Hydrothermal Carbonization
Hydrothermal Carbonization
Hydrothermal Carbonization
Hydrothermal Carbonization
Edited by
M. Toufiq Reza
www.mdpi.com/journal/energies
Hydrothermal Carbonization
Hydrothermal Carbonization
Editor
M. Toufiq Reza
Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland
This is a reprint of articles from the Special Issue published online in the open access journal
Energies (ISSN 1996-1073) (available at: https://www.mdpi.com/journal/energies/special issues/
hydrothermal carbonization).
For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:
Lastname, A.A.; Lastname, B.B. Article Title. Journal Name Year, Volume Number, Page Range.
© 2023 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license. The book as a whole is distributed by MDPI under the terms
and conditions of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND)
license.
Contents
M. Toufiq Reza
Hydrothermal Carbonization
Reprinted from: Energies 2022, 15, 5491, doi:10.3390/en15155491 . . . . . . . . . . . . . . . . . . . 1
Fabio Merzari, Jillian Goldfarb, Gianni Andreottola, Tanja Mimmo, Maurizio Volpe and
Luca Fiori
Hydrothermal Carbonization as a Strategy for Sewage Sludge Management: Influence of
Process Withdrawal Point on Hydrochar Properties
Reprinted from: Energies 2020, 13, 2890, doi:10.3390/en13112890 . . . . . . . . . . . . . . . . . . . 31
Louise Delahaye, John Thomas Hobson, Matthew Peter Rando, Brenna Sweeney, Avery
Bernard Brown, Geoffrey Allen Tompsett, et al.
Experimental and Computational Evaluation of Heavy Metal Cation Adsorption for Molecular
Design of Hydrothermal Char
Reprinted from: Energies 2020, 13, 4203, doi:10.3390/en13164203 . . . . . . . . . . . . . . . . . . . 53
Md Tahmid Islam, Nepu Saha, Sergio Hernandez, Jordan Klinger and M. Toufiq Reza
Integration of Air Classification and Hydrothermal Carbonization to Enhance Energy Recovery
of Corn Stover
Reprinted from: Energies 2021, 14, 1397, doi:10.3390/en14051397 . . . . . . . . . . . . . . . . . . . 121
Gabriel Gerner, Luca Meyer, Rahel Wanner, Thomas Keller and Rolf Krebs
Sewage Sludge Treatment by Hydrothermal Carbonization: Feasibility Study for Sustainable
Nutrient Recovery and Fuel Production
Reprinted from: Energies 2021, 14, 2697, doi:10.3390/en14092697 . . . . . . . . . . . . . . . . . . . 135
v
About the Editor
M. Toufiq Reza
M. Toufiq Reza, Ph.D., is an Assistant Professor of Chemical Engineering at the Department of
Biomedical and Chemical Engineering and Sciences of Florida Institute of Technology. He received
his M.S. and Ph.D. in Chemical Engineering from the University of Nevada, Reno, in 2011 and 2013,
respectively. He has published more than ninety peer-reviewed journal articles, five patents, two book
chapters, and more than one hundred oral and poster presentations. Dr. Reza was named as one of
recipients of the I&EC Research 2021 Class of Influential Researchers. He has also contributed to the
AIChE Journal 2022 Futures Issue. He was the recipient of the 2022 New Holland Young Researcher
award from American Society of Agricultural and Biological Engineers (ASABE) and 2022 FASABE
Young Researcher Award. He is also listed in the “World Ranking of Top 2% Scientists” in the 2021
and 2022 database published by Stanford University. Dr. Reza leads the Biofuels Research Lab at
Florida Tech, with primary research interest on upcycling wastes via hydrothermal carbonization.
His current research projects have been funded by the National Science Foundation, U.S. Department
of Agriculture, U.S. Environmental Protection Agency, American Chemical Society, Florida Sea Grant,
U.S. Agency for International Development, Sugarbush Foundation, and other agencies. Dr. Reza is a
member of the American Institute of Chemical Engineers, American Chemical Society, and American
Society of Agricultural and Biological Engineers. He is the secretary of PRS-707 Food & Organic
Waste Management & Utilization at the ASABE. He is an Associate Editor of Frontiers in Fuels, a
Section Editor of the Energies journal, and an Editorial Board member of the Biomass Conversion and
Biorefinery journal.
vii
energies
Editorial
Hydrothermal Carbonization
M. Toufiq Reza
Department of Biomedical and Chemical Engineering and Sciences, Florida Institute of Technology,
150 West University Blvd, Melbourne, FL 32901, USA; treza@fit.edu
Over the past decade, hydrothermal carbonization (HTC) has emerged as a promising
thermochemical pathway for treating and converting wet wastes into fuel, materials, and
chemicals. Many of the earlier studies have been carried out to understand HTC reaction
mechanisms and reaction kinetics using model compounds [1,2]. More recently, HTC
studies have shifted from model compounds to researching lignocellulosic biomasses [3–5];
however, these works remain focused on solid biofuel production in the form of hydrochar.
Perhaps the real potential for HTC lies in wet wastes (such as sewage sludge, agricultural
wastes, animal wastes, etc.) and their conversion to fertilizer, high value materials, and
sustainable chemicals [1]. However, the kinetics and reaction mechanism of the real wet
wastes often deviate from those of the model compounds [6]. Therefore, studying the
reaction kinetics of real wet waste is an important step towards revealing the viability of
HTC in the commercial process.
This Special Issue titled, “Hydrothermal Carbonization” has invited researchers from
around the World to contribute to these issues in the field of HTC research. Including
the editorial, ten papers are listed in this Special Issue. Researchers from eight different
countries have contributed their research to this Special Issue. The highlights of the Special
Issue are as follows:
• Maniscalco et al. [7] reviewed the recent progress made towards using HTC as a
valuable tool for energy and environmental applications. The review identified that
pristine hydrochar from organic wastes such as food waste are suitable for combustion
or soil amendment based on feedstock. However, the review also reported recent pro-
gresses with respect to further activating hydrochar for energy storage and pollutant
Citation: Reza, M.T. Hydrothermal
adsorption applications.
Carbonization. Energies 2022, 15, 5491.
• Three articles are reported aspects of HTC related to sewage sludge. Merzari et al. [8]
https://doi.org/10.3390/en15155491
have evaluated HTC systems at various points in wastewater treatment plants to treat
Received: 19 July 2022 various type of sludges (e.g., primary, secondary, and digestate sludge) for different
Accepted: 27 July 2022 applications. Meanwhile, Luhmann and With have optimized the dewaterability and
Published: 29 July 2022 phosphorus release from sewage sludge during HTC [9]. Sewage sludge is one of
Publisher’s Note: MDPI stays neutral the major sources of phosphorus and recovering phosphorus effectively from sewage
with regard to jurisdictional claims in sludge may enhance the success of HTC. Finally, Gerner et al. [10] have reported
published maps and institutional affil- the nutrient recovery and fuel production potentials from sewage sludge. The study
iations. reveals that the HTC process could be much cheaper than the current practice of
sewage sludge treatment in Switzerland; therefore, fuel production and nutrient
recovery from sewage sludge are economically viable.
• Ro et al. [11] have compared the HTC process with vapothermal carbonization (VTC).
Copyright: © 2022 by the author. When feedstock does not have an adequate volume of water to create subcritical water
Licensee MDPI, Basel, Switzerland. conditions, the thermochemical conversion is called VTC. There are distinct product
This article is an open access article formations for HTC and VTC [12]. This study reveals the minimum amount of water
distributed under the terms and
required in the feedstock to be considered conducive to HTC. This fundamental study
conditions of the Creative Commons
will clarify future research directions and ensure the practice of HTC.
Attribution (CC BY) license (https://
• Islam et al. [13] reported an approach integrating air classification with HTC. In
creativecommons.org/licenses/by/
the study, air classification separated high-ash-fraction (waste) corn stover from the
4.0/).
light ash fraction type [14]. This paper reported that the HTC of waste products can
be performed and the hydrochar can be used to enhance the pelletization of clean
corn stover.
• Gonnella et al. [15] have studied the thermal analysis and kinetics modelling of both
the pyrolysis and oxidation of hydrochar. Although hydrochar has been reported as
solid fuel, few sources have reported the combustion of hydrochar.
• Delahaye et al. [16] evaluated the cationic heavy metal adsorption of hydrochar both
experimentally and computationally. Confirmed via density functional theory (DFT)
simulation, this fundamental study reveals that hydrochar possesses superior binding
sites for copper (II) ions compared to ion exchange resins.
• Finally, Hasan et al. [17] reported how various precursors affect the carbon quantum
dot (CQD) properties via HTC. CQDs are very high-value materials with a variety
of applications, such as bioimaging, photocatalysis, and energy storage. This paper
demonstrates the possibility of separating CQDs from HTC process liquid. The utiliza-
tion of HTC process liquid has always been a concern for HTC commercialization and
this paper reveals a potential application for this very liquid.
Overall, HTC has shown tremendous potential toward commercialization. We hope
that readers will find this Special Issue informative and inspiring. The editor is thankful
to the authors for submitting their HTC research. The Energies editors and the review-
ers are also acknowledged for reviewing the manuscripts, which have made significant
improvements to this Special Issue.
Funding: The work is partially funded by USDA NIFA AFRI grant no. 2019-67019-31594 and grant
no. 2021-67022-34487.
Conflicts of Interest: The author declares no conflict of interest.
References
1. Román, S.; Libra, J.; Berge, N.; Sabio, E.; Ro, K.; Li, L.; Ledesma, B.; Álvarez, A.; Bae, S. Hydrothermal Carbonization: Modeling,
Final Properties Design and Applications: A Review. Energies 2018, 11, 216. [CrossRef]
2. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.-M.; Fühner, C.; Bens, O.; Kern, J.; et al.
Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry, processes and applications of wet and
dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
3. Islam, M.T.; Sultana, A.I.; Saha, N.; Klinger, J.L.; Reza, M.T. Pretreatment of Biomass by Selected Type-III Deep Eutectic Solvents
and Evaluation of the Pretreatment Effects on Hydrothermal Carbonization. Ind. Eng. Chem. Res. 2021, 60, 15479–15491.
[CrossRef]
4. Reza, M.T.; Yang, X.; Coronella, C.J.; Lin, H.; Hathwaik, U.; Shintani, D.; Neupane, B.P.; Miller, G.C. Hydrothermal Carbonization
(HTC) and Pelletization of Two Arid Land Plants Bagasse for Energy Densification. ACS Sustain. Chem. Eng. 2015, 4, 1106–1114.
[CrossRef]
5. Saba, A.; Saha, P.; Reza, M.T. Co-Hydrothermal Carbonization of coal-biomass blend: Influence of temperature on solid fuel
properties. Fuel Process. Technol. 2017, 167, 711–720. [CrossRef]
6. Reza, M.T.; Yan, W.; Uddin, M.H.; Lynam, J.G.; Hoekman, S.K.; Coronella, C.J.; Vasquez, V.R. Reaction kinetics of hydrothermal
carbonization of loblolly pine. Bioresour. Technol. 2013, 139, 161–169. [CrossRef] [PubMed]
7. Maniscalco, M.P.; Volpe, M.; Messineo, A. Hydrothermal Carbonization as a Valuable Tool for Energy and Environmental
Applications: A Review. Energies 2020, 13, 4098. [CrossRef]
8. Merzari, F.; Goldfarb, J.; Andreottola, G.; Mimmo, T.; Volpe, M.; Fiori, L. Hydrothermal Carbonization as a Strategy for Sewage
Sludge Management: Influence of Process Withdrawal Point on Hydrochar Properties. Energies 2020, 13, 2890. [CrossRef]
9. Lühmann, T.; Wirth, B. Sewage Sludge Valorization via Hydrothermal Carbonization: Optimizing Dewaterability and Phosphorus
Release. Energies 2020, 13, 4417. [CrossRef]
10. Gerner, G.; Meyer, L.; Wanner, R.; Keller, T.; Krebs, R. Sewage Sludge Treatment by Hydrothermal Carbonization: Feasibility
Study for Sustainable Nutrient Recovery and Fuel Production. Energies 2021, 14, 2697. [CrossRef]
11. Ro, K.S.; Libra, J.A. Alvarez-Murillo, Comparative Studies on Water- and Vapor-Based Hydrothermal Carbonization: Process
Analysis. Energies 2020, 13, 5733. [CrossRef]
12. Funke, A.; Reebs, F.; Kruse, A. Experimental comparison of hydrothermal and vapothermal carbonization. Fuel Process. Technol.
2013, 115, 261–269. [CrossRef]
13. Islam, M.T.; Saha, N.; Hernandez, S.; Klinger, J.; Reza, M.T. Integration of Air Classification and Hydrothermal Carbonization to
Enhance Energy Recovery of Corn Stover. Energies 2021, 14, 1397. [CrossRef]
2
Energies 2022, 15, 5491
14. Lacey, J.A.; Aston, J.E.; Westover, T.L.; Cherry, R.S.; Thompson, D.N. Removal of introduced inorganic content from chipped
forest residues via air classification. Fuel 2015, 160, 265–273. [CrossRef]
15. Gonnella, G.; Ischia, G.; Fambri, L.; Fiori, L. Thermal Analysis and Kinetic Modeling of Pyrolysis and Oxidation of Hydrochars.
Energies 2022, 15, 950. [CrossRef]
16. Delahaye, L.; Hobson, J.T.; Rando, M.P.; Sweeney, B.; Brown, A.B.; Tompsett, G.A.; Ates, A.; Deskins, N.A.; Timko, M.T.
Experimental and Computational Evaluation of Heavy Metal Cation Adsorption for Molecular Design of Hydrothermal Char.
Energies 2020, 13, 4203. [CrossRef]
17. Hasan, M.R.; Saha, N.; Quaid, T.; Reza, M.T. Formation of Carbon Quantum Dots via Hydrothermal Carbonization: Investigate
the Effect of Precursors. Energies 2021, 14, 986. [CrossRef]
3
energies
Review
Hydrothermal Carbonization as a Valuable Tool for
Energy and Environmental Applications: A Review
Manfredi Picciotto Maniscalco, Maurizio Volpe and Antonio Messineo *
Faculty of Engineering and Architecture, Kore University of Enna, Cittadella Universitaria, 94100 Enna, Italy;
manfredi.maniscalco@unikore.it (M.P.M.); maurizio.volpe@unikore.it (M.V.)
* Correspondence: antonio.messineo@unikore.it
Keywords: HTC; waste biomass; energy recovery; environmental remediation; nutrients recovery;
activated carbon
1. Introduction
The increasing need of finding new renewable energy alternatives to fossil fuels together with
the need to safely dispose of organic waste has pushed, in the last few years, the investigation for
more efficient and reliable technologies for waste biomass energy exploitation and conversion towards
valuable materials. Indeed, the constant rise in the global energy consumption together with the
parallel decrease in the fossil fuel reservoirs, has driven the scientific community toward the research of
more sustainable sources [1]. Waste biomass, and in particular the organic fraction of municipal solid
waste (OFMSW) and industrial and sewage sludges are among the most studied residual feedstocks to
produce energy and valuable carbonaceous materials [2–6].
One of the major drawbacks of choosing waste biomass for energy production, is the need for
one or more pre-treatments before use [7]. The low energy density due to the high moisture content
of waste biomass, makes it often not suitable for a direct energy conversion by combustion and/or
gasification [8–10]. The high moisture in residual biomass raises significantly the operational cost of
transportation and energy consumption for drying [11]. However, due to its great abundance, globally
availability, carbon neutrality and the necessity to find efficient and valuable technologies to treat
and convert it into sustainable energy sources and exploitable materials, waste biomasses has gained
increasing attention and investigation [12].
Between different valuable alternative treatments for waste biomass, wet thermochemical
technologies for their flexibility and, principle, their low investment and operating costs are receiving
5
Energies 2020, 13, 4098
increasing attention in the last years [13,14]. In particular, hydrothermal carbonization (HTC),
also known as wet pyrolysis, is a thermochemical process performed in the presence of sub-critical
water at temperature usually between 180 and 280 ◦ C and autogenous saturated vapor conditions
(10–80 bars). Residence time is varied from minutes up to several hours [3,15,16]. At high temperature
and pressure, even at subcritical conditions, water undergoes a dramatic properties change acting more
as an organic solvent and its increased ion product favors reactions that are typically catalyzed by acids
or bases, promoting biomass decomposition through hydrolysis, dehydration and decarboxylation
reactions [17,18]. By means of these reactions, it is possible to increase the carbon content of the initial
feedstock, by removing most of the more volatile oxygenated compounds (furans and low molecular
fatty acids) that are typically moved to the aqueous phase [19]. It is also well documented in the
literature that, during HTC, part of the inorganics are moved to the liquid phase [19–21] thus, on the
one hand, reducing in the solid residue (hydrochar), minerals and ash formation during combustion,
and on the other hand, increasing the high heating values (HHV) [17,22]. This behavior leads to
a reduced risk of slugging and fouling phenomena during combustion of hydrochars in boilers if
compared to the use of pyrochars [2,23,24]. Hydrochar can be used in numerous applications, as a bio
fuel for energy production [1,13,25], as a source of carbon and nutrients in soil application [19,26,27] or
even as starting material for further advanced utilization in supercapacitors or a porous matrix for the
adsorption of pollutants [28,29]. Moreover, the possible scalability of the process up to the industrial
level [30], as well as relatively mild process conditions, make HTC a valid solution for the near future
for the transformation of a wide range of raw biomass into valuable materials [13,30]. In Figure 1
is reported a flowchart of hydrochar production from waste biomass and its possible energetic and
environmental applications.
