractionation in subcritical & supercritical water. J Supercrit Fluids
Hohenheim for their efforts in experimentation and analytics as well as 133:550–565
Sonja Habicht, Hermann Köhler, and Armin Lautenbach from the 15. van Putten R-J, van der Waal JC, de Jong E, Rasrendra CB, Heeres
Institute of Catalysis Research and Technology of the Karlsruhe HJ, de Vries JG (2013) Hydroxymethylfurfural, a versatile plat-
Institute of Technology for the technical support regarding the analysis form chemical made from renewable resources. Chem Rev 113(3):
of solid and liquid samples. 1499–1597
16. Rapp KM Process for preparing pure 5-hydroxymethylfurfuraldehyde:
Funding information The work was financially supported by the Federal United States patent
Ministry of Education and Research within the project Humboldt 17. van Zandvoort I, Wang Y, Rasrendra CB, Van Eck ERH,
Reloaded during the winter semester 2017/2018. Bruijnincx PCA, Heeres HJ et al (2013) Formation, molecular
structure, and morphology of humins in biomass conversion: in-
Compliance with ethical standards fluence of feedstock and processing conditions. ChemSusChem
6(9):1745–1758
18. Kambo HS, Minaret J, Dutta A (2018) Process water from the
Conflict of interests The authors declare that they have no conflicts of
hydrothermal carbonization of biomass: a waste or a valuable
interest.
product? Waste Biomass Valor 9(7):1181–1189
19. Tsilomelekis G, Orella MJ, Lin Z, Cheng Z, Zheng W, Nikolakis
V, Vlachos DG (2016) Molecular structure, morphology and
growth mechanisms and rates of 5-hydroxymethyl furfural
(HMF) derived humins. Green Chem 18(7):1983–1993
References 20. Reiche S, Kowalew N, Schlögl R (2015) Influence of synthesis
pH and oxidative strength of the catalyzing acid on the morphol-
1. Kruse A, Dahmen N (2018) Hydrothermal biomass conversion: ogy and chemical structure of hydrothermal carbon.
Quo vadis? J Supercrit Fluids 134:114–123 Chemphyschem 16(3):579–587
2. Kim D, Yoshikawa K, Park K (2015) Characteristics of biochar 21. Uddin MH, Reza MT, Lynam JG, Coronella CJ (2013) Effects of
obtained by hydrothermal carbonization of cellulose for renewable water recycling in hydrothermal carbonization of loblolly pine.
energy. Energies 8(12):14040–14048 Environ Prog Sustain Energy 34:n/a-n/a
Biomass Conv. Bioref.

22. Stemann J, Putschew A, Ziegler F (2013) Hydrothermal carboni- 43. Zakzeski J, Bruijnincx PCA, Jongerius AL, Weckhuysen BM
zation: process water characterization and effects of water recircu- (2010) The catalytic valorization of lignin for the production of
lation. Bioresour Technol 143:139–146 renewable chemicals. Chem Rev 110(6):3552–3599
23. Kabadayi Catalkopru A, Kantarli IC, Yanik J (2017) Effects of 44. Bhatty RS (1986) The potential of hull-less barley - a review.