Since HTC is performed under wet conditions, this technology can find its best application for
biomasses that present high initial moisture content. Indeed, when dealing with direct combustion,
a detrimental overall efficiency was shown when the starting biomass presents high moisture
content [10]. The main reaction pathways involved in the hydrothermal carbonization process
are still under discussion among the scientific community [31]. One possible route was discussed
by Jatzwauck and Schumpe [32], according to which the HTC process is essentially divided into
three steps. The first one is mainly related to the hydrolyzation of cellulose, hemicellulose and part
of the lignin. During the second step, the generated intermediates endure further re-arrangement
resulting in the generation of gaseous products. As an alternative to the formation of the gas phase,
intermediate products can conglomerate to form secondary char, during the third step. The other
mostly discussed route, regarding the evolution of HTC reactions, involves two main pathways:
a solid–solid transformation, according to which hydrochar is directly generated from the dehydration
of the initial biomass, and a vapor–solid reaction in which the dissolved organics present in the solution,
back polymerize to form again a solid material, often named coke or secondary char [33–35].
The present work attempted to review the most relevant research on HTC of the last 5 years,
in order to derive the best operative conditions for the production of valuable hydrochar materials.
Moreover, whenever possible, the present review was integrated with a further pyrolysis or gasification
process in order to evaluate how HTC behaves as a thermo-chemical pre-treatment for the production
of porous material or precursor for energy application purposes.
6
Energies 2020, 13, 4098
7
Energies 2020, 13, 4098
by Titirici [31]. According to the author, cellulose starts to be converted above 210 ◦ C according to
two different pathways, as depicted in Figure 2 [31]. In their experiments, Olszewski et al. [43] and
Volpe et al. [40] reported a complete cellulose conversion at a reaction temperature of 260 ◦ C.
Due to its very complex structure, lignin undergoes decomposition under a wider temperature
range with respect to hemicellulose and cellulose [44]. The deterioration starts at around 200 ◦ C, leading
to the disruption of highly reactive alkyl-aryl-ether bonds, which further react with other intermediate
compounds to form highly stable products [45]. Volpe et al. [19], when performing thermos gravimetric
analysis (TGA) on olive mill waste hydrochar, found that while cellulose peak is significantly reduced
for HTC temperature of 250 ◦ C, the peak associated to lignin increases [44]. Through the same analysis,
authors also confirmed the production of lower molecular compounds and/or tarry products coming
from the decomposition of lignin. Indeed, a new peak in TGA curve at around 250 ◦ C was found for
the sample produced at 220 and 250 ◦ C. From these considerations, it can be deduced that different
reaction temperatures lead to different compositions in terms of cellulose, hemicellulose and lignin.
The study conducted by Kim et al. [46] reported the energy retention efficiency (ERE) for each of these
components in relation to reaction temperature. Indeed, as previously reported, as HTC temperature is
raised, the calorific value increases as well, but the mass yield is reduced. It is therefore of paramount
importance to define the best conditions to maximize the ERE, in order to obtain a valuable biochar for
energy purposes. The results of this study showed that ERE for lignin and cellulose was maximized,
respectively, at 200 ◦ C and 220 ◦ C, while the best temperature for the conversion of biomass into high
energetic medium appeared to be 220 ◦ C.
The higher thermal stability of lignocellulosic against non-lignocellulosic biomass was also
confirmed by Zhuang et al. [47], testing herb tea waste (lignocellulosic) and penicillin mycelial
waste (non-lignocellulosic). The analysis performed on the hydrochars produced showed a superior
thermal stability for the tea waste, due to the presence of hemicellulose and lignin which start to
degrade above 210 ◦ C, while polysaccharide and protein present in the penicillin mycelial waste were
hydrolyzed at temperatures lower than 180 ◦ C. Finally, some considerations should be addressed
regarding the influence of pressure, particle size, and the possibility of water recirculation within
the system. With respect to the first parameter, the autogenous pressure that is developed inside the
reactor maintains the water in the liquid state, favoring its ability to dissolve polar compounds [48].
Higher pressure levels result in smaller hydrochar dimensions and higher pore volume, together with
a mild increase in the HHV. Only when pressure is raised above 100 MPa, significant increases in the
HHV can be achieved, but the higher cost of the reactor can limit the application [48]. Regarding
particle size and water recirculation, Heidari et al. [49] found that when biomass dimension was raised
from <0.25 mm to within the range of 1.25–3 mm, a slight increase in the mass yield was observed,
8
Energies 2020, 13, 4098
but the calorific value was lower. This condition is explained by the more severe conversion that
biomass endures since water can more easily penetrate and reach the inner core of the biomass particle.
Water recirculation affects almost only the mass yield, increasing it after the first recycle of about 12%,
while the contribution to HHV is almost negligible, since it was raised by just 2%.
Proximate and ultimate analysis have been widely used to characterize the hydrochars properties.
Proximate analysis is used to get an overall picture of the char composition in terms of fixed carbon
(FC), volatile matter (VC), ash and moisture content (M), while the ultimate analysis give the weight
percentages of their major elements (C, H, N, S, O). Although not all the reactions involved in HTC
treatment are well known, the general evolution pathways of the main biomass elements are fairly well
understood. As long as reaction severity is increased, most of the volatile compounds are released,
increasing carbon content and heating value. The main reactions occurring are dehydration and
de-hydrogenation which imply, respectively, the removal of hydroxyl groups and the rupture of the
of the long chain carboxylic acids [50]. Different studies evaluated the effect of residence time and
reaction temperature on the final hydrochar yields and properties [4,17,51,52]. The results of the studies
assert that the main contribution to biomass degradation and increase in the calorific value is reaction
temperature rather than residence time [53]. Lucian et al. [51] reported an increase in the heating value
of olive trimming from 22.6 to 27.8 MJ kg−1 when reaction temperature was raised from 180 to 250 ◦ C.
At the same time, increasing the residence time from 1 to 6 h led to a maximum increase in HHV of
+2 MJ kg−1 , at the temperature of 220 ◦ C, passing from 24.7 to 26.7 MJ kg−1 . The increase in the calorific
value is mainly due to the combined effect of the removal of oxygen and volatile compounds and the
parallel rise in carbon content. The evolution of the concentration of carbon, oxygen and hydrogen
is usually represented through the van Krevelen diagram in H/C and O/C ratios which are plotted.
It is fairly stated that an increase in the hydrothermal carbonization temperature boosts dehydration
reactions, removing oxygen from the initial biomass structure, resulting in a decrease in the O/C ratio.
The evolution of the hydrochar composition is represented in the van Krevelen diagram through
a shift toward lower values of O/C, characteristic of highly carbonaceous fuels like lignite or coals.
Ulbrich et al. stated that the removal of oxygen for low HTC temperature is only mildly affected by
residence time, while in the temperature range of 230–280 ◦ C a more pronounced connection between
oxygen reduction and residence time was observed [52]. Figure 3 displays the van Krevelen diagrams
for three biomass sources reported in the literature: a sugar rich waste, namely rotten apples and
coming from the fruit industry, a lignocellulosic residue (olive trimming) and a protein based biomass
(Penicillin mycelial waste) [1,51,54]. It can clearly be seen that each waste endures a specific evolution
pathway, in relation to its initial structure and composition. Evidences of the different transformations
that biomass endures during HTC, in relation to its origin process parameters, are reported in Table 1.
Lignocellulosic material, like olive trimming, presents higher thermal stability than proteins and
polysaccharides, resulting in higher mass yield (approx. 50%) and fixed carbon (9%) with respect to
fruit and protein-based wastes.
Hydrochar produced from fruit wastes resulted in an increase in fixed carbon as long as reaction
temperature is raised, due to the conversion of glucose which starts at around 225 ◦ C [1]. Different
research groups showed that mass yield is mainly affected by reaction temperature, rather than
residence time. For low reaction temperature, residence time can even have a positive effect on the
final mass yield. Chen et al. [55], as well as Lucian et al. [51], reported an increase in mass yield when
reaction temperature was set respectively at 190 and 180 ◦ C for a reaction time of 6 h, when compared
to mass yields obtained at 1-h reaction time. The increase in hydrochar mass is associated with a
re-arrangement and back polymerization of organic compounds from liquid to solid phase [55].
9
Energies 2020, 13, 4098
Figure 3. van Krevelen diagram of hydrothermal carbonization (HTC) treatment of rotten apples,
Penicillin mycelial and olive trimming.
As previously reported, the last parameter that can be varied when performing HTC treatment is
the biomass-to-water ratio (B/W). Volpe et al. [17] found that when biomass is increased in the mix up to
a B/W ratio equal to 0.2, the share of secondary char production is increased as well, boosting the HHV
of the hydrochar. In terms of ash content, HTC presents a huge advantage over other pre-treatment
process such as pyrolysis. In fact, although during “dry” pre-treatment all the inorganics are kept
inside the biomass structure, increasing their content in the final product due to the mass loss during
the heating, in HTC some of the ash forming elements are removed from the biomass and washed
away with the process water [24].
10
Table 1. Effect of HTC reaction parameters on hydrochar composition, calorific value and mass yield, together with coal properties as reference.
FC VM C H O HHVraw HTC Res. Time FC_HTC VM_HTC C_HTC H_HTC O_HTC HHV_HC MY Ash_HC
Biomass Ref.
[%] [%] [%] [%] [%] [MJ/kg] Temp [◦ C] [h] [%] [%] [%] [%] [%] [MJ/kg] [%] [%]
Pine 13.1 87.8 48.1 6.6 45.1 19.2 220 4.0 30.8 68.8 66.0 5.9 27.5 25.9 76 0.4 [56]
Straw 14.8 80.9 48.2 6.5 40.6 18.2 220 4.0 25.2 72.6 57.9 5.0 34.4 22.4 56 2.2 [56]
150 0.5 15.3 68.8 54.9 6.9 34.8 19.8 84 15.8
Energies 2020, 13, 4098
11
200 4.0 23.0 70.5 61.2 7.1 27.2 26.9 63 4.1
220 4.0 29.6 66.2 67.1 6.9 21.1 29.3 55 4.2
180 2.0 22.4 74.2 60.2 7.4 28.3 26.7 66 3.4
200 2.0 25.1 71.7 62.3 7.2 26.2 27.5 62 3.2
Brewer spent grains 220 2.0 28.7 68.0 66.5 7.3 21.9 29.6 52 3.2
16 76.2 51.2 6.9 36.5 22.3 [4]
(90% moisture) 180 4.0 23.7 73.1 59.9 7.0 29.1 26.2 65 3.1
200 4.0 26.5 70.3 63.5 7.2 24.9 28.0 60 3.2
220 4.0 31.9 64.8 66.6 6. 9 21.8 29.1 51 3.3
190 0.25 34.2 65.5 62.4 5.4 30.8 24.8 36 0.3
Rotten apples 14.8 83.6 43.5 6.2 47.5 17.5 225 0.25 35.9 63.9 64.2 5.3 29.4 25.7 36 0.2 [1]
260 0.25 37.9 61.6 66.8 5.0 26.8 26.0 39 0.4
Table 1. Cont.
FC VM C H O HHVraw HTC Res. Time FC_HTC VM_HTC C_HTC H_HTC O_HTC HHV_HC MY Ash_HC
Biomass Ref.
[%] [%] [%] [%] [%] [MJ/kg] Temp [◦ C] [h] [%] [%] [%] [%] [%] [MJ/kg] [%] [%]
190 0.25 28.8 68.6 55.7 5.5 34.5 21.8 38 2.6
Grape pomace 11.6 87.5 44.1 6.2 41.9 17.5 225 0.25 35.4 62.8 61.4 5.1 29.9 24.5 40 1.7 [1]
260 0.25 35.0 60.6 64.9 5.0 24.9 24.8 45 4.3
Energies 2020, 13, 4098
12
Energies 2020, 13, 4098
13
Energies 2020, 13, 4098
experiments with food waste, proposed a simulation model of a plant processing 1 ton of fresh refuse
per day, resulting in a positive net energy balance for the whole process.
Food waste in the form of digestate from biogas plants, was experimented by Cao et al. [80].
The digestion significantly reduced the potential for energy conversion, since it converted a great share
of the carbonaceous content to methane and carbon oxides and increased the ash amount with respect
to the raw material. HTC treatment did not show significant improvement in terms of combustion
potential. Poor energetic properties were also reported in [81] when dealing with the wet fraction of
municipal solid waste digestate, but mostly due to the scarce potential of the initial feedstock.
A good potential for solid fuel conversion was proposed in [82,83] for the lignocellulosic material.
In [82] authors tested the efficacy of a 220 ◦ C, 90 min HTC treatment on six different biomasses, namely:
olive pomace, walnut shell, hazelnut shell, apricot seed, tea stalk and wood sawdust. Each of the
biomasses studied, presented better combustion properties with respect to the initial feedstock, in terms
of heating value, carbon content and ignition conditions. The best energetic properties were found to
be related to the olive pomace with an increase in the HHV of up to 25.6 MJ kg-1 and an ash content of
5.5%, while hazelnut husks showed the poorest properties with a HHV of 20.6 MJ kg−1 and an ash
percentage of approximately 12%. Moreover, all the hydrochar presented higher ignition temperatures
with respect to the raw biomasses, which significantly improve the handling and storage properties
of the fuels, reducing the risk of self-ignition and combustion. Similar properties of the obtained
hydrochar were also reported by Chen et al. when dealing with HTC of sweet potato peels [84].
14
Energies 2020, 13, 4098
treatment on the final product, a blank assay was also conducted. The results showed a remarkable
increase in surface area and pore development, as well as in the capacitance properties of the hydrochar.
Melamine for hydrochar activation was also studied by Sevilla et al. [91], mixing it with K2 C2 O4 and
the hydrochar obtained from eucalyptus sawdust produced at 250 ◦ C for 4 h. The use of K2 C2 O4 ,
instead of the more frequently used potassium hydroxide, according to the authors, is due to three main
reasons: its lower corrosive potential, the higher obtainable product yield and its less disruptive action
on the morphology of chars. The use of K2 C2 O4 instead of KOH can almost double the char production,
for the same amount of reactant adopted, while increasing at the same time the electrical properties.
Hydrochar from pine cones, hemp waste, tobacco rods, peony pollen and argy worm-wood were
successfully tested as precursors for supercapacitors application through KOH activation [63,92–95].
Excellent capacitance properties were found by Zhao et al. [63] when working with hydrothermally
carbonized tobacco rods. In this work, authors firstly treated the biomass through HTC at 200 ◦ C for
12 h and then activated it at 800 ◦ C for 1 h with a char-to-KOH ratio of 1:3 in mass. Results showed
that the activation stage increased the specific surface area up to 2115 m2 g−1 , raising significantly
also the specific capacitance. Indeed, the produced electrodes showed superb electrical properties
with a capacitance as high as 286 and 212 F g−1 , when the current density was, respectively, 0.5 and
30 A g−1 . Higher capacitance was recorded when using malva nut as biomass precursor by Ye et al. [96].
HTC was performed at 200 ◦ C for 18 h and hydrochars were further activated at 700 ◦ C in a ratio of 3:1
with KOH. Capacitance up to 279 and 219 F g−1 were achieved for current densities of 1 and 20 A g−1 ,
respectively. Acid HTC with KOH activation resulted in lower performance when pine cones were
used as starting biomass [93]. Manyala et al. tested the addition of 0.5 ml of sulfuric acid into 80 ml of
de ionized (DI) water, as medium for the HTC reaction. The further activation was performed for 1 h at
the temperatures of 600, 700, 800 and 900 ◦ C, with a KOH-to-char ratio of 1:1 (w/w). When temperature
reached 900 ◦ C a significant reduction in porosity and electrical properties of the char was observed.
This condition is probably due to pore walls collapsing at higher temperature, which enlarge their
average dimensions, reducing at the same time the amount of micropores which act as a storage site
during charging. A wider inspection of the possible activating agent was conducted by Jain et al. [97]
on coconut shells. Biomass was firstly treated in autoclave with H2 O2 or ZnCl2 , testing different
dosages and temperatures, and further activated under CO2 at 800 ◦ C for 2 h. Three different HTC
conditions were tested in this work:
Each set of tests was further proposed under a wide range of biomass concentrations in the
solution, ranging from 0.05 to 0.66 mg L−1 . For the case of ZnCl2 -to-shell ratio of 1:1, authors reported
a decrease in the specific surface area as long as biomass concentration was raised, passing from
1700 m2 g −1 for a biomass concentration of 0.05 mg L−1 , to 1350 m2 g −1 when the concentration was
raised to 0.22 mg L−1 . The tests using H2 O2 presented a similar trend, with an increase in surface area
from 1750 to 2450 m2 g−1 when the biomass concentration was raised from 0.16 to 0.5 mg L−1 , but a
further increase up to 0.66 mg L−1 led to an almost 20% reduction in specific surface area. In terms of
electrical properties, the highest energy density as well as the highest capacitance was found to be
related to carbon produced with H2 O2 and ZnCl2 with a biomass concentration of 0.5 mg L−1 , with a
capacitance of 246 and 221 F g−1 for a current density of 0.25 and 5 A g−1 , respectively. Impressive
performances were also reported in [98] when testing ZnCl2 -activated hydrocar produced from Coca
Cola® . Authors found one of the highest capacitances ever recorded for waste precursors, equal to
352.7 F g−1 for a current density of 1 A g−1 . Table 2 summarizes some of the studies reported above
with also some reference values for synthetized materials used for supercapacitor applications.