spent liquor recirculation in hydrothermal carbonization. Cereal Chem 62:97–103
Bioresour Technol 226:89–93 45. Fincher GB (1975) Morphology and chemical composition of
24. Weiner B, Poerschmann J, Wedwitschka H, Koehler R, Kopinke barley endosperm cell walls. J Inst Brew 81(2):116–122
F-D (2014) Influence of process water reuse on the hydrothermal 46. Krawielitzki S, Kläusli TM (2015) Modified hydrothermal car-
carbonization of paper. ACS Sustain Chem Eng 2(9):2165–2171 bonization process for producing biobased 5-HMF platform
25. Poerschmann J, Weiner B, Koehler R, Kopinke F-D (2015) chemical. Ind Biotechnol 11(1):6–8
Organic breakdown products resulting from hydrothermal carbon- 47. Asghari SF, Yoshida H (2006) Acid-catalyzed production of 5-
ization of brewer’s spent grain. Chemosphere 131:71–77 hydroxymethyl furfural from d-fructose in subcritical water. Ind
26. Kruse A, Badoux F, Grandl R, Wüst D (2012) Hydrothermale Eng Chem Res 45(7):2163–2173
Karbonisierung: 2. Kinetik der Biertreber-Umwandlung Chemie 48. Kritzer P (1998) Die Korrosion der Nickel-Basis-Legierung 625
Ingenieur Technik 84(4):509–512 unter hydrothermalen Bedingungen: Einfluss der Parameter
27. Lynch KM, Steffen EJ, Arendt EK (2016) Brewers’ spent grain: a Temperatur, Druck, pH-Wert und Anwesenheit von Sauerstoff
review with an emphasis on food and health. J Inst Brew 122(4): sowie der Anionen Chlorid, Sulfat, Nitrat und Phosphat auf das
553–568 Korroionsverhalten
28. McCarthy AL, O'Callaghan YC, Piggott CO, FitzGerald RJ, 49. Wegertseder GmbH. Microsoft Word - TechDat.doc
O'Brien NM (2013) Brewers’ spent grain; bioactivity of phenolic 50. ARS. National Nutrient Database for Standard Reference
component, its role in animal nutrition and potential for incorpo- 51. Funke A, Ziegler F (2010) Hydrothermal carbonization of bio-
ration in functional foods: a review. Proc Nutr Soc 72(1):117–125 mass: a summary and discussion of chemical mechanisms for
29. Robertson JA, I'Anson KJA, Treimo J, Faulds CB, Brocklehurst process engineering. Biofuels Bioprod Bioref 4(2):160–177
TF, Eijsink VGH, Waldron KW (2010) Profiling brewers’ spent 52. Titirici M-M (ed) (2013) Sustainable carbon materials from hydro-
grain for composition and microbial ecology at the site of produc- thermal processes. Wiley, Hoboken
tion. LWT Food Sci Technol 43(6):890–896 53. Patil SKR, Lund CRF (2011) Formation and growth of humins via
30. Mussatto SI, Roberto IC (2006) Chemical characterization and aldol addition and condensation during acid-catalyzed conversion
liberation of pentose sugars from brewer's spent grain. J Chem of 5-hydroxymethylfurfural. Energy Fuel 25(10):4745–4755
Technol Biotechnol 81(3):268–274 54. Patil SKR, Heltzel J, Lund CRF (2012) Comparison of structural
31. Berge ND, Ro KS, Mao J, Flora JRV, Chappell MA, Bae S (2011) features of humins formed catalytically from glucose, fructose,
Hydrothermal carbonization of municipal waste streams. Environ and 5-hydroxymethylfurfuraldehyde. Energy Fuel 26(8):5281–
Sci Technol 45(13):5696–5703 5293
32. Dinjus E, Kruse A, Tröger N (2011) Hydrothermal carbonization - 55. Volpe M, Fiori L (2017) From olive waste to solid biofuel through
1. Influence of lignin in lignocelluloses. Chem Eng Technol hydrothermal carbonisation: the role of temperature and solid load
34(12):2037–2043 on secondary char formation and hydrochar energy properties. J