15
Energies 2020, 13, 4098
Aside from applications for supercapacitors, hydrochar has been successfully used for the
production of electrode material in lithium or sodium ion batteries as well as in the vanadium redox
ones [101–104]. In [101], authors used corn stalk as substrate for the production of a bio-based
lamellar molybdenum disulfide electrode to be used as anode in a lithium ion battery. To obtain the
desired material, a thiourea and ammonium molybdate solution was used as reaction medium for the
hydrothermal carbonization of cornstalks. Hydrochars were further treated in a furnace at 1000 ◦ C in
N2 atmosphere and then placed into an ammonium chloride solution at 60 ◦ C, to separate the excess of
calcium. The obtained material presented a discharge capacity, after 100 cycles, for a current density of
0.1 and 1 A g−1 of, respectively, 1129 and 339 mAh g−1 . A lithium ion battery anode was also produced
from a cellulose-derived carbon nanosphere obtained from corn straw by Yu et al. [102]. A sulfuric
acid bath and a following sodium hydroxide bath were used to extract cellulose from corn straw.
The obtained products were hydrothermally carbonized at 200 ◦ C for 24, 36, 48 and 60 h before being
further carbonized at 600 ◦ C in Ar, to obtain carbon nanospheres. Authors found that if reaction time
is too long, it can induce particles agglomeration, reducing the available specific area and, therefore,
the storage capacity. The best carbonization time was found to be 36 h, which showed a specific
discharge capacity after 100 cycles of 577 mA g−1 for a current density of 74 mA g−1 . Aside from lithium
ion, also applications for sodium ion batteries were studied. In [103], authors tested lath-shaped carbon
made through HTC followed by a 800 ◦ C carbonization step produced from peanut shells, to be used as
anode in Sodium ion battery. In this case, the discharge capacity was found to be equal to 265 mA g−1
at 30 mA g−1 after 100 cycles. Moreover, tests for a possible cathode material were performed by
Palomares et al. [105], using waste from vine shoots and eucalyptus wood as precursor to realize a
sodium vanadium fluorophosphate electrode. Authors found that when hydrochar is further subjected
to a flash thermal treatment (700 ◦ C, 10 min in N2 ), its electrochemical properties reached a specific
capacity of more than twofold that of pristine hydrochar.
Promising results were also achieved in vanadium redox flow batteries when testing activated
hydrochar [104] as electrode material.
4. Environmental Remediation
16
Energies 2020, 13, 4098
Two studies involving the use of digestate material as precursor for the realization of activated
carbon for the adsorption of contaminants were found in the literature [107,108]. The research
conducted by Bernardo et al. [108] on the use of hydrochar from digestate of municipal solid waste,
showed a promising perspective for its application on the removal of phosphate from waste water.
Digestate was hydrothermally carbonized at 250 ◦ C for 1 h, in acid (with addition of H2 SO4 ) or native
conditions, before being activated at 600 ◦ C for 2 h with a KOH-to-hydrochar ratio of 3:1 in mass.
Despite the reduction in specific surface area when biomass is subjected to acid treatment, the two
sets of chars present the same adsorption capacity of 12 mg of orthophosphate for each gram of
carbon. Results indicate that morphological properties do not play a significant role in orthophosphate
adsorption. According to the authors, the presence of Al3+ , Ca2+ , Fe2+/3+ and Mg2+ ions in the acid
hydrochar, can boost the formation of mineral complexes with phosphate ions, reducing its content.
Similar behaviors were also reported in [109–111], when operating with bio-chars for the removal of
phosphate from aqueous streams.
Hydrochars were also successfully used to remove CO2 from gaseous streams by using
agro-industrial waste or Coca Cola® [98,107,112]. Bio chars from agro-industrial waste were produced
at different temperatures (190–250 ◦ C), carbonization times (3 and 6 h) and pH levels (5 and 7),
in order to evaluate the optimal reaction conditions [107]. Activated carbons were then realized by
mixing hydrocarbons (HCs) with KOH in a mass ratio of 1:4 and then heated up to 600 ◦ C for 2 h.
Activation considerably boosted surface area and micro-meso porosity of the material. When testing the
adsorption capacity, authors found a much greater affinity toward CO2 than CH4 , with an adsorption
capacity of 8.8 molCO2 kg−1 , for the carbon produced at 250 ◦ C for 6 h at a pH of 5. Tests on the
adsorption of CO2 were also performed with activated carbons realized from the HTC of garlic peel
and Coca Cola® [98,112]. Hydrochar from garlic peels [112] were activated under different activation
temperatures (600, 700 and 800 ◦ C) and KOH-to-char ratios (0:1; 2:1 and 4:1) in order to investigate their
effect on porosity development and adsorption capacity. Tests revealed that an increase in activation
temperature and KOH amount led to a rise in specific surface area and pore volume. However,
CO2 adsorption is mainly driven by micropores availability, resulting in lower adsorption capacity
as long as temperature and potassium hydroxide increases. Indeed, too severe activation conditions
can induce pore enlargement, as well as pore walls collapse, reducing the possible sites for CO2
immobilization. Superior performances were achieved when using Coca Cola® as precursor in the
HTC treatment [98]. The work conducted by Boyjoo et al. showed outstanding capacity when Coca
Cola® was hydrothermally carbonized at 200 ◦ C for 4 h and further activated either with ZnCl2 or
KOH. This latter condition induced the highest increment in the adsorption ability, leading to an
adsorption capacity of 5.22 mmol g−1 at 25 ◦ C and 1 Atm, which is one of the highest ever recorded for
biomass precursors.
Bernardo et al. produced activated carbon by hydrothermal carbonization of digested sludge
and tested their activity toward phosphorous adsorption. They demonstrated that the high porosity
together with a high concentration of cations as in Ca, Al, Fe etc., favored phosphates removal from
wastewater [108].
Experiments for heavy metals removal were also conducted starting from agro-industrial wastes
as reported in [113–115]. In [114], authors tested the removal capacity of Antimony (III) and Cadmium
(II) on pyro-char and hydrochar realized from animal manure. The results of the adsorption tests
showed higher yields when pyrolytic chars were used, as well as an increase in the removal capacity as
long as pH was raised from 3 to 6. The total adsorption potentials for Antimony were 2.24–3.98 and
4.44–16.28 mg g−1 , respectively, for hydrochar and pyrochar, while higher capacities were reported
for the case of Cadmium, achieving a removal of 19.80–27.18 and 33.48–81.32 mg g−1 . Aside from
Cadmium removal, promising results on the adsorption of heavy metals contaminants such as
lead (Pb) [116,117], copper (Cu) [118–120], Zinc (Zn) [55,113,115] as well as pharmaceutical and
chemical waste were obtained in different researches [106,121–124]. For copper removal, different
studies proved that surface charge, pH and adsorbent dosage were the main parameters affecting
17
Energies 2020, 13, 4098
the adsorption capacity, while for lead superior efficiency (>99.5%) was achieved with Ni/Fe-doped
hydrochar [125]. Additionally, organic compounds like methylene blue, methyl orange and Congo
red were removed from contaminated solutions through the adsorption on hydrochar produced from
biomass waste [126–132]. Up to 655.7 mg of methylene blue for each gram of hydrochar was absorbed
when bamboo sawdust was carbonized into 1 M hydrochloric acid solution, before being further
treated with NaOH for 1 h. Acid modification led to 100%–200% increase in the adsorption capacity
when the solution was kept in the pH range of 10–12 [127].
The removal of Congo red and 2-naptol was studied by Li et al. in two following works,
using bamboo sawdust as starting material [133,134]. Both the two pollutants can have toxic effects on
both aquatic life and human organisms and, therefore, need to be removed from the water stream.
Moreover, due to their recalcitrant behavior to biological and thermal treatment, the use of adsorbent is
more and more studied. In both the studies published by the authors, they found that pore development
and the presence of oxygen functional groups on the surface can positively influence the adsorption
capacity of the activated carbon. HTC reaction was conducted according to three main procedures:
with pure water, with acid medium (HNO3 , H2 SO4 and H3 PO4 ) or through a two-stage operation with
a first hydrothermal carbonization in a 5% weight NaOH solution followed by an acid one in HNO3 ,
H2 SO4 or H3 PO4 . The two-stage process was used to perform a first delignification in order to favor
the penetration of the acid inside the material and enhance pore development. The results showed
that, among all, the best adsorption performances were achieved in a two-stage process, with H2 SO4
or H3 PO4 as acid medium and were equal to 90.51 and 72.93 mg g−1 , respectively, for Congo red
and 2-naphtol.
Adsorption of crystal violet (C25 H30 IN3 ) and malachite green was also investigated in
biomass-derived hydrochar in [135–137]. For both contaminants, pH represents a key factor for
adsorption since it is responsible for the surface charge modification of the hydrochars and the ionic
charge modification of the contaminants. It was found that for crystal violet, the optimum pH was 10,
while for malachite green it was found to be 7. Food leftovers were also transformed in adsorbent
material to remove rare earth ions [138] and toluene [139] from contaminated streams.
Magnetic modified hydrochar, produced from different biomasses through the addition of iron
compounds in the reaction media, were also used to remove tetracycline, roxarsone and persistent
free radicals from waste water [140–142]. The magnetic properties of the obtained material can reduce
the whole purification process time, by shortening the time needed to remove the adsorbent from
the solution and saving costs of operation. In Table 3, some of the presented works are reviewed,
according to the starting biomass, HTC condition and contaminant removed.
Table 3. Biomass type and HTC condition for removal of contaminants from waste waters.
To sum up, the last five years of studies on the possible use of hydrochar material as precursor for
the realization of the adsorption of different contaminants, showed interesting potential for the further
development of substrates with enhanced capacities that could represent, in the near future, cheaper
and more environmentally friendly options with respect to artificially synthetized materials.
18
Energies 2020, 13, 4098
19
Energies 2020, 13, 4098
with the contaminated soil in a proportion equal to 1, 2.5 and 5% by weight to evaluate the effect on
leaching and contaminants immobilization. Results of the trials showed a remarkable increase in the
removal of contaminants as long as the share of hydrochar was raised, achieving an improvement
in the immobilization efficiency of 95.1% (Pb) and 64.4% (Cd), with respect to pristine hydrochar.
Indeed, as long as hydrochar amount and lime content were raised, pH of soil shifted toward basic
values, increasing electrostatic interaction between positively charged metals ions and the anions in
hydrochar. Moreover, the formation of precipitates like metal (hydr)oxide or carbonate at high soil pH
helped to further reduce the amount of Pb and Cd ions in the samples. Copper was also successfully
removed from contaminated soil by Xia et al. [172], by using amino-functionalized hydrochar derived
from pinewood sawdust. Authors reported an efficiency in Cu removal from the soil equal to 96.2%,
together with a reduction of 98.1% in Cu amount in the leachate.
6. Conclusions
HTC represents a valid pre-treatment technology to convert waste biomass into new valuable
products which could find applications in a wide range of fields, from energy production to environment
remediation and soil conditioning. Since HTC operates in aqueous medium, it is perfectly suitable to
convert wet biomass, without any drying step. The most common waste biomass materials include
those produced by agro-industries, municipalities (organic fraction of municipal solid waste and
sewage sludge) but also from forestry and paper mill industries. The removal of part of the volatile
compounds and ashes increases the combustion properties of the hydrochar, making them suitable for
direct combustion in energy production applications. Chemical and physical activation, by using KOH
or CO2 , respectively, were successfully proposed to increase the surface area and internal porosity
of the material. By modifying the morphology and chemical surface, activated hydrochar can be
successfully used for the production of energy storage devices for the immobilization of pollutants
in gaseous or liquid streams. Pores development as well as elemental and surface composition can
be tailored by changing the carbonization and activation parameters. Finally, the addition in soil can
induce important benefits on pant growth, increasing water retention and reducing at the same time
possible leaching of contaminants into ground water.
Recently, some researchers focused their attention on how to mitigate the energetic impact of
the hydrothermal carbonization process by coupling it with renewable energy production systems.
Indeed, biochar can represent a valuable product to increase the efficiency of co-digestion systems,
with respect to its pristine composition. HC produced from coffee spent, rice straw or microalgae,
as well as the liquid phase produced from the hydrothermal carbonization, induced higher biogas
20
Energies 2020, 13, 4098
production as well as an increased methane content [184–188]. In such a way, the energy content of the
produced biogas can be used to run the HTC reactor, ensuring cleaner combustion with respect to the
hydrochar itself. Another option that has recently been studied is to couple the HTC reactor with a
solar concentration system which can satisfy the thermal load. Linear parabolic and parabolic disc
concentrators ensured a carbonization temperature up to 200 and 250 ◦ C respectively, ensuring heating
conditions of traditional systems [189,190].
Nevertheless, there are still some mechanisms in the HTC process that need further study to be
completely understood, in order to gain a stronger control over the process parameters, their influence
on the final product properties and thus their possible applications. Depending on the HTC operating
conditions and the nature of the feedstock, the technology still presents some limitations for its full
development. Between them, the correct process liquid treatment and management often present
challenges due to the presence of toxic hydrocarbons (such as polycyclic aromatic hydrocarbons (PHAs)
and high heavy metals concentrations. The presence of toxic compounds severely limits the possibility
to use such residue (as for example in agriculture as a possible source of nutrients) increasing the
overall costs of the HTC process. Investigations related to toxic hydrocarbons removal and heavy
metals recovery represent a necessary step for further development of the HTC technology.
Author Contributions: Conceptualization, M.P.M., M.V. and A.M.; Writing-Original Draft Preparation, M.P.M.;
Writing-Review & Editing, M.P.M., M.V. and A.M.; Supervision & final review, A.M. All authors have read and
agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Zhang, B.; Heidari, M.; Regmi, B.; Salaudeen, S.; Arku, P.; Thimmannagari, M.; Dutta, A. Hydrothermal
carbonization of fruit wastes: A promising technique for generating hydrochar. Energies 2018, 11, 2022.
[CrossRef]
2. Volpe, M.; Fiori, L.; Volpe, R.; Messineo, A. Upgrading of olive tree trimmings residue as biofuel by
hydrothermal carbonization and torrefaction: A comparative study. In Chemical Engineering Transactions;
AIDIC: Milan, Italy, 2016; Volume 50, pp. 113–118. ISBN 9788895608419.
3. Lucian, M.; Volpe, M.; Gao, L.; Piro, G.; Goldfarb, J.L.; Fiori, L. Impact of hydrothermal carbonization
conditions on the formation of hydrochars and secondary chars from the organic fraction of municipal solid
waste. Fuel 2018, 233, 257–268. [CrossRef]
4. Arauzo, P.J.; Olszewski, M.P.; Kruse, A. Hydrothermal carbonization brewer’s spent grains with the focus on
improving the degradation of the feedstock. Energies 2018, 11, 3226. [CrossRef]
5. Ferrentino, R.; Ceccato, R.; Marchetti, V.; Andreottola, G. Sewage Sludge Hydrochar: An Option for Removal
of Methylene Blue from Wastewater. Appl. Sci. 2020, 10, 3445. [CrossRef]
6. Gopu, C.; Gao, L.; Volpe, M.; Fiori, L.; Goldfarb, J.L. Valorizing municipal solid waste: Waste to energy and
activated carbons for water treatment via pyrolysis. J. Anal. Appl. Pyrolysis 2018, 133, 48–58. [CrossRef]
7. Cavalaglio, G.; Coccia, V.; Cotana, F.; Gelosia, M.; Nicolini, A.; Petrozzi, A. Energy from poultry waste:
An Aspen Plus-based approach to the thermo-chemical processes. Waste Manag. 2018, 73, 496–503. [CrossRef]
[PubMed]
8. Prins, M.J.; Ptasinski, K.J.; Janssen, F.J.J.G. More efficient biomass gasification via torrefaction. Energy 2006,
31, 3458–3470. [CrossRef]
9. Messineo, A.; Ciulla, G.; Messineo, S.; Volpe, M.; Volpe, R. Evaluation of equilibrium moisture content in
ligno-cellulosic residues of olive culture. ARPN J. Eng. Appl. Sci. 2014, 9, 5–11.
10. Heidari, M.; Salaudeen, S.; Norouzi, O.; Acharya, B.; Dutta, A. Numerical comparison of a combined
hydrothermal carbonization and anaerobic digestion system with direct combustion of biomass for power
production. Processes 2020, 8, 43. [CrossRef]
11. Li, H.-Y.; Tsai, G.-L.; Chao, S.-M.; Yen, Y.-F. Measurement of thermal and hydraulic performance of a plate-fin
heat sink with a shield. Exp. Therm. Fluid Sci. 2012, 42, 71–78. [CrossRef]
21
Energies 2020, 13, 4098
12. Ingrao, C.; Bacenetti, J.; Adamczyk, J.; Ferrante, V.; Messineo, A.; Huisingh, D. Investigating energy and
environmental issues of agro-biogas derived energy systems: A comprehensive review of Life Cycle
Assessments. Renew. Energy 2019, 136, 296–307. [CrossRef]
13. Lucian, M.; Fiori, L. Hydrothermal carbonization of waste biomass: Process design, modeling,
energy efficiency and cost analysis. Energies 2017, 10, 211. [CrossRef]
14. Munir, M.T.; Mansouri, S.S.; Udugama, I.A.; Baroutian, S.; Gernaey, K.V.; Young, B.R. Resource recovery
from organic solid waste using hydrothermal processing: Opportunities and challenges. Renew. Sustain.
Energy Rev. 2018, 96, 64–75. [CrossRef]
15. Álvarez-Murillo, A.; Román, S.; Ledesma, B.; Sabio, E. Study of variables in energy densification of olive
stone by hydrothermal carbonization. J. Anal. Appl. Pyrolysis 2015, 113, 307–314. [CrossRef]
16. Mäkelä, M.; Benavente, V.; Fullana, A. Hydrothermal carbonization of lignocellulosic biomass: Effect of
process conditions on hydrochar properties. Appl. Energy 2015, 155, 576–584. [CrossRef]
17. Volpe, M.; Goldfarb, J.L.; Fiori, L. Hydrothermal carbonization of Opuntia ficus-indica cladodes: Role of
process parameters on hydrochar properties. Bioresour. Technol. 2018, 247, 310–318. [CrossRef]
18. Kruse, A.; Funke, A.; Titirici, M.-M. Hydrothermal conversion of biomass to fuels and energetic materials.
Curr. Opin. Chem. Biol. 2013, 17, 515–521. [CrossRef]
19. Volpe, M.; Wüst, D.; Merzari, F.; Lucian, M.; Andreottola, G.; Kruse, A.; Fiori, L. One stage olive mill waste
streams valorisation via hydrothermal carbonisation. Waste Manag. 2018, 80, 224–234. [CrossRef]
20. Reza, M.; Lynam, J.G.; Helal Uddin, M.; Coronella, C.J. Hydrothermal carbonization: Fate of inorganics.
Biomass Bioenergy 2013, 49, 86–94. [CrossRef]
21. Reza, M.T.; Emerson, R.; Uddin, M.H.; Gresham, G.; Coronella, C.J. Ash reduction of corn stover by mild
hydrothermal preprocessing. Biomass Convers. Biorefinery 2015, 5, 21–31. [CrossRef]
22. Gao, L.; Volpe, M.; Lucian, M.; Fiori, L.; Goldfarb, J.L. Does Hydrothermal Carbonization as a Biomass
Pretreatment Reduce Fuel Segregation of Coal-Biomass Blends During Oxidation? Energy Convers. Manag.
2019, 181, 93–104. [CrossRef]
23. Smith, A.M.; Singh, S.; Ross, A.B. Fate of inorganic material during hydrothermal carbonisation of biomass:
Influence of feedstock on combustion behaviour of hydrochar. Fuel 2016, 169, 135–145. [CrossRef]
24. Liu, Z.; Balasubramanian, R. Upgrading of waste biomass by hydrothermal carbonization (HTC) and low
temperature pyrolysis (LTP): A comparative evaluation. Appl. Energy 2014, 114, 857–864. [CrossRef]
25. Unrean, P.; Lai Fui, B.C.; Rianawati, E.; Acda, M. Comparative techno-economic assessment and environmental
impacts of rice husk-to-fuel conversion technologies. Energy 2018, 151, 581–593. [CrossRef]
26. Kruse, A.; Koch, F.; Stelzl, K.; Wüst, D.; Zeller, M. Fate of Nitrogen during Hydrothermal Carbonization.
Energy Fuels 2016, 30, 8037–8042. [CrossRef]
27. Rodriguez Correa, C.; Otto, T.; Kruse, A. Influence of the biomass components on the pore formation of
activated carbon. Biomass Bioenergy 2017, 97, 53–64. [CrossRef]
28. Liu, F.; Gao, Y.; Zhang, C.; Huang, H.; Yan, C.; Chu, X.; Xu, Z.; Wang, Z.; Zhang, H.; Xiao, X.; et al.
Highly microporous carbon with nitrogen-doping derived from natural biowaste for high-performance
flexible solid-state supercapacitor. J. Colloid Interface Sci. 2019, 548, 322–332. [CrossRef]
29. Yu, X.; Liu, S.; Lin, G.; Yang, Y.; Zhang, S.; Zhao, H.; Zheng, C.; Gao, X. KOH-activated hydrochar with
engineered porosity as sustainable adsorbent for volatile organic compounds. Colloids Surfaces A Physicochem.
Eng. Asp. 2020, 588, 124372. [CrossRef]
30. Hitzl, M.; Corma, A.; Pomares, F.; Renz, M. The hydrothermal carbonization (HTC) plant as a decentral
biorefinery for wet biomass. Catal. Today 2015, 257, 154–159. [CrossRef]
31. Titirici, M.M. Hydrothermal Carbons: Synthesis, Characterization, and Applications. In Novel Carbon
Adsorbents; Elsevier: Amsterdam, The Netherlands, 2012; ISBN 9780080977447.
32. Jatzwauck, M.; Schumpe, A. Kinetics of hydrothermal carbonization (HTC) of soft rush. Biomass Bioenergy
2015, 75, 94–100. [CrossRef]
33. Volpe, M.; Fiori, L. From olive waste to solid biofuel through hydrothermal carbonisation: The role of
temperature and solid load on secondary char formation and hydrochar energy properties. J. Anal. Appl.
Pyrolysis 2017, 124, 63–72. [CrossRef]
34. Knežević, D.; Van Swaaij, W.; Kersten, S. Hydrothermal conversion of biomass. II. conversion of wood,
pyrolysis oil, and glucose in hot compressed water. Ind. Eng. Chem. Res. 2010, 49, 104–112. [CrossRef]
22
Energies 2020, 13, 4098
35. Kruse, A.; Dahmen, N. Water—A magic solvent for biomass conversion. J. Supercrit. Fluids 2015, 96, 36–45.
[CrossRef]
36. Sabio, E.; Álvarez-Murillo, A.; Román, S.; Ledesma, B. Conversion of tomato-peel waste into solid fuel
by hydrothermal carbonization: Influence of the processing variables. Waste Manag. 2016, 47, 122–132.
[CrossRef]
37. Heidari, M.; Norouzi, O.; Salaudeen, S.; Acharya, B.; Dutta, A. Prediction of Hydrothermal Carbonization
with Respect to the Biomass Components and Severity Factor. Energy Fuels 2019, 33, 9916–9924. [CrossRef]
38. Borrero-López, A.M.; Masson, E.; Celzard, A.; Fierro, V. Modelling the reactions of cellulose, hemicellulose
and lignin submitted to hydrothermal treatment. Ind. Crops Prod. 2018, 124, 919–930. [CrossRef]
39. Mäkelä, M.; Volpe, M.; Volpe, R.; Fiori, L.; Dahl, O. Spatially resolved spectral determination of polysaccharides
in hydrothermally carbonized biomass. Green Chem. 2018, 20, 1114–1120. [CrossRef]
40. Volpe, M.; Messineo, A.; Mäkelä, M.; Barr, M.R.; Volpe, R.; Corrado, C.; Fiori, L. Reactivity of cellulose during
hydrothermal carbonization of lignocellulosic biomass. Fuel Process. Technol. 2020, 206, 106456. [CrossRef]
41. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.M.; Fühner, C.; Bens, O.;
Kern, J.; et al. Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry,
processes and applications of wet and dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
42. Sevilla, M.; Fuertes, A.B. The production of carbon materials by hydrothermal carbonization of cellulose.
Carbon N. Y. 2009, 47, 2281–2289. [CrossRef]
43. Olszewski, M.P.; Nicolae, S.A.; Arauzo, P.J.; Titirici, M.M.; Kruse, A. Wet and dry? Influence of hydrothermal
carbonization on the pyrolysis of spent grains. J. Clean. Prod. 2020, 260, 121101. [CrossRef]
44. Kruse, A.; Zevaco, T.A. Properties of hydrochar as function of feedstock, reaction conditions and
post-treatment. Energies 2018, 11, 674. [CrossRef]
45. Islam, M.A.; Asif, M.; Hameed, B.H. Pyrolysis kinetics of raw and hydrothermally carbonized Karanj
(Pongamia pinnata) fruit hulls via thermogravimetric analysis. Bioresour. Technol. 2015, 179, 227–233.
[CrossRef] [PubMed]
46. Kim, D.; Lee, K.; Park, K.Y. Upgrading the characteristics of biochar from cellulose, lignin, and xylan for
solid biofuel production from biomass by hydrothermal carbonization. J. Ind. Eng. Chem. 2016, 42, 95–100.
[CrossRef]
47. Zhuang, X.; Zhan, H.; Song, Y.; Yin, X.; Wu, C. Structure-reactivity relationships of biowaste-derived
hydrochar on subsequent pyrolysis and gasification performance. Energy Convers. Manag. 2019, 199, 112014.
[CrossRef]
48. Heidari, M.; Dutta, A.; Acharya, B.; Mahmud, S. A review of the current knowledge and challenges of
hydrothermal carbonization for biomass conversion. J. Energy Inst. 2019, 92, 1779–1799. [CrossRef]
49. Heidari, M.; Salaudeen, S.; Dutta, A.; Acharya, B. Effects of Process Water Recycling and Particle Sizes on
Hydrothermal Carbonization of Biomass. Energy Fuels 2018. [CrossRef]
50. Li, J.; Zhao, P.; Li, T.; Lei, M.; Yan, W.; Ge, S. Pyrolysis behavior of hydrochar from hydrothermal carbonization
of pinewood sawdust. J. Anal. Appl. Pyrolysis 2020, 146, 104771. [CrossRef]
51. Lucian, M.; Volpe, M.; Fiori, L. Hydrothermal carbonization kinetics of lignocellulosic agro-wastes:
Experimental data and modeling. Energies 2019, 12, 516. [CrossRef]
52. Ulbrich, M.; Preßl, D.; Fendt, S.; Gaderer, M.; Spliethoff, H. Impact of HTC reaction conditions on the
hydrochar properties and CO2 gasification properties of spent grains. Fuel Process. Technol. 2017, 167,
663–669. [CrossRef]
53. Basso, D.; Patuzzi, F.; Castello, D.; Baratieri, M.; Rada, E.C.; Weiss-Hortala, E.; Fiori, L. Agro-industrial waste
to solid biofuel through hydrothermal carbonization. Waste Manag. 2016, 47, 114–121. [CrossRef] [PubMed]
54. Zhuang, X.; Song, Y.; Zhan, H.; Yin, X.; Wu, C. Gasification performance of biowaste-derived hydrochar:
The properties of products and the conversion processes. Fuel 2020, 260, 116320. [CrossRef]
55. Chen, X.; Lin, Q.; He, R.; Zhao, X.; Li, G. Hydrochar production from watermelon peel by hydrothermal
carbonization. Bioresour. Technol. 2017, 241, 236–243. [CrossRef] [PubMed]
56. Magdziarz, A.; Wilk, M.; Wadrzyk, ˛ M. Pyrolysis of hydrochar derived from biomass—Experimental
investigation. Fuel 2020, 167, 117246. [CrossRef]
57. Roy, P.; Dutta, A.; Gallant, J. Hydrothermal carbonization of peat moss and herbaceous biomass (Miscanthus):
A potential route for bioenergy. Energies 2018, 11, 2794. [CrossRef]
23
Energies 2020, 13, 4098
58. Khatami, R.; Stivers, C.; Joshi, K.; Levendis, Y.A.; Sarofim, A.F. Combustion behavior of single particles from
three different coal ranks and from sugar cane bagasse in O2 /N2 and O2 /CO2 atmospheres. Combust. Flame
2012, 159, 1253–1271. [CrossRef]
59. Wilk, M.; Magdziarz, A.; Kalemba-Rec, I.; Szymańska-Chargot, M. Upgrading of green waste into carbon-rich
solid biofuel by hydrothermal carbonization: The effect of process parameters on hydrochar derived from
acacia. Energy 2020, 202, 117717. [CrossRef]
60. Babinszki, B.; Jakab, E.; Sebestyén, Z.; Blazsó, M.; Berényi, B.; Kumar, J.; Krishna, B.B.; Bhaskar, T.; Czégény, Z.
Comparison of hydrothermal carbonization and torrefaction of azolla biomass: Analysis of the solid products.
J. Anal. Appl. Pyrolysis 2020, 104844. [CrossRef]
61. Sharma, H.B.; Sarmah, A.K.; Dubey, B. Hydrothermal carbonization of renewable waste biomass for solid
biofuel production: A discussion on process mechanism, the influence of process parameters, environmental
performance and fuel properties of hydrochar. Renew. Sustain. Energy Rev. 2020, 123, 109761. [CrossRef]
62. Zhang, X.; Gao, B.; Zhao, S.; Wu, P.; Han, L.; Liu, X. Optimization of a “coal-like” pelletization technique
based on the sustainable biomass fuel of hydrothermal carbonization of wheat straw. J. Clean. Prod. 2020,
242, 118426. [CrossRef]
63. Peng, C.; Zhai, Y.; Zhu, Y.; Xu, B.; Wang, T.; Li, C.; Zeng, G. Production of char from sewage sludge employing
hydrothermal carbonization: Char properties, combustion behavior and thermal characteristics. Fuel 2016,
176, 110–118. [CrossRef]
64. Gunarathne, D.S.; Mueller, A.; Fleck, S.; Kolb, T.; Chmielewski, J.K.; Yang, W.; Blasiak, W. Gasification
characteristics of hydrothermal carbonized biomass in an updraft pilot-scale gasifier. Energy Fuels 2014, 28,
1992–2002. [CrossRef]
65. Huang, G.; Wang, Y.; Zhang, T.; Wu, X.; Cai, J. High-performance hierarchical N-doped porous carbons from
hydrothermally carbonized bamboo shoot shells for symmetric supercapacitors. J. Taiwan Inst. Chem. Eng.
2019, 96, 672–680. [CrossRef]
66. Saari, J.; Sermyagina, E.; Kaikko, J.; Vakkilainen, E.; Sergeev, V. Integration of hydrothermal carbonization
and a CHP plant: Part 2—Operational and economic analysis. Energy 2016, 113, 574–585. [CrossRef]
67. He, C.; Wang, K.; Yang, Y.; Wang, J.Y. Utilization of sewage-sludge-derived hydrochars toward efficient
cocombustion with different-rank coals: Effects of subcritical water conversion and blending scenarios.
Energy Fuels 2014, 28, 6140–6150. [CrossRef]
68. Reza, M.T.; Andert, J.; Wirth, B.; Busch, D.; Pielert, J.; Lynam, J.G.; Mumme, J. Hydrothermal Carbonization
of Biomass for Energy and Crop Production. Appl. Bioenergy 2014. [CrossRef]
69. Saha, N.; Saba, A.; Saha, P.; McGaughy, K.; Franqui-Villanueva, D.; Orts, W.J.; Hart-Cooper, W.M.; Toufiq
Reza, M. Hydrothermal carbonization of various paper mill sludges: An observation of solid fuel properties.
Energies 2019, 125, 858. [CrossRef]
70. Volpe, M.; Fiori, L.; Merzari, F.; Messineo, A.; Andreottola, G. Hydrothermal carbonization as an efficient tool
for sewage sludge valorization and phosphorous recovery. Chem. Eng. Trans. 2020, 80, 199–204. [CrossRef]
71. Merzari, F.; Goldfarb, J.; Andreottola, G.; Mimmo, T.; Volpe, M.; Fiori, L. Hydrothermal carbonization as a
strategy for sewage sludge management: Influence of process withdrawal point on hydrochar properties.
Energies 2020, 13, 2890. [CrossRef]
72. Wang, R.; Wang, C.; Zhao, Z.; Jia, J.; Jin, Q. Energy recovery from high-ash municipal sewage sludge by
hydrothermal carbonization: Fuel characteristics of biosolid products. Energy 2019, 186, 115848. [CrossRef]
73. Chen, C.; Liu, G.; An, Q.; Lin, L.; Shang, Y.; Wan, C. From wasted sludge to valuable biochar by low
temperature hydrothermal carbonization treatment: Insight into the surface characteristics. J. Clean. Prod.
2020, 263, 121600. [CrossRef]
74. Park, K.Y.; Lee, K.; Kim, D. Characterized hydrochar of algal biomass for producing solid fuel through
hydrothermal carbonization. Bioresour. Technol. 2018, 258, 119–124. [CrossRef] [PubMed]
75. Saba, A.; McGaughy, K.; Toufiq Reza, M. Techno-economic assessment of co-hydrothermal carbonization of a
coal-Miscanthus blend. Energies 2019, 12, 630. [CrossRef]
76. Zheng, C.; Ma, X.; Yao, Z.; Chen, X. The properties and combustion behaviors of hydrochars derived from
co-hydrothermal carbonization of sewage sludge and food waste. Bioresour. Technol. 2019, 258, 121347.
[CrossRef] [PubMed]
24
Energies 2020, 13, 4098
77. Wang, T.; Zhai, Y.; Zhu, Y.; Gan, X.; Zheng, L.; Peng, C.; Wang, B.; Li, C.; Zeng, G. Evaluation of the
clean characteristics and combustion behavior of hydrochar derived from food waste towards solid biofuel
production. Bioresour. Technol. 2018, 266, 275–283. [CrossRef]
78. McGaughy, K.; Toufiq Reza, M. Hydrothermal carbonization of food waste: Simplified process simulation
model based on experimental results. Biomass Convers. Biorefinery 2018, 8, 283–292. [CrossRef]
79. Cao, Z.; Jung, D.; Olszewski, M.P.; Arauzo, P.J.; Kruse, A. Hydrothermal carbonization of biogas digestate:
Effect of digestate origin and process conditions. Waste Manag. 2019, 100, 138–150. [CrossRef]
80. Pawlak-Kruczek, H.; Sieradzka, M.; Mlonka-M˛edrala, A.; Baranowski, M.; Serafin-Tkaczuk, M.; Magdziarz, A.;
Niedźwiecki, Ł. Structural and energetic properties of hydrochars obtained from agricultural and municipal
solid waste digestates. In Proceedings of the 32nd International Conference on Efficiency, Cost, Optimization,
Simulation and Environmental Impact of Energy Systems (ECOS 2019), Wroclaw, Poland, 23–28 June 2019.