Anal Appl Pyrolysis 124:63–72
33. van Soest PJ, Robertson JB, Lewis BA (1991) Methods for dietary
56. Aida TM, Sato Y, Watanabe M, Tajima K, Nonaka T, Hattori H,
fiber, neutral detergent fiber, and nonstarch polysaccharides in
Arai K (2007) Dehydration of d-glucose in high temperature water
relation to animal nutrition. J Dairy Sci 74(10):3583–3597
at pressures up to 80MPa. J Supercrit Fluids 40(3):381–388
34. Muthusamy N (2014) Chemical composition of brewer’s spent
57. Möller M, Nilges P, Harnisch F, Schröder U (2011) Subcritical
grains - a review. Int J Sci Environ Technol (3):2109–2112
water as reaction environment: fundamentals of hydrothermal bio-
35. Fărcaş AC, Socaci SA, Dulf FV, Tofană M, Mudura E, Diaconeasa mass transformation. ChemSusChem 4(5):566–579
Z (2015) Volatile profile, fatty acids composition and total pheno-
58. Jin F, Zhou Z, Moriya T, Kishida H, Higashijima H, Enomoto H
lics content of brewers' spent grain by-product with potential use
(2005) Controlling hydrothermal reaction pathways to improve
in the development of new functional foods. J Cereal Sci 64:34–42
acetic acid production from carbohydrate biomass. Environ Sci
36. Poerschmann J, Weiner B, Wedwitschka H, Baskyr I, Koehler R, Technol 39(6):1893–1902
Kopinke F-D (2014) Characterization of biocoals and dissolved 59. Chuntanapum A, Matsumura Y (2009) Formation of tarry material
organic matter phases obtained upon hydrothermal carbonization from 5-HMF in subcritical and supercritical water. Ind Eng Chem
of brewer’s spent grain. Bioresour Technol 164:162–169 Res 48(22):9837–9846
37. Santos M, Jiménez JJ, Bartolomé B, Gómez-Cordovés C, del 60. Srokol Z, Bouche A-G, van Estrik A, Strik RCJ, Maschmeyer T,
Nozal MJ (2003) Variability of brewer’s spent grain within a Peters JA (2004) Hydrothermal upgrading of biomass to biofuel;
brewery. Food Chem 80(1):17–21 studies on some monosaccharide model compounds. Carbohydr
38. Hardwick WA (ed.). Handbook of Brewing; 1994 Res 339(10):1717–1726
39. Yang H, Wang G, Ding N, Yin C, Chen Y (2016) Size-controllable 61. Fang Y, Zeng X, Yan P, Jing Z, Jin F (2012) An acidic two-step
synthesis of carbon spheres with assistance of metal ions. Synth hydrothermal process to enhance acetic acid production from car-
Met 214:1–4 bohydrate biomass. Ind Eng Chem Res 51(12):4759–4763
40. Jung D, Zimmermann M, Kruse A (2018) Hydrothermal carbon- 62. Kabyemela BM, ADSCHIRI T, Malaluan RM, ARAI K (1999)
ization of fructose: growth mechanism and kinetic model. ACS Glucose and fructose decomposition in subcritical and supercriti-
Sustainable Chem Eng cal water: detailed reaction pathway, mechanisms, and kinetics.
41. Falco C, Baccile N, Titirici M-M (2011) Morphological and struc- Ind Eng Chem Res 38(8):2888–2895
tural differences between glucose, cellulose and lignocellulosic 63. Akgül G, Kruse A (2013) Hydrothermal disproportionation of form-
biomass derived hydrothermal carbons. Green Chem 13(11):3273 aldehyde at subcritical conditions. J Supercrit Fluids 73:43–50
42. Chen W-H, Ye S-C, Sheen H-K (2012) Hydrolysis characteristics 64. Koido K, Ishida Y, Kumabe K, Matsumoto K, Hasegawa T (2010)
of sugarcane bagasse pretreated by dilute acid solution in a micro- Kinetics of ethanol oxidation in subcritical water. J Supercrit
wave irradiation environment. Appl Energy 93:237–244 Fluids 55(1):246–251
Biomass Conv. Bioref.