81. Başakçılardan Kabakcı, S.; Baran, S.S. Hydrothermal carbonization of various lignocellulosics: Fuel
characteristics of hydrochars and surface characteristics of activated hydrochars. Waste Manag. 2019,
100, 259–268. [CrossRef]
82. Yang, W.; Wang, H.; Zhang, M.; Zhu, J.; Zhou, J.; Wu, S. Fuel properties and combustion kinetics of hydrochar
prepared by hydrothermal carbonization of bamboo. Bioresour. Technol. 2016, 205, 199–204. [CrossRef]
83. Chen, X.; Ma, X.; Peng, X.; Lin, Y.; Yao, Z. Conversion of sweet potato waste to solid fuel via hydrothermal
carbonization. Bioresour. Technol. 2018, 249, 900–907. [CrossRef]
84. Celiktas, M.S.; Alptekin, F.M. Conversion of model biomass to carbon-based material with high conductivity
by using carbonization. Energy 2019, 188, 116089. [CrossRef]
85. Sinan, N.; Unur, E. Hydrothermal conversion of lignocellulosic biomass into high-value energy storage
materials. J. Energy Chem. 2017, 26, 783–789. [CrossRef]
86. Wang, C.; Wu, D.; Wang, H.; Gao, Z.; Xu, F.; Jiang, K. Biomass derived nitrogen-doped hierarchical porous
carbon sheets for supercapacitors with high performance. J. Colloid Interface Sci. 2018, 523, 133–143. [CrossRef]
[PubMed]
87. Antero, R.V.P.; Alves, A.C.F.; de Oliveira, S.B.; Ojala, S.A.; Brum, S.S. Challenges and alternatives for the
adequacy of hydrothermal carbonization of lignocellulosic biomass in cleaner production systems: A review.
J. Clean. Prod. 2020, 252, 119899. [CrossRef]
88. Guo, N.; Luo, W.; Guo, R.; Qiu, D.; Zhao, Z.; Wang, L.; Jia, D.; Guo, J. Interconnected and hierarchical porous
carbon derived from soybean root for ultrahigh rate supercapacitors. J. Alloys Compd. 2020, 834, 155115.
[CrossRef]
89. Maniscalco, M.P.; Corrado, C.; Volpe, R.; Messineo, A. Evaluation of the optimal activation parameters for
almond shell bio-char production for capacitive deionization. Bioresour. Technol. Reports 2020, 11, 100435.
[CrossRef]
90. Sevilla, M.; Ferrero, G.A.; Fuertes, A.B. Beyond KOH activation for the synthesis of superactivated carbons
from hydrochar. Carbon N. Y. 2017, 114, 50–58. [CrossRef]
91. Liu, Y.; An, Z.; Wu, M.; Yuan, A.; Zhao, H.; Zhang, J.; Xu, J. Peony pollen derived nitrogen-doped activated
carbon for supercapacitor application. Chinese Chem. Lett. 2019, 31, 1644–1647. [CrossRef]
92. Manyala, N.; Bello, A.; Barzegar, F.; Khaleed, A.A.; Momodu, D.Y.; Dangbegnon, J.K. Coniferous pine
biomass: A novel insight into sustainable carbon materials for supercapacitors electrode. Mater. Chem. Phys.
2016, 182, 139–147. [CrossRef]
93. Zhao, Y.Q.; Lu, M.; Tao, P.Y.; Zhang, Y.J.; Gong, X.T.; Yang, Z.; Zhang, G.Q.; Li, H.L. Hierarchically porous
and heteroatom doped carbon derived from tobacco rods for supercapacitors. J. Power Sources 2016, 307,
391–400. [CrossRef]
94. Sun, W.; Lipka, S.M.; Swartz, C.; Williams, D.; Yang, F. Hemp-derived activated carbons for supercapacitors.
Carbon N. Y. 2016, 103, 181–192. [CrossRef]
95. Dai, C.; Wan, J.; Yang, J.; Qu, S.; Jin, T.; Ma, F.; Shao, J. H3 PO4 solution hydrothermal carbonization combined
with KOH activation to prepare argy wormwood-based porous carbon for high-performance supercapacitors.
Appl. Surf. Sci. 2018, 444, 105–117. [CrossRef]
96. Ye, R.; Cai, J.; Pan, Y.; Qiao, X.; Sun, W. Microporous carbon from malva nut for supercapacitors: Effects of
primary carbonizations on structures and performances. Diam. Relat. Mater. 2020, 105, 107816. [CrossRef]
25
Energies 2020, 13, 4098
97. Jain, A.; Xu, C.; Jayaraman, S.; Balasubramanian, R.; Lee, J.Y.; Srinivasan, M.P. Mesoporous activated carbons
with enhanced porosity by optimal hydrothermal pre-treatment of biomass for supercapacitor applications.
Microporous Mesoporous Mater. 2015, 218, 55–61. [CrossRef]
98. Boyjoo, Y.; Cheng, Y.; Zhong, H.; Tian, H.; Pan, J.; Pareek, V.K.; Jiang, S.P.; Lamonier, J.F.; Jaroniec, M.;
Liu, J. From waste Coca Cola® to activated carbons with impressive capabilities for CO2 adsorption and
supercapacitors. Carbon N. Y. 2017, 116, 490–499. [CrossRef]
99. Du, P.; Hu, X.; Yi, C.; Liu, H.C.; Liu, P.; Zhang, H.L.; Gong, X. Self-powered electronics by integration of
flexible solid-state graphene-based supercapacitors with high performance perovskite hybrid solar cells.
Adv. Funct. Mater. 2015, 25, 2420–2427. [CrossRef]
100. Gupta, V.; Miura, N. Electrochemically deposited polyaniline nanowire’s network a high-performance
electrode material for redox supercapacitor. Electrochem. Solid-State Lett. 2005, 8, A630–A632. [CrossRef]
101. Zhao, G.; Cheng, Y.; Sun, P.; Ma, W.; Hao, S.; Wang, X.; Xu, X.; Xu, Q.; Liu, M. Biocarbon based template
synthesis of uniform lamellar MoS2 nanoflowers with excellent energy storage performance in lithium-ion
battery and supercapacitors. Electrochim. Acta 2020, 331, 135262. [CrossRef]
102. Yu, K.; Wang, J.; Song, K.; Wang, X.; Liang, C.; Dou, Y. Hydrothermal synthesis of cellulose-derived carbon
nanospheres from corn straw as anode materials for lithium ion batteries. Nanomaterials 2019, 9, 93. [CrossRef]
103. Ren, X.; Xu, S.D.; Liu, S.; Chen, L.; Zhang, D.; Qiu, L. Lath-shaped biomass derived hard carbon as anode
materials with super rate capability for sodium-ion batteries. J. Electroanal. Chem. 2019, 841, 63–72. [CrossRef]
104. Maharjan, M.; Wai, N.; Veksha, A.; Giannis, A.; Lim, T.M.; Lisak, G. Sal wood sawdust derived highly
mesoporous carbon as prospective electrode material for vanadium redox flow batteries. J. Electroanal. Chem.
2019, 834, 94–100. [CrossRef]
105. Palomares, V.; Blas, M.; Serras, P.; Iturrondobeitia, A.; Peña, A.; Lopez-Urionabarrenechea, A.; Lezama, L.;
Rojo, T. Waste Biomass as in Situ Carbon Source for Sodium Vanadium Fluorophosphate/C Cathodes for
Na-Ion Batteries. ACS Sustain. Chem. Eng. 2018, 6, 16386–16398. [CrossRef]
106. Rodríguez Correa, C.; Ngamying, C.; Klank, D.; Kruse, A. Investigation of the textural and adsorption
properties of activated carbon from HTC and pyrolysis carbonizates. Biomass Convers. Biorefinery 2018, 8,
317–328. [CrossRef]
107. Rodriguez Correa, C.; Bernardo, M.; Ribeiro, R.P.P.L.; Esteves, I.A.A.C.; Kruse, A. Evaluation of hydrothermal
carbonization as a preliminary step for the production of functional materials from biogas digestate. J. Anal.
Appl. Pyrolysis 2017, 124, 461–474. [CrossRef]
108. Bernardo, M.; Correa, C.R.; Ringelspacher, Y.; Becker, G.C.; Lapa, N.; Fonseca, I.; Esteves, I.A.A.C.; Kruse, A.
Porous carbons derived from hydrothermally treated biogas digestate. Waste Manag. 2020, 105, 170–179.
[CrossRef] [PubMed]
109. Vikrant, K.; Kim, K.H.; Ok, Y.S.; Tsang, D.C.W.; Tsang, Y.F.; Giri, B.S.; Singh, R.S. Engineered/designer biochar
for the removal of phosphate in water and wastewater. Sci. Total Environ. 2018, 616, 1242–1260. [CrossRef]
110. Takaya, C.A.; Fletcher, L.A.; Singh, S.; Anyikude, K.U.; Ross, A.B. Phosphate and ammonium sorption
capacity of biochar and hydrochar from different wastes. Chemosphere 2016, 145, 518–527. [CrossRef]
111. Shepherd, J.G.; Joseph, S.; Sohi, S.P.; Heal, K.V. Biochar and enhanced phosphate capture: Mapping
mechanisms to functional properties. Chemosphere 2017, 179, 57–74. [CrossRef]
112. Huang, G.G.; Liu, Y.F.; Wu, X.X.; Cai, J.J. Activated carbons prepared by the KOH activation of a hydrochar
from garlic peel and their CO2 adsorption performance. Xinxing Tan Cailiao/New Carbon Mater. 2019, 34,
247–257. [CrossRef]
113. Chen, Y.; Chen, J.; Chen, S.; Tian, K.; Jiang, H. Ultra-high capacity and selective immobilization of Pb through
crystal growth of hydroxypyromorphite on amino-functionalized hydrochar. J. Mater. Chem. A 2015, 3,
9843–9850. [CrossRef]
114. Han, L.; Sun, H.; Ro, K.S.; Sun, K.; Libra, J.A.; Xing, B. Removal of antimony (III) and cadmium (II) from
aqueous solution using animal manure-derived hydrochars and pyrochars. Bioresour. Technol. 2017, 234,
77–85. [CrossRef] [PubMed]
115. Sun, K.; Tang, J.; Gong, Y.; Zhang, H. Characterization of potassium hydroxide (KOH) modified hydrochars
from different feedstocks for enhanced removal of heavy metals from water. Environ. Sci. Pollut. Res. 2015,
22, 16640–16651. [CrossRef] [PubMed]
26
Energies 2020, 13, 4098
116. Zhou, N.; Chen, H.; Xi, J.; Yao, D.; Zhou, Z.; Tian, Y.; Lu, X. Biochars with excellent Pb(II) adsorption property
produced from fresh and dehydrated banana peels via hydrothermal carbonization. Bioresour. Technol. 2017,
232, 204–210. [CrossRef] [PubMed]
117. Petrović, J.T.; Stojanović, M.D.; Milojković, J.V.; Petrović, M.S.; Šoštarić, T.D.; Laušević, M.D.; Mihajlović, M.L.
Alkali modified hydrochar of grape pomace as a perspective adsorbent of Pb2+ from aqueous solution.
J. Environ. Manage. 2016, 182, 292–300. [CrossRef] [PubMed]
118. Semerciöz, A.S.; Göğüş, F.; Çelekli, A.; Bozkurt, H. Development of carbonaceous material from grapefruit
peel with microwave implemented-low temperature hydrothermal carbonization technique for the adsorption
of Cu (II). J. Clean. Prod. 2017, 165, 599–610. [CrossRef]
119. Deng, J.; Li, X.; Wei, X.; Liu, Y.; Liang, J.; Tang, N.; Song, B.; Chen, X.; Cheng, X. Sulfamic acid modified
hydrochar derived from sawdust for removal of benzotriazole and Cu(II) from aqueous solution: Adsorption
behavior and mechanism. Bioresour. Technol. 2019, 290, 121765. [CrossRef]
120. Zuo, X.J.; Liu, Z.; Chen, M.D. Effect of H2 O2 concentrations on copper removal using the modified
hydrothermal biochar. Bioresour. Technol. 2016, 207, 262–267. [CrossRef]
121. Fernandez, M.E.; Ledesma, B.; Román, S.; Bonelli, P.R.; Cukierman, A.L. Development and characterization
of activated hydrochars from orange peels as potential adsorbents for emerging organic contaminants.
Bioresour. Technol. 2015, 183, 221–228. [CrossRef]
122. Lima, H.H.C.; Maniezzo, R.S.; Kupfer, V.L.; Guilherme, M.R.; Moises, M.P.; Arroyo, P.A.; Rinaldi, A.W.
Hydrochars based on cigarette butts as a recycled material for the adsorption of pollutants. J. Environ. Chem.
Eng. 2018, 6, 7054–7061. [CrossRef]
123. Khataee, A.; Kayan, B.; Kalderis, D.; Karimi, A.; Akay, S.; Konsolakis, M. Ultrasound-assisted removal of
Acid Red 17 using nanosized Fe3 O4 -loaded coffee waste hydrochar. Ultrason. Sonochem. 2017, 35, 72–80.
[CrossRef]
124. Khoshbouy, R.; Takahashi, F.; Yoshikawa, K. Preparation of high surface area sludge-based activated
hydrochar via hydrothermal carbonization and application in the removal of basic dye. Environ. Res. 2019,
175, 457–467. [CrossRef] [PubMed]
125. Tang, Z.; Deng, Y.; Luo, T.; Xu, Y.; Zhu, N. Enhanced removal of Pb(II) by supported nanoscale Ni/Fe on
hydrochar derived from biogas residues. Chem. Eng. J. 2016, 292, 224–232. [CrossRef]
126. Buapeth, P.; Watcharin, W.; Dechtrirat, D.; Chuenchom, L. Carbon Adsorbents from Sugarcane Bagasse
Prepared through Hydrothermal Carbonization for Adsorption of Methylene Blue: Effect of Heat Treatment
on Adsorption Efficiency. IOP Conf. Ser. Mater. Sci. Eng. 2019, 515, 12003. [CrossRef]
127. Qian, W.C.; Luo, X.P.; Wang, X.; Guo, M.; Li, B. Removal of methylene blue from aqueous solution by
modified bamboo hydrochar. Ecotoxicol. Environ. Saf. 2018, 157, 300–306. [CrossRef]
128. Islam, M.A.; Benhouria, A.; Asif, M.; Hameed, B.H. Methylene blue adsorption on factory-rejected tea
activated carbon prepared by conjunction of hydrothermal carbonization and sodium hydroxide activation
processes. J. Taiwan Inst. Chem. Eng. 2015, 52, 57–64. [CrossRef]
129. Fang, J.; Gao, B.; Chen, J.; Zimmerman, A.R. Hydrochars derived from plant biomass under various
conditions: Characterization and potential applications and impacts. Chem. Eng. J. 2015, 267, 253–259.
[CrossRef]
130. Wu, J.; Yang, J.; Feng, P.; Huang, G.; Xu, C.; Lin, B. High-efficiency removal of dyes from wastewater by fully
recycling litchi peel biochar. Chemosphere 2020, 246, 125734. [CrossRef]
131. Zheng, H.; Sun, Q.; Li, Y.; Du, Q. Biosorbents prepared from pomelo peel by hydrothermal technique and its
adsorption properties for congo red. Mater. Res. Express 2020, 7, 45505. [CrossRef]
132. Li, B.; Wang, Q.; Guo, J.Z.; Huan, W.W.; Liu, L. Sorption of methyl orange from aqueous solution by
protonated amine modified hydrochar. Bioresour. Technol. 2018, 268, 454–459. [CrossRef]
133. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X. Production and optimization of bamboo hydrochars for adsorption
of Congo red and 2-naphthol. Bioresour. Technol. 2016, 207, 379–386. [CrossRef]
134. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X.; Wang, H.; Tsend, N. Hydrochars from bamboo sawdust through
acid assisted and two-stage hydrothermal carbonization for removal of two organics from aqueous solution.
Bioresour. Technol. 2018, 261, 257–264. [CrossRef] [PubMed]
135. Wei, J.; Liu, Y.; Li, J.; Yu, H.; Peng, Y. Removal of organic contaminant by municipal sewage sludge-derived
hydrochar: Kinetics, thermodynamics and mechanisms. Water Sci. Technol. 2018, 78, 947–956. [CrossRef]
[PubMed]
27
Energies 2020, 13, 4098
136. Zhang, H.; Zhang, F.; Huang, Q. Highly effective removal of malachite green from aqueous solution by
hydrochar derived from phycocyanin-extracted algal bloom residues through hydrothermal carbonization.
RSC Adv. 2017, 7, 5790–5799. [CrossRef]
137. Hammud, H.H.; Shmait, A.; Hourani, N. Removal of Malachite Green from water using hydrothermally
carbonized pine needles. RSC Adv. 2015, 5, 7909–7920. [CrossRef]
138. Feng, Y.; Sun, H.; Han, L.; Xue, L.; Chen, Y.; Yang, L.; Xing, B. Fabrication of hydrochar based on food waste
(FWHTC) and its application in aqueous solution rare earth ions adsorptive removal: Process, mechanisms
and disposal methodology. J. Clean. Prod. 2019, 212, 1423–1433. [CrossRef]
139. Xiao, K.; Liu, H.; Li, Y.; Yang, G.; Wang, Y.; Yao, H. Excellent performance of porous carbon from urea-assisted
hydrochar of orange peel for toluene and iodine adsorption. Chem. Eng. J. 2020, 382, 122997. [CrossRef]
140. Rattanachueskul, N.; Saning, A.; Kaowphong, S.; Chumha, N.; Chuenchom, L. Magnetic carbon composites
with a hierarchical structure for adsorption of tetracycline, prepared from sugarcane bagasse via hydrothermal
carbonization coupled with simple heat treatment process. Bioresour. Technol. 2017, 226, 164–172. [CrossRef]
141. Zhu, X.; Qian, F.; Liu, Y.; Matera, D.; Wu, G.; Zhang, S.; Chen, J. Controllable synthesis of magnetic carbon
composites with high porosity and strong acid resistance from hydrochar for efficient removal of organic
pollutants: An overlooked influence. Carbon N. Y. 2016, 99, 338–347. [CrossRef]
142. Yu, J.; Zhu, Z.; Zhang, H.; Chen, T.; Qiu, Y.; Xu, Z.; Yin, D. Efficient removal of several estrogens in water
by Fe-hydrochar composite and related interactive effect mechanism of H2 O2 and iron with persistent free
radicals from hydrochar of pinewood. Sci. Total Environ. 2019, 658, 1013–1022. [CrossRef]
143. European Committee. EU Report on Critical Raw Materials List; EC: Brussels, Belgium, 2017.
144. Ava Green Chemistry Development GMBH. Sewage sludge reuse Phosphate recovery with an innovative HTC
technology (HTCycle); Ava Green Chemistry Development GMBH: Munich, Germany, 2015.
145. Marin-Batista, J.D.; Mohedano, A.F.; Rodriguez, J.J.; de la Rubia, M.A. Energy and Phosphorous Recovery
Through Hydrothermal Carbonization of Digested Sewage Sludge. Waste Manag. 2020, 105, 566–574.
[CrossRef]
146. Becker, G.C.; Wüst, D.; Köhler, H.; Lautenbach, A.; Kruse, A. Novel approach of phosphate-reclamation as
struvite from sewage sludge by utilising hydrothermal carbonization. J. Environ. Manage. 2019, 238, 119–125.
[CrossRef] [PubMed]
147. Bhatt, D.; Shrestha, A.; Dahal, R.K.; Acharya, B.; Basu, P.; MacEwen, R. Hydrothermal carbonization of
biosolids from Waste water treatment plant. Energies 2018, 11, 2286. [CrossRef]
148. Shi, Y.; Luo, G.; Rao, Y.; Chen, H.; Zhang, S. Hydrothermal conversion of dewatered sewage sludge: Focusing
on the transformation mechanism and recovery of phosphorus. Chemosphere 2019, 228, 619–628. [CrossRef]
[PubMed]
149. Bento, L.R.; Castro, A.J.R.; Moreira, A.B.; Ferreira, O.P.; Bisinoti, M.C.; Melo, C.A. Release of nutrients
and organic carbon in different soil types from hydrochar obtained using sugarcane bagasse and vinasse.
Geoderma 2019, 334, 24–32. [CrossRef]
150. Ji, M.; Sang, W.; Tsang, D.C.W.; Usman, M.; Zhang, S.; Luo, G. Molecular and microbial insights towards
understanding the effects of hydrochar on methane emission from paddy soil. Sci. Total Environ. 2020, 714,
136769. [CrossRef]
151. Eibisch, N.; Durner, W.; Bechtold, M.; Fuß, R.; Mikutta, R.; Woche, S.K.; Helfrich, M. Does water repellency of
pyrochars and hydrochars counter their positive effects on soil hydraulic properties? Geoderma 2015, 245,
31–39. [CrossRef]
152. Zhang, S.; Zhu, X.; Zhou, S.; Shang, H.; Luo, J.; Tsang, D.C.W. Hydrothermal carbonization for hydrochar
production and its application. In Biochar from Biomass and Waste: Fundamentals and Applications; Elsevier:
Amsterdam, The Netherlands, 2018; ISBN 9780128117293.
153. Abel, S.; Peters, A.; Trinks, S.; Schonsky, H.; Facklam, M.; Wessolek, G. Impact of biochar and hydrochar
addition on water retention and water repellency of sandy soil. Geoderma 2013, 202, 183–191. [CrossRef]
154. Scheifele, M.; Hobi, A.; Buegger, F.; Gattinger, A.; Schulin, R.; Boller, T.; Mäder, P. Impact of pyrochar
and hydrochar on soybean (Glycine max L.) root nodulation and biological nitrogen fixation. Zeitschrift fur
Pflanzenernahrung und Bodenkunde 2017, 180, 199–211. [CrossRef]
155. George, E.; Ventura, M.; Panzacchi, P.; Scandellari, F.; Tonon, G. Can hydrochar and pyrochar affect nitrogen
uptake and biomass allocation in poplars? Zeitschrift fur Pflanzenernahrung und Bodenkunde 2017, 180, 178–186.
[CrossRef]
28
Energies 2020, 13, 4098
156. Eibisch, N.; Schroll, R.; Fuß, R. Effect of pyrochar and hydrochar amendments on the mineralization of the
herbicide isoproturon in an agricultural soil. Chemosphere 2015, 134, 528–535. [CrossRef]
157. Eibisch, N.; Schroll, R.; Fuß, R.; Mikutta, R.; Helfrich, M.; Flessa, H. Pyrochars and hydrochars differently alter
the sorption of the herbicide isoproturon in an agricultural soil. Chemosphere 2015, 119, 155–162. [CrossRef]
[PubMed]
158. Bera, T.; Purakayastha, T.J.; Patra, A.K.; Datta, S.C. Comparative analysis of physicochemical, nutrient,
and spectral properties of agricultural residue biochars as influenced by pyrolysis temperatures. J. Mater.
Cycles Waste Manag. 2018. [CrossRef]
159. Gronwald, M.; Vos, C.; Helfrich, M.; Don, A. Stability of pyrochar and hydrochar in agricultural soil—A new
field incubation method. Geoderma 2016, 284, 85–92. [CrossRef]
160. Malghani, S.; Jüschke, E.; Baumert, J.; Thuille, A.; Antonietti, M.; Trumbore, S.; Gleixner, G. Carbon
sequestration potential of hydrothermal carbonization char (hydrochar) in two contrasting soils; results of a
1-year field study. Biol. Fertil. Soils 2015, 51, 123–134. [CrossRef]
161. De Jager, M.; Röhrdanz, M.; Giani, L. The influence of hydrochar from biogas digestate on soil improvement
and plant growth aspects. Biochar 2020, 2, 177–194. [CrossRef]
162. Yuan, H.; Lu, T.; Wang, Y.; Chen, Y.; Lei, T. Sewage sludge biochar: Nutrient composition and its effect on the
leaching of soil nutrients. Geoderma 2016, 267, 17–23. [CrossRef]
163. Riedel, T.; Hennessy, P.; Iden, S.C.; Koschinsky, A. Leaching of soil-derived major and trace elements in an
arable topsoil after the addition of biochar. Eur. J. Soil Sci. 2015, 66, 823–834. [CrossRef]
164. Hitzl, M.; Mendez, A.; Owsianiak, M.; Renz, M. Making hydrochar suitable for agricultural soil: A thermal
treatment to remove organic phytotoxic compounds. J. Environ. Chem. Eng. 2018, 6, 7029–7034. [CrossRef]
165. Busch, D.; Stark, A.; Kammann, C.I.; Glaser, B. Genotoxic and phytotoxic risk assessment of fresh and treated
hydrochar from hydrothermal carbonization compared to biochar from pyrolysis. Ecotoxicol. Environ. Saf.
2013, 97, 59–66. [CrossRef]
166. Yue, Y.; Yao, Y.; Lin, Q.; Li, G.; Zhao, X. The change of heavy metals fractions during hydrochar decomposition
in soils amended with different municipal sewage sludge hydrochars. J. Soils Sediments 2017, 17, 763–770.
[CrossRef]
167. Chu, Q.; Xue, L.; Cheng, Y.; Liu, Y.; Feng, Y.; Yu, S.; Meng, L.; Pan, G.; Hou, P.; Duan, J.; et al. Microalgae-derived
hydrochar application on rice paddy soil: Higher rice yield but increased gaseous nitrogen loss. Sci. Total
Environ. 2020, 717, 137127. [CrossRef] [PubMed]
168. Chu, Q.; Xue, L.; Singh, B.P.; Yu, S.; Müller, K.; Wang, H.; Feng, Y.; Pan, G.; Zheng, X.; Yang, L. Sewage
sludge-derived hydrochar that inhibits ammonia volatilization, improves soil nitrogen retention and rice
nitrogen utilization. Chemosphere 2020, 245, 125558. [CrossRef] [PubMed]
169. Subedi, R.; Kammann, C.; Pelissetti, S.; Taupe, N.; Bertora, C.; Monaco, S.; Grignani, C. Does soil amended
with biochar and hydrochar reduce ammonia emissions following the application of pig slurry? Eur. J. Soil
Sci. 2015, 66, 1044–1053. [CrossRef]
170. Andert, J.; Mumme, J. Impact of pyrolysis and hydrothermal biochar on gas-emitting activity of soil
microorganisms and bacterial and archaeal community composition. Appl. Soil Ecol. 2015, 96, 225–239.
[CrossRef]
171. Xia, Y.; Liu, H.; Guo, Y.; Liu, Z.; Jiao, W. Immobilization of heavy metals in contaminated soils by modified
hydrochar: Efficiency, risk assessment and potential mechanisms. Sci. Total Environ. 2019, 685, 1201–1208.
[CrossRef] [PubMed]
172. Xia, Y.; Luo, H.; Li, D.; Chen, Z.; Yang, S.; Liu, Z.; Yang, T.; Gai, C. Efficient immobilization of toxic heavy
metals in multi-contaminated agricultural soils by amino-functionalized hydrochar: Performance, plant
responses and immobilization mechanisms. Environ. Pollut. 2020, 261, 114217. [CrossRef]
173. Du, F.L.; Du, Q.S.; Dai, J.; Tang, P.D.; Li, Y.M.; Long, S.Y.; Xie, N.Z.; Wang, Q.Y.; Huang, R.B. A comparative
study for the organic byproducts from hydrothermal carbonizations of sugarcane bagasse and its bio-refined
components cellulose and lignin. PLoS ONE 2018, 13, e0197188. [CrossRef]
174. Chang, M.Y.; Huang, W.J. Hydrothermal biorefinery of spent agricultural biomass into value-added
bio-nutrient solution: Comparison between greenhouse and field cropping data. Ind. Crops Prod. 2018, 126,
186–189. [CrossRef]
29
Energies 2020, 13, 4098
175. Ruiz, H.A.; Conrad, M.; Sun, S.N.; Sanchez, A.; Rocha, G.J.M.; Romaní, A.; Castro, E.; Torres, A.;
Rodríguez-Jasso, R.M.; Andrade, L.P.; et al. Engineering aspects of hydrothermal pretreatment: From batch
to continuous operation, scale-up and pilot reactor under biorefinery concept. Bioresour. Technol. 2020, 299,
122685. [CrossRef]
176. Ashraf, M.T.; Schmidt, J.E. Process simulation and economic assessment of hydrothermal pretreatment and
enzymatic hydrolysis of multi-feedstock lignocellulose—Separate vs combined processing. Bioresour. Technol.
2018, 249, 835–843. [CrossRef]
177. Aguilar, D.L.; Rodríguez-Jasso, R.M.; Zanuso, E.; de Rodríguez, D.J.; Amaya-Delgado, L.; Sanchez, A.;
Ruiz, H.A. Scale-up and evaluation of hydrothermal pretreatment in isothermal and non-isothermal regimen
for bioethanol production using agave bagasse. Bioresour. Technol. 2018, 263, 112–119. [CrossRef] [PubMed]
178. Nascimento, V.M.; Rossell, C.E.V.; de Moraes Rocha, G.J. Scale-up hydrothermal pretreatment of sugarcane
bagasse and straw for second-generation ethanol production. In Hydrothermal Processing in Biorefineries:
Production of Bioethanol and High Added-Value Compounds of Second and Third Generation Biomass; Springer:
New York, NY, USA, 2017; ISBN 9783319564579.
179. Zakaria, M.R.; Hirata, S.; Hassan, M.A. Hydrothermal pretreatment enhanced enzymatic hydrolysis and
glucose production from oil palm biomass. Bioresour. Technol. 2015, 176, 142–148. [CrossRef] [PubMed]
180. Rivas, S.; Vila, C.; Alonso, J.L.; Santos, V.; Parajó, J.C.; Leahy, J.J. Biorefinery processes for the valorization
of Miscanthus polysaccharides: From constituent sugars to platform chemicals. Ind. Crops Prod. 2019, 134,
309–317. [CrossRef]
181. Deng, A.; Ren, J.; Li, H.; Peng, F.; Sun, R. Corncob lignocellulose for the production of furfural by hydrothermal
pretreatment and heterogeneous catalytic process. RSC Adv. 2015, 5, 60264–60272. [CrossRef]
182. Li, H.; Wang, X.; Liu, C.; Ren, J.; Zhao, X.; Sun, R.; Wu, A. An efficient pretreatment for the selectively
hydrothermal conversion of corncob into furfural: The combined mixed ball milling and ultrasonic
pretreatments. Ind. Crops Prod. 2016, 94, 721–728. [CrossRef]
183. Paze, A.; Brazdausks, P.; Rizhikovs, J.; Puke, M.; Tupciauskas, R.; Andzs, M.; Meile, K.; Vedernikovs, N.
Changes in the Polysaccharide Complex of Lignocellulose after Catalytic Hydrothermal Pre-Treatment
Process of Hemp (Cannabis Sativa L.) Shives. In Proceedings of the 23th European Biomass Conference and
Exhibition, Vienna, Austria, 1–4 June 2015; pp. 1063–1069. [CrossRef]
184. Codignole Luz, F.; Volpe, M.; Fiori, L.; Manni, A.; Cordiner, S.; Mulone, V.; Rocco, V. Spent coffee enhanced
biomethane potential via an integrated hydrothermal carbonization-anaerobic digestion process. Bioresour.
Technol. 2018, 256, 102–109. [CrossRef]
185. Passos, F.; Ferrer, I. Influence of hydrothermal pretreatment on microalgal biomass anaerobic digestion and
bioenergy production. Water Res. 2015, 68, 364–373. [CrossRef]
186. Weide, T.; Brügging, E.; Wetter, C. Anaerobic and aerobic degradation of wastewater from hydrothermal
carbonization (HTC) in a continuous, three-stage and semi-industrial system. J. Environ. Chem. Eng. 2019, 7,
102912. [CrossRef]
187. He, L.; Huang, H.; Zhang, Z.; Lei, Z.; Lin, B. Le Energy Recovery from Rice Straw through Hydrothermal
Pretreatment and Subsequent Biomethane Production. Energy Fuels 2017, 31, 10850–10857. [CrossRef]
188. Lucian, M.; Volpe, M.; Merzari, F.; Wüst, D.; Kruse, A.; Andreottola, G.; Fiori, L. Hydrothermal carbonization
coupled with anaerobic digestion for the valorization of the organic fraction of municipal solid waste.
Bioresour. Technol. 2020, 314, 123734. [CrossRef]
189. Ischia, G.; Orlandi, M.; Fendrich, M.A.; Bettonte, M.; Merzari, F.; Miotello, A.; Fiori, L. Realization of a solar
hydrothermal carbonization reactor: A zero-energy technology for waste biomass valorization. J. Environ.
Manage. 2020, 259, 110067. [CrossRef] [PubMed]
190. Xiao, C.; Liao, Q.; Fu, Q.; Huang, Y.; Chen, H.; Zhang, H.; Xia, A.; Zhu, X.; Reungsang, A.; Liu, Z.
A solar-driven continuous hydrothermal pretreatment system for biomethane production from microalgae
biomass. Appl. Energy 2019, 236, 1011–1018. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
30
energies
Article
Hydrothermal Carbonization as a Strategy for Sewage
Sludge Management: Influence of Process
Withdrawal Point on Hydrochar Properties
Fabio Merzari 1 , Jillian Goldfarb 2 , Gianni Andreottola 1 , Tanja Mimmo 3 , Maurizio Volpe 1,4 and
Luca Fiori 1, *
1 Department of Civil, Environmental and Mechanical Engineering, University of Trento, via Mesiano 77,
38123 Trento, Italy; fabio.merzari@unitn.it (F.M.); gianni.andreottola@unitn.it (G.A.);
maurizio.volpe@unikore.it (M.V.)
2 Department of Biological and Environmental Engineering, Cornell University, 226 Riley-Robb Hall, Ithaca,
NY 14853, USA; jlg459@cornell.edu
3 Faculty of Science and Technology, Free University of Bolzano, Piazza Università 5, 39100 Bolzano, Italy;
tanja.mimmo@unibz.it
4 Faculty of Engineering and Architecture, University of Enna Kore, Cittadella Universitaria, 94100 Enna, Italy
* Correspondence: luca.fiori@unitn.it; Tel.: +39-0461-282692
Abstract: Conventional activated sludge systems, still widely used to treat wastewater, produce
large amounts of solid waste that is commonly landfilled or incinerated. This study addresses the
potential use of Hydrothermal Carbonization (HTC) to valorize sewage sludge residues examining the
properties of hydrochars depending on HTC process conditions and sewage sludge withdrawal point.
With increasing HTC severity (process residence time and temperature), solid yield, total Chemical
Oxygen Demand (COD) and solid pH decrease while ash content increases. Hydrochars produced
from primary (thickened) and secondary (digested and dewatered) sludge show peculiar distinct
properties. Hydrochars produced from thickened sludge show good fuel properties in terms of
Higher Heating Value (HHV) and reduced ash content. However, relatively high volatile matter and
O:C and H:C ratios result in thermal reactivity significantly higher than typical coals. Both series
of carbonized secondary sludges show neutral pH, low COD, enhanced phosphorous content and
low heavy metals concentration: as a whole, they show properties compatible with their use as
soil amendments.
Keywords: sewage sludge; hydrothermal carbonization; hydrochar; solid biofuel; soil amendment
1. Introduction
In 1991, the European Union (EU) Directive 91/271/EEC set new benchmarks for the collection,
treatment and monitoring of wastewater in urban areas [1]. In 2000, the EU produced over 10 million
dry tons of sewage sludge (latest available official EU data [2]). Since then, production has steadily
increased, increasing burdens on municipal wastewater treatment systems [3]. In 2017, the global
sewage sludge production reached approximately 45 million of dry tons per year [4]. In Italy alone,
over 1100 kilotons of dry sludge matter were produced in 2010, with almost 40% going to landfills and
less than 30% used in agricultural applications [2]. In 2016, the European Commission reported that
the 1991 directive was successful in terms of current improvements seen in EU water quality despite
an increasing population growth. However, the European Commission also underlined the need to
develop “innovative solutions to increase resource efficiency, such as solutions for energy recovery,
nutrient recovery, and processing to marketable products and water re-use” [5]. Beyond the EU, global
urbanization and the growth of the middle class, combined with stricter environmental regulations,
31
Energies 2020, 13, 2890
have forced municipalities to re-examine their sludge management practices—transitioning away from
traditional disposal via incineration, landfilling, or discharge to oceans/waterways—to favor beneficial
reuse [6,7].