65. Aydıncak K, Yumak T, Sınağ A, Esen B (2012) Synthesis and hydroxymethyl-2-furaldehyde and d-fructose. Carbohydr Res
characterization of carbonaceous materials from saccharides (glu- 242:131–139
cose and lactose) and two waste biomasses by hydrothermal car- 85. Kruse A, Gawlik A (2003) Biomass conversion in water at 330
bonization. Ind Eng Chem Res 51(26):9145–9152 −410 °C and 30−50 MPa. Identification of key compounds for
66. Girisuta B, Janssen LPBM, Heeres HJ (2006) A kinetic study on indicating different chemical reaction pathways. Ind Eng Chem
the decomposition of 5-hydroxymethylfurfural into levulinic acid. Res 42(2):267–279
Green Chem 8(8):701 86. Antal MJ, Mok WSL, Richards GN (1990) Four-carbon model
67. Gökkaya DS, Sağlam M, Yüksel M, Madenoğlu TG, Ballice L compounds for the reactions of sugars in water at high tempera-
(2016) Characterization of products evolved from supercritical ture. Carbohydr Res 199(1):111–115
water gasification of xylose (principal sugar in hemicellulose). 87. Horvat J, Klaic B, Metelko B, Sunjic V (1985) Mechanism of
Energy Sources Part A 38(11):1503–1511 levulinic acid formation. Tetrahedron Lett 26:2111–2114
68. Karagöz S, Bhaskar T, MUTO A, SAKATA Y (2005) 88. Flannelly T, Lopes M, Kupiainen L, Dooley S, Leahy JJ (2016)
Comparative studies of oil compositions produced from sawdust, Non-stoichiometric formation of formic and levulinic acids from
rice husk, lignin and cellulose by hydrothermal treatment. Fuel the hydrolysis of biomass derived hexose carbohydrates. RSC
84(7–8):875–884 Adv 6(7):5797–5804
69. Peterson AA, Vogel F, Lachance RP, Fröling M, Antal JMJ, Tester 89. Baccile N, Laurent G, Babonneau F, Fayon F, Titirici M-M,
JW (2008) Thermochemical biofuel production in hydrothermal Antonietti M (2009) Structural characterization of hydrothermal
media: a review of sub- and supercritical water technologies. carbon spheres by advanced solid-state MAS 13 C NMR investi-
Energy Environ Sci 1(1):32 gations. J Phys Chem C 113(22):9644–9654
70. Yang L, Tsilomelekis G, Caratzoulas S, Vlachos DG (2015) 90. Cheng Z, Everhart JL, Tsilomelekis G, Nikolakis V, Saha B,
Mechanism of Brønsted acid-catalyzed glucose dehydration. Vlachos DG (2018) Structural analysis of humins formed in the
ChemSusChem 8(8):1334–1341 Brønsted acid catalyzed dehydration of fructose. Green Chem
71. Zhuang X, Zhan H, Song Y, He C, Huang Y, Yin X, Wu C (2019) 20(5):997–1006
Insights into the evolution of chemical structures in lignocellulose 91. Li M, Li W, Liu S (2012) Control of the morphology and chemical
and non-lignocellulose biowastes during hydrothermal carboniza- properties of carbon spheres prepared from glucose by a hydro-
tion (HTC). Fuel 236:960–974 thermal method. J Mater Res 27(08):1117–1123
72. Ryu J, Suh Y-W, Suh DJ, Ahn, Dong JA (2010) Hydrothermal 92. Funke A, Reebs F, Kruse A (2013) Experimental comparison of
preparation of carbon microspheres from mono-saccharides and hydrothermal and vapothermal carbonization. Fuel Process
phenolic compounds. Carbon 48(7):1990–1998 Technol 115:261–269
73. Koch H, Pein J (1985) Condensation reactions between phenol,
93. Rasrendra CB, Windt M, Wang Y, Adisasmito S, Makertihartha
formaldehyde and 5-hydroxymethylfurfural, formed as intermedi-
IGBN, van Eck ERH, Meier D, Heeres HJ (2013) Experimental
ate in the acid catalyzed dehydration of starchy products. Polym
studies on the pyrolysis of humins from the acid-catalysed dehy-
Bull 13:525–532
dration of C6-sugars. J Anal Appl Pyrolysis 104:299–307
74. Lucian M, Volpe M, Fiori L (2019) Hydrothermal carbonization
94. Sykes P (ed) (1988) Reaktionsmechanismen der organischen
kinetics of lignocellulosic agro-wastes: experimental data and
Chemie: Eine Einführung, 9th edn. VCH, Weinheim
modeling. Energies 12(3):516
95. Castello D, Kruse A, Fiori L (2015) Low temperature supercritical
75. Dussan K, Girisuta B, Lopes M, Leahy JJ, Hayes MHB (2015)
water gasification of biomass constituents: glucose/phenol mix-
Conversion of hemicellulose sugars catalyzed by formic acid: ki-
tures. Biomass Bioenergy 73:84–94
netics of the dehydration of D-xylose, L-arabinose, and D-glucose.