Sludge is a complex, heterogeneous mixture comprised of organic compounds such as proteins,
peptides, polysaccharides, phenols, aliphatic, aromatic and furan compounds, as well as inorganic
materials such as nutrients (phosphorous, potassium, nitrogen), silica, and heavy metals [8] and
pathogens and other microbiological pollutants [9]. In a typical wastewater treatment process, primary
sludge is characterized as the sludge following mechanical processing (screening, grit removal,
sedimentation) containing between 93 and 99.5 wt% water, with a high content of suspended and
dissolved organics. Secondary sludge (also known as Waste Activated Sludge, WAS) follows from
biological treatment and contains a high amount of microbial cells with a total solids concentration
ranging between 0.8 and 1.2 wt% depending on the type of process used [10].
Anaerobic Digestion (AD) is the most widely used sludge management technique. AD converts
the organic solids to biogas (predominantly CH4 and CO2 ) via hydrolysis, acidogenesis, acetogenesis,
and methanogenesis [11,12]. Despite its popularity, a considerable amount of solid remains after AD;
as little as 20–30 wt% of the total organic matter is mineralized [9]. While these solids were once
thought to be environmentally benign, their use into the soil may well be a potentially large source of
greenhouse gas emissions and point pollutant sources for mercury, lead, cadmium, and copper on
arable land, contributing to environmental acidification [13] and posing a toxicological risk in terms
of pathogens present [9]. The microbiological processes occurring on land-applied sludge (mainly
anaerobic decomposition, nitrification, and denitrification) lead to considerable emissions of methane
and nitrous oxide, as well as ammonia and nitrate. While land application of sludge offsets the
use of industrial nitrogen-containing fertilizers and represents a considerable net reduction of N2 O,
NH3 , and NO3 − emissions, CH4 emissions are still estimated to be 6.3 kgton−1 of applied sludge [13].
As such, direct land application of secondary sludge may not be the optimal nutrient recovery pathway
and may well represent a waste of a renewable energy source.
Hydrothermal Carbonization (HTC) is a process to concentrate the carbon in a given biomass,
occurring in water at elevated temperatures in the range of 160–280 ◦ C [14], above saturated vapor
pressure, where water’s dielectric constant decreases so drastically that it catalyzes the carbonization of
biomass while acting as an organic solvent [15,16]. HTC proceeds via a series of mechanisms, including
hydrolysis, dehydration, decarboxylation, decarbonylation, and demethanation. HTC is performed
in 80–95 vol% water [17–20], making it an ideal processing pathway for wet biomasses [21,22] such
as sewage sludge. As well recognized in the literature, HTC not only leads to reduced volume and
energy densification of the solid residue, considerably improving its dewaterability [23,24], but also
significantly increases its solid fuel properties while stabilizing and disinfecting the sludge [25,26].
Interestingly, HTC of sludge enhances NO + NH3 reactions during solid combustion, significantly
reducing NOx emissions across combustion modes [27].
While multiple studies focused on nutrients recovery such as phosphorous [28–30] or probed
the impact of processing conditions on the solid and energy yields of hydrochars produced from the
HTC of sludge from one withdrawal point in the wastewater treatment (WWT) process [23,31–39],
the impact of upstream WWT processes on hydrochars is often not part of the experimental design.
Thus, a primary goal of the present work was to determine the most viable point in the WWT process
to employ HTC for sludge treatment from a solid fuel production and potential nutrient recovery
standpoint. There are two sets of variables of interest in the present work: (1) feedstock withdrawal
point from the WWT process; (2) hydrothermal reaction conditions on the composition and oxidation
properties of resulting hydrochars. While some research has been done to quantify the environmental
and economic benefits of using HTC to treat sewage sludge, understanding the impact of both feedstock
and processing conditions on resulting hydrochars is critical to enable better optimization of this
waste-to-energy conversion pathway [40].
32
Energies 2020, 13, 2890
Figure 1. Schematic of wastewater treatment plant in Trento, Italy, identifying three sludge samples
(Thickened—Pipe 3; Digested—Pipe 4; Dewatered—Pipe 5) used in the present work.
33
Energies 2020, 13, 2890
Masshydrochar,dry
Yhydrochar = (2)
MassSludge,dry
The liquid yield was computed as the complement to 1 of the gas and solid yields.
34
Energies 2020, 13, 2890
The solid hydrochar was characterized according to the same methods described in Section 2.1.
for Higher Heating Value (HHV), proximate, and ultimate analyses. The hydrochar’s relative solid
reactivity was measured using a Mettler-Toledo Thermogravimetric Analyzer–Differential Scanning
Calorimeter (TGA-DSC-1, Mettler-Toledo LLC, Columbus, Ohio, USA) in an oxidative atmosphere.
The TGA-DSC was calibrated with NIST-traceable gold, indium, and aluminum and the mass was
measured to ±0.1 μg and temperature to ±0.1 ◦ C. Approximately 10 mg of sample was loaded into a
70 μL alumina crucible. Samples were heated at 20 ◦ C min−1 up to 110 ◦ C in air flowing at 50 mL min−1
and held for 30 min to drive off any residual moisture. They were subsequently heated at 20 ◦ C min−1
up to 950 ◦ C and held for 30 min to oxidize all material. The mass fraction of sample converted (X) at
any time, t, was calculated as:
m − mt
X= i (3)
mi − m f
where mi is the initial mass, mt is the mass at any time, t, and mf is the final mass after the hold at
950 ◦ C. Derivative thermogravimetric (DTG) curves were plotted as dX/dt (s−1 ) versus temperature.
Differential scanning calorimeter (DSC) data was normalized as heat flow per sample mass at any
given instant (mt ). DTG curves are compared to those from an in-house sample of Illinois No. 6 coal,
a high volatile bituminous coal from the Illinois #6 (Herrin) seam from the Argonne Premium Coal
Bank [47]. The coal sample is well characterized in the literature and is often used as a standard on
which to compare solid fuel oxidation [47–49].
Dewaterability—and the improvement due to HTC was determined by measuring the Capillary
Suction Time (CST) required for water to be separated from sludge across a filter paper (Whatman 17
CHR, VWR International, Milan, Italy) using a Triton Electronics Ltd. capillary suction timer type 304B
according to standard methods [43]. CST provides a quantitative assessment of how readily sludge
releases water.
The liquid phase remaining after hydrothermal carbonization was characterized by measuring
pH, COD, organic nitrogen, ammonia nitrogen, and phosphorous as described above. Measurements
of Readily Biodegradable COD (RBCOD) were performed following the procedure described in
literature [50]. To measure NH4 + nitrogen and soluble COD, samples were screened through a 0.45 μm
filter [43].
3. Results
To assess the optimal point to withdraw sludge from the wastewater treatment process for
hydrothermal carbonization in terms of resulting hydrochar properties, three samples were pulled
from various points along the process: Thickened, Digested, and Dewatered (Figure 1). The feedstocks
characteristics are presented in Tables 1 and 2. These three samples were subjected to hydrothermal
carbonization at three temperatures (190 ◦ C, 220 ◦ C, 250 ◦ C) and two residence times (30 min, 60 min)
each, producing a total of 18 hydrochar samples for analysis.
Looking at the data of the different raw sludges, it is clear that they differ substantially. Thickened
sludge contains about 46 wt% elemental carbon and 15 wt% ash. In digested sludge, the carbon
decreases to about 26 wt%, and the ash content increases to 45 wt%, due to stabilization during
anaerobic digestion. Dewatered sludge contains 36 wt% carbon and about 28 wt% ash. Thus, even if
dewatering is a mechanical process, it greatly modifies the sludge characteristics. The supernatant from
sludge dewatering has high concentrations of inorganic compounds such as N-NH4 + , P compounds,
CaCO3 , Mg, K, Na, and other minerals that contribute to the ash content [51]. The dewatering unit
washes away these inorganics strongly decreasing the ash content of the dewatered sludge: this reflects
on an increase in elemental C, H, N, and O, and also in FC and VM. VM variation is extremely
significant, passing from a value of 50 wt% in the digested sludge to about 66 wt% in the dewatered
sludge. These differences are also due to the fact that the digested sludge, immediately upstream
of the dewatering operation, is chemically conditioned with organic polyelectrolyte (1–10 g/kg dry
solids [52]) that is quickly adsorbed on the sludge particles.
35
Table 1. Characteristics of raw thickened sludge, raw digested sludge, and raw dewatered sludge and products of hydrothermal carbonization of thickened, digested,
and dewatered sludge.
36
Raw 175.00 ± 9.00 187.00 ± 9.00 10.00 ± 1.00 a
190 32.5 ± 0.1 38.4 ± 0.1 19.6 ± 0.3 22.7 ± 0.1 0.3 ± 0.0 0.4 ± 0.0
Total Phosphorous in Liquid (mg/L)
220 20.2 ± 0.1 16.4 ± 0.2 19.8 ± 0.3 17.8 ± 0.1 0.2 ± 0.0 0.2 ± 0.0
250 16.1 ± 0.1 11.1 ± 0.2 19.3 ± 0.3 12.2 ± 0.1 0.2 ± 0.0 0.2 ± 0.0
Raw 5.22 ± 0.10 5.81 ± 0.13 9.22 ± 0.33
190 6.2 ± 0.7 6.8 ± 0.4 7.2 ± 0.7 7.4 ± 1.4 9.2 ± 0.5 9.4 ± 0.4
Total Phosphorous in Solid (mg/g)
220 7.8 ± 0.3 9.4 ± 0.2 7.5 ± 0.5 7.5 ± 1.5 10.5 ± 0.2 10.6 ± 0.2
250 9.9 ± 0.5 10.2 ± 4.6 7.4 ± 0.9 6.7 ± 1.2 10.8 ± 0.8 10.9 ± 0.9
Raw 0.470 ± 0.230 0.880 ± 0.010 11.023 ± 0.551 a
190 0.1 ± 0.0 0.1 ± 0.0 1.3 ± 0.0 1.2 ± 0.0 4.5 ± 1.0 4.1 ± 0.2
Organic Nitrogen in Liquid (g/L)
220 0.1 ± 0.0 0.1 ± 0.0 1.4 ± 0.0 0.9 ± 0.0 2.4 ± 0.6 1.5 ± 0.1
250 0.1 ± 0.0 0.1 ± 0.0 1.3 ± 0.0 1.5 ± 0.0 1.7 ± 0.3 1.4 ± 0.7
Raw 0.490 ± 0.010 0.880 ± 0.010 0.598 ± 0.030 a
190 0.4 ± 0.0 0.6 ± 0.0 0.7 ± 0.0 0.7 ± 0.0 2.7 ± 0.1 3.8 ± 0.2
Ammonia Nitrogen in Liquid (g/L)
220 0.4 ± 0.0 0.5 ± 0.0 0.8 ± 0.0 1.0 ± 0.0 4.4 ± 0.3 4.6 ± 0.6
250 0.6 ± 0.0 0.6 ± 0.0 1.0 ± 0.0 0.8 ± 0.0 6.9 ± 0.4 6.5 ± 0.4
aras received, n.a.—not applicable, a Based on 1g of dry biomass dissolved in one liter of distilled water, HTC—Hydrothermal Carbonization.
Table 2. Products yields and analysis of solid fuel characteristics of raw feedstocks and hydrochars from thickened, digested, and dewatered sludge.
Solid Yield (wt%) 220 60.6 ± 8.8 57.2 ± 3.4 64.6 ± 6.4 70.6 ± 3.4 75.1 ± 1.0 75.2 ± 0.3
250 49.4 ± 4.8 52.0 ± 6.2 78.1 ± 7.8 64.9 ± 3.7 67.6 ± 2.5 66.8 ± 0.2
190 2.5 ± 0.0 2.7 ± 0.2 3.7 ± 0.9 4.7 ± 0.3 2.1 ± 1.4 2.6 ± 0.2
Gas Yield (wt%) 220 4.1 ± 0.7 4.2 ± 0.4 5.1 ± 1.3 5.6 ± 1.0 3.7 ± 0.1 4.3 ± 0.6
250 6.3 ± 1.6 7.2 ± 1.3 6.1 ± 2.1 7.6 ± 1.9 5.1 ± 0.2 5.9 ± 0.1
190 20.3 ± 2.4 34.9 ± 5.4 43.2 ± 4.9 43.1 ± 4.6 45.2 ± 1.7 44.1 ± 0.1
Liquid Yield (wt%) 220 35.3 ± 4.7 30.7 ± 1.9 34.8 ± 3.9 38.1 ± 2.2 39.4 ± 0.6 39.7 ± 0.4
250 44.3 ± 3.2 29.6 ± 3.8 42.1 ± 5.0 36.3 ± 2.8 36.4 ± 1.3 36.4 ± 0.2
Ultimate Analysis (dry basis)
Raw 45.96 ± 0.20 45.96 ± 0.20 25.60 ± 0.33 25.60 ± 0.33 35.91 ± 0.25 35.91 ± 0.25
190 44.56 ± 0.39 46.11 ± 0.39 19.22 ± 1.37 14.16 ± 2.57 36.61 ± 0.02 35.07 ± 0.23
C (wt%)
220 44.86 ± 0.51 43.15 ± 0.19 11.70 ± 0.64 10.21 ± 0.58 35.19 ± 0.17 35.75 ± 0.31
37
250 41.68 ± 0.29 41.21 ± 0.52 12.51 ± 0.26 12.02 ± 0.05 35.30 ± 0.18 35.57 ± 0.08
Raw 6.57 ± 0.02 6.57 ± 0.02 3.96 ± 0.06 3.96 ± 0.06 5.42 ± 0.00 5.42 ± 0.00
190 6.24 ± 0.08 6.46 ± 0.06 2.63 ± 0.16 2.00 ± 0.29 4.92 ± 0.02 4.61 ± 0.03
H (wt%)
220 5.95 ± 0.01 5.69 ± 0.02 1.61 ± 0.08 1.37 ± 0.07 4.32 ± 0.01 4.36 ± 0.03
250 5.03 ± 0.05 5.01 ± 0.03 1.65 ± 0.01 1.56 ± 0.02 4.11 ± 0.00 4.05 ± 0.01
Raw 4.26 ± 0.00 4.26 ± 0.00 3.59 ± 0.10 3.59 ± 0.10 5.81 ± 0.02 5.81 ± 0.02
190 2.23 ± 0.07 2.10 ± 0.07 1.58 ± 0.07 1.07 ± 0.17 4.27 ± 0.06 3.95 ± 0.01
N (wt%)
220 1.87 ± 0.24 1.86 ± 0.07 0.80 ± 0.08 0.66 ± 0.01 3.48 ± 0.02 3.45 ± 0.02
250 1.89 ± 0.05 1.99 ± 0.04 0.70 ± 0.00 0.69 ± 0.02 3.16 ± 0.04 3.12 ± 0.02
Raw 27.61 ± 4.86 27.61 ± 4.86 21.16 ± 0.79 21.16 ± 0.79 23.52 ± 0.52 23.52 ± 0.52
190 28.52 ± 6.69 25.97 ± 4.63 18.43 ± 3.50 15.67 ± 6.89 18.55 ± 0.17 19.33 ± 0.58
O (wt%)
220 25.51 ± 6.34 26.43 ± 5.06 12.35 ± 1.03 18.93 ± 1.10 16.53 ± 0.33 15.34 ± 0.49
250 22.47 ± 6.44 21.59 ± 3.83 11.72 ± 2.00 7.17 ± 2.53 14.07 ± 0.44 14.12 ± 2.18
Table 2. Cont.
190 4.25 ± 2.50 4.58 ± 1.75 1.70 ± 0.09 2.06 ± 0.34 5.37 ± 0.23 9.94 ± 6.93
Fixed Carbon (wt%)
220 5.37 ± 4.00 5.23 ± 3.97 1.26 ± 0.07 0.13 ± 0.01 4.04 ± 2.06 9.83 ± 5.58
250 5.31 ± 5.85 6.12 ± 5.93 0.62 ± 0.05 1.38 ± 0.09 6.27 ± 0.61 7.02 ± 0.03
Raw 72.48 ± 3.19 72.48 ± 3.19 50.25 ± 3.51 50.25 ± 3.51 65.67 ± 0.50 65.67 ± 0.50
190 70.84 ± 6.65 71.03 ± 5.36 41.09 ± 3.22 30.84 ± 4.20 58.98 ± 0.29 53.02 ± 6.61
Volatile Matter (wt%)
220 66.05 ± 5.56 63.57 ± 7.81 26.17 ± 1.59 31.99 ± 0.90 55.48 ± 2.19 49.07 ± 5.72
250 60.06 ± 2.19 56.70 ± 3.95 25.97 ± 1.68 21.65 ± 4.68 50.37 ± 0.39 49.85 ± 2.03
Raw 14.96 ± 2.31 14.96 ± 2.31 44.99 ± 3.65 44.99 ± 3.65 28.43 ± 0.56 28.43 ± 0.56
190 24.91 ± 4.57 24.40 ± 3.56 57.21 ± 1.61 67.10 ± 2.27 35.66 ± 0.26 37.05 ± 6.77
Ash (Inorganic) (wt%)
220 28.58 ± 4.78 31.20 ± 5.89 72.57 ± 0.80 67.88 ± 0.45 40.48 ± 2.13 41.10 ± 5.65
250 34.63 ± 4.02 37.18 ± 4.94 73.42 ± 0.86 76.97 ± 2.34 43.36 ± 0.50 43.14 ± 1.03
Combustion Analysis
38
Raw 20.50 ± 0.14 20.50 ± 0.14 10.66 ± 1.79 10.66 ± 1.79 16.02 ± 0.09 16.02 ± 0.09
190 19.45 ± 0.64 20.71 ± 0.03 9.27 ± 1.73 7.97 ± 1.77 16.30 ± 0.19 15.96 ± 0.07
HHV (MJ/kg)
220 20.06 ± 0.90 18.72 ± 0.03 8.96 ± 1.77 7.86 ± 0.39 15.70 ± 0.09 15.47 ± 0.37
250 18.66 ± 0.26 19.17 ± 0.61 8.59 ± 1.78 9.37 ± 1.39 15.98 ± 0.11 15.33 ± 0.26
Energies 2020, 13, 2890
3.1. Hydrochar Properties Depend on Sludge Withdrawal Point and Carbonization Conditions
Solid hydrochar yields and properties are strongly influenced by both sewage sludge feedstock
and HTC temperature and residence time, as shown in Table 2 (see also Supplementary Materials
Figure S1). In general, as the harshness of carbonization increases, the solid hydrochar yield decreases,
as is the case with biomasses across the literature [53]. The one exception to this general trend was
the digested sludge sample carbonized at 220 ◦ C for 30 min, which had a lower solid yield than
its 250 ◦ C and 190 ◦ C counterparts. This anomaly was observed for multiple experimental runs.