ChemSusChem 8(8):1411–1428 96. Kruse A (2008) Supercritical water gasification. Biofuels Bioprod
76. Poerschmann J, Weiner B, Koehler R, Kopinke F-D (2017) Bioref 2:415–437
Hydrothermal carbonization of glucose, fructose, and xylose— 97. Chuntanapum A, Yong TL-K, Miyake S, Matsumura Y (2008)
identification of organic products with medium molecular masses. Behavior of 5-HMF in subcritical and supercritical water. Ind
ACS Sustain Chem Eng 5(8):6420–6428 Eng Chem Res 47(9):2956–2962
77. van Dam HE, Kieboom APG, van Bekkum H (1986) The conver- 98. Williams PT, Onwudili J (2005) Composition of products from the
sion of fructose and glucose in acidic media: formation of supercritical water gasification of glucose: a model biomass com-
hydroxymethylfurfural. Starch/Stärke 38(3):95–101 pound. Ind Eng Chem Res 44(23):8739–8749
78. Cottier L, Descotes G. 5-Hydroxymethylfurfural syntheses and 99. Nikolov PY, Yaylayan VA (2011) Thermal decomposition of 5-
chemical transformations 1991(2):233–248 (hydroxymethyl)-2-furaldehyde (HMF) and its further transforma-
79. Speck JC The Lobry De Bruyn-Alberda Van Ekenstein tions in the presence of glycine. J Agric Food Chem 59(18):
Transformation 13:63–103 10104–10113
80. Kuster BFM, Temmink HMG (1977) The influence of ph and 100. Gökkaya DS, Sağlam M, Yüksel M, Ballice L (2016)
weak-acid anions on the dehydration of d-fructose. Carbohydr Hydrothermal gasification of xylose: effects of reaction tempera-
Res 54:185–191 ture, pressure, and K2CO3 as a catalyst on product distribution.
81. Cao X, Peng X, Sun S, Zhong L, Chen W, Wang S, Sun RC (2015) Biomass Bioenergy 91:26–36
Hydrothermal conversion of xylose, glucose, and cellulose under 101. Rawn JD (ed) (2014) Organic chemistry: structure, mechanism
the catalysis of transition metal sulfates. Carbohydr Polym 118: and synthesis. Elsevier
44–51 102. Bicker M, Endres S, Ott L, Vogel H (2005) Catalytical conversion
82. Antal MJ, Mok WSL, Richards GN (1990) Mechanism of forma- of carbohydrates in subcritical water: a new chemical process for
tion of 5-(hydroxymethyl)-2-furaldehyde from d-fructose and su- lactic acid production. J Mol Catal A Chem 239(1–2):151–157
crose. Carbohydr Res 199(1):91–109 103. Hirosaka K, Koido K, Fukayama M, Ouryoji K, Hasegawa T
83. Körner P, Jung D, Kruse A (2018) The effect of different Brønsted (2008) Experimental and numerical study of ethanol oxidation in
acids on the hydrothermal conversion of fructose to HMF. Green sub-critical water. J Supercrit Fluids 44(3):347–355
Chem 20(10):2231–2241 104. Zhai Z, Li X, Tang C, Peng J, Jiang N, Bai W, Gao H, Liao Y
84. Luijkx GCA, van Rantwijk F, van Bekkum H (1993) (2014) Decarbonylation of lactic acid to acetaldehyde over alumi-
Hydrothermal formation of 1,2,4-benzenetriol from 5- num sulfate catalyst. Ind Eng Chem Res 53(25):10318–10327
Biomass Conv. Bioref.

105. Ma XJ, Yang XF, Zheng X, Lin L, Chen LH, Huang LL, Cao SL 124. Faboya OO (1990) The interaction between oxalic acid and diva-
(2014) Degradation and dissolution of hemicelluloses during bam- lent ions - Mg2+, Zn2+ and Ca2+ - in aqueous medium. Food
boo hydrothermal pretreatment. Bioresour Technol 161:215–220 Chem 38:179–187
106. Bajpai P (ed) Pretreatment of lignocellulosic biomass for biofuel 125. Bell DJ, Blake JD, Prazak M, Rowell D, Wilson PN (1991)
production: structure of lignocellulosic biomass. Springer, Studies on yeast differentiation using organic acid metabolites part
Singapore 1. Development of methodology using highperformance liquid
107. Román S, Nabais JMV, Laginhas C, Ledesma B, González JF chromatography. J Inst Brew 97:297–305
(2012) Hydrothermal carbonization as an effective way of densi- 126. Jin F, Zhou Z, Enomoto H, Moriya T, Higashijima H (2004)
fying the energy content of biomass. Fuel Process Technol 103: Conversion mechanism of cellulosic biomass to lactic acid in sub-
78–83 critical water and acid–base catalytic effect of subcritical water.
108. Bobleter O (1994) Hydrothermal degradation of polymers derived Chem Lett 33(2):126–127
from plants. Prog Polym Sci 19(5):797–841 127. Li Y, Lu X, Yuan L, Liu X (2009) Fructose decomposition kinetics
109. Reza MT, Yan W, Uddin MH, Lynam JG, Hoekman SK, in organic acids-enriched high temperature liquid water. Biomass
Coronella CJ, Vásquez VR (2013) Reaction kinetics of hydrother- Bioenergy 33(9):1182–1187
mal carbonization of loblolly pine. Bioresour Technol 139:161– 128. Kumalaputri AJ, Randolph C, Otten E, Heeres HJ, Deuss PJ
169 (2018) Lewis acid catalyzed conversion of 5-
110. Rissanen JV, Grénman H, Xu C, Willför S, Murzin DY, Salmi T hydroxymethylfurfural to 1,2,4-benzenetriol, an overlooked
(2014) Obtaining spruce hemicelluloses of desired molar mass by biobased compound. ACS Sustain Chem Eng 6(3):3419–3425
using pressurized hot water extraction. ChemSusChem 7(10): 129. Rodríguez Correa C, Stollovsky M, Hehr T, Rauscher Y, Rolli B,
2947–2953 Kruse A (2017) Influence of the carbonization process on activat-
111. Lucian M, Fiori L (2017) Hydrothermal carbonization of waste ed carbon properties from lignin and lignin-rich biomasses. ACS
biomass: process design, modeling, energy efficiency and cost Sustain Chem Eng 5(9):8222–8233
analysis. Energies 10(2):211 130. Kang S, Li X, Fan J, Chang J (2012) Characterization of
112. Yan W, Hoekman SK, Broch A, Coronella CJ (2014) Effect of hydrochars produced by hydrothermal carbonization of lignin,
hydrothermal carbonization reaction parameters on the properties cellulose, d-xylose, and wood meal. Ind Eng Chem Res 51(26):
of hydrochar and pellets. Environ Prog Sustain Energy 33(3):676– 9023–9031
680 131. Ravber M (2015) Hydrothermal degradation of fats, carbohydrates
113. Abdilla RM, Rasrendra CB, Heeres HJ (2018) Kinetic studies on and proteins in sunflower seeds after treatment with subcritical
the conversion of Levoglucosan to glucose in water using water. Chem Biochem Eng Q 29(3):351–355
Brønsted acids as the catalysts. Ind Eng Chem Res 57(9):3204–
132. Abdelmoez W, Nakahasi T, Yoshida H (2007) Amino acid trans-
3214
formation and decomposition in saturated subcritical water condi-
114. Zan Y, Sun Y, Kong L, Miao G, Bao L, Wang H, Li S, Sun Y
tions. Ind Eng Chem Res 46(16):5286–5294
(2018) Formic acid-induced controlled-release hydrolysis of
133. Titirici M-M, Antonietti M, Baccile N (2008) Hydrothermal car-
microalgae (Scenedesmus) to lactic acid over Sn-beta catalyst.
bon from biomass: a comparison of the local structure from poly-
ChemSusChem 11(15):2492–2496
to monosaccharides and pentoses/hexoses. Green Chem 10(11):
115. Cabeza A, Piqueras CM, Sobrón F, García-Serna J (2016)
1204
Modeling of biomass fractionation in a lab-scale biorefinery: sol-
134. Peterson AA, Lachance RP, Tester JW (2010) Kinetic evidence of
ubilization of hemicellulose and cellulose from holm oak wood
the Maillard reaction in hydrothermal biomass processing: glucose
using subcritical water. Bioresour Technol 200:90–102
−glycine interactions in high-temperature, high-pressure water.
116. Choudhary V, Mushrif SH, Ho C, Anderko A, Nikolakis V,
Ind Eng Chem Res 49(5):2107–2117
Marinkovic NS, Frenkel AI, Sandler SI, Vlachos DG (2013)
Insights into the interplay of Lewis and Brønsted acid catalysts 135. Fan Y, Hornung U, Dahmen N, Kruse A. Formation of N-
in glucose and fructose conversion to 5-(hydroxymethyl)furfural containing heterocycles from hydrothermal liquefaction of modell
and levulinic acid in aqueous media. J Am Chem Soc 135(10): compounds and sewage sludge. Proceedings of the 24th European
3997–4006 Biomass Conference 2017
117. Klingler D, Berg J, Vogel H (2007) Hydrothermal reactions of 136. Fan Y, Hornung U, Dahmen N, Kruse A (2018) Hydrothermal
alanine and glycine in sub- and supercritical water. J Supercrit liquefaction of protein-containing biomass: study of model com-
Fluids 43(1):112–119 pounds for Maillard reactions. Biomass Conv Bioref 8(4):909–
118. Sato N, Quitain AT, Kang K, Daimon H, Fujie K (2004) Reaction 923
kinetics of amino acid decomposition in high-temperature and 137. Wang T, Zhai Y, Zhu Y, Peng C, Xu B, Wang T, Li C, Zeng G
high-pressure water. Ind Eng Chem Res 43(13):3217–3222 (2018) Influence of temperature on nitrogen fate during hydrother-
119. Piqueras CM, Cabeza Á, Gallina G, Cantero DA, García-Serna J, mal carbonization of food waste. Bioresour Technol 247:182–189
Cocero MJ (2017) Online integrated fractionation-hydrolysis of 138. Teri G, Luo L, Savage PE (2014) Hydrothermal treatment of pro-
lignocellulosic biomass using sub- and supercritical water. Chem tein, polysaccharide, and lipids alone and in mixtures. Energy Fuel
Eng J 308:110–125 28(12):7501–7509
120. BIO-RAD. guide to Aminex HPLC columns for food and bever- 139. Qiao L, Chen J, Ying Y, Zheng J-W, Jiang L (2013) Influence of
age, biotechnology, and bio-organic analysis. MP 2016;82(1) NH4 + on the preparation of carbonaceous spheres by a hydro-
121. Novotný O, Cejpek K, Velíšek J (2008) Formation of carboxylic thermal process. J Mater Sci 48(9):3341–3346
acids during degradation of monosaccharides. Czech JFood Sci 140. Biller P, Riley R, Ross AB (2011) Catalytic hydrothermal process-
26(2):113–131 ing of microalgae: decomposition and upgrading of lipids.
122. Smith AM, Singh S, Ross AB Fate of inorganic material during Bioresour Technol 102(7):4841–4848
hydrothermal carbonisation of biomass: influence of feedstock on 141. Broch A, Hoekman S, Jena U, Langford J (2014) Analysis of solid
combustion behaviour of hydrochar. Fuel 169:135–145 and aqueous phase products from hydrothermal carbonization of
123. Volpe M, Wüst D, Merzari F, Lucian M, Andreottola G, Kruse A, whole and lipid-extracted algae. Energies 7(1):62–79
Fiori L (2018) One stage olive mill waste streams valorisation via 142. Heilmann SM, Jader LR, Harned LA, Sadowsky MJ, Schendel FJ,
hydrothermal carbonisation. Waste Manag 80:224–234 Lefebvre PA, von Keitz MG, Valentas KJ (2011) Hydrothermal
Biomass Conv. Bioref.

carbonization of microalgae II. Fatty acid, char, and algal nutrient 148. Simsir H, Eltugral N, Karagöz S (2017) Hydrothermal carboniza-
products. Appl Energy 88(10):3286–3290 tion for the preparation of hydrochars from glucose, cellulose,
143. Lucian M, Volpe M, Gao L, Piro G, Goldfarb JL, Fiori L (2018) chitin, chitosan and wood chips via low-temperature and their
Impact of hydrothermal carbonization conditions on the formation characterization. Bioresour Technol 246:82–87
of hydrochars and secondary chars from the organic fraction of 149. Álvarez-Murillo A, Sabio E, Ledesma B, Román S, González-
municipal solid waste. Fuel 233:257–268 García CM (2016) Generation of biofuel from hydrothermal car-
144. Gao L, Volpe M, Lucian M, Fiori L, Goldfarb JL (2019) Does bonization of cellulose. Kinetics modelling. Energy 94:600–608
hydrothermal carbonization as a biomass pretreatment reduce fuel 150. Liu X, Song P, Hou J, Wang B, Xu F, Zhang X (2017) Revealing
segregation of coal-biomass blends during oxidation? Energy the dynamic formation process and mechanism of hollow carbon
Convers Manag 181:93–104 spheres: from bowl to sphere. ACS Sustain Chem Eng 6(2):2797–
145. Mumme J, Eckervogt L, Pielert J, Diakité M, Rupp F, Kern J 2805
(2011) Hydrothermal carbonization of anaerobically digested 151. Li R, Wang L, Shahbazi A (2015) A review of hydrothermal
maize silage. Bioresour Technol 102(19):9255–9260 carbonization of carbohydrates for carbon spheres preparation.
146. Rathsack P, Reichel D, Krzack S, Otto M (2014) Komprehensive Tr Ren Energy 1(1):43–56
Gaschromatographie-Massenspektrometrie von Alkylbenzolen in 152. Zhang M, Yang H, Liu Y, Sun X, Zhang D, Xue D (2012)
Pyrolyseölen aus Biomasse und Kohle. Chem Ing Tech 86(10): Hydrophobic precipitation of carbonaceous spheres from fructose
1779–1789 by a hydrothermal process. Carbon 50(6):2155–2161
147. Oka H, Yamago S, Yoshida J, Kajimoto O (2002) Evidence for a
hydroxide ion catalyzed pathway in ester hydrolysis in supercrit- Publisher’s note Springer Nature remains neutral with regard to jurisdic-
ical water. Angew Chem Int Ed 41(4):623–625 tional claims in published maps and institutional affiliations.

You might also like