Such behavior could be attributed to potential re-condensing of tarry materials onto the hydrochar
matrix, increasing the observed “solid” yield at 250 ◦ C. Such materials were previously identified on
the surface of heterogeneous biomasses that can be maximized at carbonization conditions specific to
each biomass [17]. As the solid hydrochar is collected on a 10–100 μm filter, any particles smaller than
this size range would be considered part of the liquid yield, which may also explain this anomaly.
As a result of the carbonization process, which releases organic compounds from the solid matrix
and forces them into the water phase while generating CO2 , the total COD present substantially
decreased upon carbonization (at any time, temperature) of the dewatered sludge sample, with a
similar behavior noted for the thickened and digested samples (the effect on these two samples is less
pronounced given the lower initial concentration in the raw samples) as shown in Table 1 and Figure S2
in Supplementary Materials. Conversely, soluble COD increased with increasing severity, most notably
for the thickened and digested samples, with a significantly higher soluble COD that is maintained
throughout carbonization for the dewatered samples. Such behavior has been described in the literature
for other biomasses, including sludge, secondary sludge, wood waste, and dairy waste [54]. Given
the nature of the carbonization process, where water acts as both a solvent and transport medium,
simultaneously drawing organics into the water phase and generating CO2 , this COD mass balance
suggests that indeed the insoluble fraction is being converted to CO2 (and lesser amounts of CO and
other non-condensable gases) while the soluble portion is retained in the process water.
The decrease in hydrochar pH for increasingly severe carbonization of the thickened sludge is
to be expected; Volatile Fatty Acids (VFAs) present in primary sludge and formed via hydrolysis of
triacylglycerols [55], especially acetic acid, are known to remain stable in the liquid phase, only breaking
down at higher carbonization temperatures [54]. It was found here that the pH for the thickened
sludge dropped from 7.2 (raw) to 6.0 for both the 190 ◦ C and 220 ◦ C conditions and to 5.8 for the
250 ◦ C condition of the 30 min HTC samples and to 5.3, 6.1, and 6.0 for the 190 ◦ C, 220 ◦ C, and 250 ◦ C
60 min samples, respectively (Table 1; plotted in Figure S3 in Supplementary Materials). This suggests
that carbonization is able to release some VFAs from the primary matrix, but that the overall acidic
nature of the hydrochar increases slightly. It should be noted that these measurements were taken in
the same manner used to measure biochar pH, by equilibrating the hydrochar with deionized water
and reading the pH of the liquid following settling. Anecdotally, when the pH of the process water is
measured immediately following carbonization, the pH increases slightly upon severe carbonization,
which would indicate hydrolysis of released VFAs occurs in the process water. For the digested
samples, carbonization at 250 ◦ C yields hydrochars with slightly higher pH than their 190 ◦ C and
220 ◦ C counterparts, suggesting a reduced acid content of the char. For the dewatered secondary
sludge, the pH decreases with increasing carbonization but is overall lowest to begin with, suggesting
that the acid content decreased during anaerobic digestion.
The proposed loss of semi-volatile compounds from the sludge matrices upon carbonization is
further supported by the proximate analysis (Figure 2; Table 2; data of Table 2 plotted in Figure S4 in
Supplementary Materials). The sludge treatment of WWTP results in a decrease in fixed carbon and
volatile matter content upon digestion, with a slight increase in both upon dewatering. Changes in
proximate analysis as a function of carbonization conditions versus the raw sludge withdrawn at each
point are given in Figure 2. Figure 2 reports the percentage change in value of the various variables
(FC, VM, Ash, all data on a dry basis) due to HTC, in respect to the raw biomasses. As compared
to the raw sludge sample from each withdrawal point, the volatile matter content decreased with
39
Energies 2020, 13, 2890
increasingly harsh carbonization. The fixed carbon content also decreases upon HTC, shifting the
balance of the hydrochar composition to the inorganic (loosely termed “ash” phase). The change
in proximate analysis appears to be more heavily temperature dependent than time dependent for
the thickened samples. In this case, the ash content increased by 50 wt%, 100 wt%, and 150 wt%
over the raw thickened sample as the carbonization temperature increased from 190 ◦ C to 220 ◦ C to
250 ◦ C, respectively. This increase in inorganic content is offset more so by decreases in fixed carbon
than volatile matter. For the digested samples, the increase in ash content—ranging from 25 wt% to
50 wt% as temperature increases—is offset by decreases in both VM and FC, with a slightly higher
decrease in FC than VM across HTC conditions. The behavior of the dewatered samples is somewhat
different. While for HTC at 30 min there is a slight increase in ash (especially at 220 ◦ C and 250 ◦ C
carbonization), the 220 ◦ C 30 min sample is offset more by changes in FC, whereas the 250 ◦ C 60 min
sample’s higher ash content is due to a decrease in VM. The dewatered 60 min HTC samples show
even further divergent behavior; for all temperatures, the FC and ash contents increase, offset by
a decrease in VM content. For the 190 ◦ C and 220 ◦ C samples, the change in FC weight percent is
actually greater than that of the ash, whereas the 250 ◦ C sample has a larger increase in ash than fixed
carbon. This is likely due to the preparation of the samples. Both thickened and digested samples were
used as received. The dewatered samples had too low of a moisture content to ensure all solids were
submerged in the reactor, such that deionized water was added to the samples prior to carbonization.
It is possible that some of the inorganics migrated into the water phase (to establish a concentration
equilibrium), and thus, the relative change in ash content was lower for these samples.
Figure 2. Changes in volatile matter, fixed carbon, and ash composition as a result of hydrothermal
carbonization: (a) Thickened sludge, 30 min of Hydrothermal Carbonization (HTC); (b) Thickened
sludge, 60 min HTC; (c) Digested sludge, 30 min HTC; (d) Digested sludge, 60 min HTC; (e) Dewatered
sludge, 30 min HTC—note 190 ◦ C Fixed Carbon (FC) and Volatile Matter (VM) data points overlap; (f)
Dewatered sludge, 60 min HTC.
This behavior is echoed by the ultimate analysis, whereby HTC temperature has a greater impact
on organic element composition than time (Figure 3; Figure 4; Table 2; data of Table 2 plotted in Figure S5
in Supplementary Materials). As expected, the biological transformations of the primary sludge lead
to overall lower oxygen and carbon contents of the raw digested and dewatered samples [56]. Figure 3
40
Energies 2020, 13, 2890
reports the percentage change in content of atomic species C, H, N, and O (all data on a dry basis)
due to HTC, in respect to the raw biomasses. As shown in Figure 3, in general the weight percent of
the organic elements decrease upon carbonization of the sludge samples, the balance made up by an
increasing (relative) inorganic content. The elemental carbon contents for the thickened 190 ◦ C 30 min
and 60 min and all dewatered samples remain constant within statistical significance. Interestingly,
the relative percent changes of C, H, and O were fairly minimal for the thickened sludge carbonized at
190 ◦ C and 220 ◦ C, suggesting minimal decarboxylation and dehydration, both of which would lower
the O:C and H:C ratios, respectively [57]. However, this is not to imply a stagnant system; for the
thickened (and indeed all sludge samples) the nitrogen content of the solids decreases considerably
upon even mild carbonization, ranging from hydrochars with 40 wt% to those with 80 wt% lower N
content than their raw counterparts (Figure 3). This loss of nitrogen supports the hypothesis that there
is significant hydrolysis of fat and protein components present in the sludge. The organic nitrogen in
the process liquid tends to decrease by 15–85 wt% upon carbonization of all samples. The ammonia
nitrogen present in the liquid varies in a certain range, without any clear trend with time or temperature,
as shown in Table 1. Given the substantial elemental decrease in N of the hydrochar, and liquid-phase
organic nitrogen upon carbonization, closure of a mass balance would suggest either production of
nitrogen gases (N2 , NOx ) or formation of nitrate or nitrite in the liquid phase, which could precipitate
out as salts onto the hydrochar. As reported by Kruse and co-workers, this concentration would be too
low to detect with N content measurements [58].
Figure 3. Changes in ultimate analysis as a result of hydrothermal carbonization: (a) Thickened sludge,
30 min HTC; (b) Thickened sludge, 60 min HTC; (c) Digested sludge, 30 min HTC; (d) Digested sludge,
60 min HTC; (e) Dewatered sludge, 30 min HTC; (f) Dewatered sludge, 60 min HTC.
41
Energies 2020, 13, 2890
Figure 4. Van Krevelen diagram of raw and hydrothermally carbonized sludge samples with:
(a) all samples; (b) thickened sludge; (c) dewatered sludge; (d) digested sludge.
The carbon and fixed carbon values reported in Table 2 and the trends of the same variables
shown in Figures 2 and 3 testify to the peculiarity of these kinds of substrates which behave differently
during HTC in respect to the vast majority of the other biomasses. While HTC applied to agro-waste,
lignocellulosic feedstock, organic fraction of municipal solid waste, or compost [20,45,53,59,60] results
in an increase in the values of C and FC, this is not the case for the majority of the sludge samples
here investigated. The FC content decreases after HTC for thickened and digested sludges (Figure 2).
The C content in the hydrochars is equal to or lower than C content in the raw sludges (Figure 3).
This apparently strange behavior depends on the very high ash content of the sludge samples and
was also previously reported in the literature [31,57,61]: in all these cases, the C and FC contents were
expressed on a dry basis. Reverse C and FC trends were reported when the data was expressed on a
dry ash free (daf) basis [23]. Here, on a daf basis, thickened and dewatered sludge increase their C
content after HTC, and this applies to the digestate treated at the highest HTC temperature of 250 ◦ C,
too; FC data, conversely, even on a daf basis, remain still lower in the hydrochars in respect to the
parent biomasses. To summarize, the organics in the sludge carbonized as should be expected in an
HTC process: the increase in C content is evident when the basis of calculation is represented by the
organics themselves (i.e., data on a daf basis), and the HTC was sufficiently severe. On a dry basis,
conversely, the relative increase in ash content often prevails over the relative increase in C, as recently
also discussed by Ferrentino et al. [62].
3.2. Distribution of Nutrients and Inorganics: Potential for Use as Soil Amendment
As shown in Table 1, the total phosphorous concentrated in the solid hydrochar
(all samples/processing conditions) from 5 wt% in the raw samples to up to 10 wt% in the carbonized
samples, with the thickened and dewatered 250 ◦ C 30 min and 60 min samples having the highest
amounts. The thickened sludge (all times/temperatures) showed lower organic nitrogen concentrations
in the liquid phase, suggesting it remains in the solid phase or exits as a gas, as mentioned previously.
This, coupled with the higher FC and VM contents of the thickened and dewatered hydrochars,
suggests they would make better soil amendments than the digested sludge hydrochars [63–65].
42
Energies 2020, 13, 2890
Thermally treated biomasses tend to show enhanced P fertilizer values as a result of various
mechanisms, including structural surface changes and improved association of P to inorganics such
as Mg, Ca, and Al [66]. In the present work, the type of sludge sample had a larger impact on the
retention and concentration of these nutrients than the carbonization conditions. As shown in Figure 5,
the dewatered samples had higher overall Al concentrations—from 4 mgAl /gsludge for the raw sample
as compared to ~ 2 mgAl /gsludge for the thickened and digested raw samples. Carbonization at 220
and 250 ◦ C for both 30 min and 60 min doubled all of these concentrations, to 8 mgAl /gsludge for the
carbonized dewatered sample and ~ 4 mgAl /gsludge for the others. Conversely, the digested sludge
had almost twice the concentration of magnesium than the thickened or dewatered samples (which
again almost doubled upon carbonization at 220 ◦ C and 250 ◦ C). The calcium concentration was quite
similar for all sludge samples at 23, 33, and 28 mgCa /gsludge , for the thickened, digested, and dewatered
raw samples, respectively. Upon carbonization at 250 ◦ C for 30 min, the Ca content increased by over
100% for the thickened sample but only by 39% and 43% for the digested and dewatered raw samples,
respectively. While soil incubation studies are beyond the scope of the present work, prior work by
Thomsen and co-workers [66] suggests that the more heavily oxidized hydrochar samples (e.g., those
that released more CO2 and thus had higher gas yield) containing more Mg, Ca, and Al would have
a higher P availability. This corresponds to the digested 250 ◦ C 60 min and dewatered 220 ◦ C and
250 ◦ C for 30 min and 60 min samples. Future work will investigate the degree to which hydrothermal
carbonization plays a role in P plant availability, as well as the impact of HTC processing conditions
itself on the volatilization and re-condensation of P species on the hydrochar surfaces.
Figure 5. Impact of hydrothermal carbonization on nutrient and heavy metals concentrations for:
(a) Al; (b) Ca; (c) Mg; (d) K; (e) Ag; (f) V; (g) Cr; (h) Ni; (i) Co; (j) Mo; (k) As; (l) Pb.
43
Energies 2020, 13, 2890
The concentrations of a series of additional inorganic elements in the sludges and hydrochars were
measured. Figure 5 shows a representative set of these elements (all data available in Supplementary
Materials: Tables S1 and S2, Figures S6 and S7). Silver was included in the analysis, given its
increasing prevalence in consumer materials as an incorporated nanomaterial and resulting detection
in wastewater treatment systems [67]. Prior work indicates that silver will more likely accumulate
in the biosolids than in the WWTP effluent [68]. Here, considerably higher levels of silver in the
digested sludge are found as compared to the thickened or dewatered sludge, suggesting that the
digestion indeed concentrates the silver in the biosolid, but that the dewatering process shifts the
silver to the liquid phase. The silver concentration in the digested hydrochar decreases with increasing
carbonization. Both of these indicate that the silver is in an easily mobile state in the digested sludge,
equilibrating with its aqueous phase to lower solid concentration.
Arsenic and vanadium were included in this analysis given their high prevalence in Italian
drinking water sources; the acceptable limit for arsenic in the Trentino-Alto Adige region increases
to 50 μg L−1 over the nationwide 10 μg L−1 limit due to natural lithology [69]. As seen in Figure 5,
the arsenic concentrations in the solid samples decreased with increasing harshness for the thickened
and digested samples (with the exception of the 190 ◦ C and 250 ◦ C, 60 min digested samples). All other
hydrochars had arsenic concentrations below 2 μg g−1 . The International Biochar Initiative (IBI) [70]
suggests an acceptable range for As in biochars of between 13 and 100 μg g−1 if they are to be applied
as a soil amendment.
In Sicily, concentrations of V in drinking water sources are routinely above E.U. and U.S. maximum
contaminant levels (MCLs) due to the underlying geology [71–73]. As such, it is important to ensure
that concentrations of these metals were below MCLs before considering the potential for hydrochars to
be land-applied, which could exacerbate the metal contamination issue. The vanadium concentration
was slightly increased upon carbonization, but never exceeded 10 μgg−1 ; the MCL for drinking water
in Italy is 50 μgL−1 . V is known to co-precipitate with iron (III) [74]; while the oxidation states
are not known here, the concentrations of iron in all solids exceeded 2 mgg−1 (therefore higher by
several orders of magnitude), such that it is possible that V is present in the hydrochars in an iron
(hydr)oxide precipitate.
As shown in Figure 5, the concentrations of chromium, cobalt, nickel, molybdenum, and lead
(and a series of additional inorganics, as given in SI) are only modestly affected by carbonization and
sludge withdrawal point. The cobalt concentration of the digested sludge was notably higher than
the thickened or dewatered, but upon carbonization reached the same levels as the other samples of
hydrochars, all at less than 3 μgg−1 . The concentrations of all of these metals are below IBI maximum
allowable thresholds [70], making them reasonable candidates for use as soil amendments on the basis
of heavy metal content.
3.3. Energy Content and Oxidative Reactivity: Potential for Use as Solid Fuel
As Table 2 shows, the solid hydrochars have similar higher heating values to their raw counterparts.
In this case, HTC has not substantially improved the “energy density” of the solid fuel. However,
HTC does enable a more efficient solid–liquid separation and dewaterability. For the raw dewatered
sludge, CST was not complete after 60 min, while it was complete for the 190 ◦ C 60 min hydrochar
after only 380 s. With an increase in HTC temperature, CST was 209 and 90 s at 220 ◦ C and 250 ◦ C
(60 min), respectively. Images of the CST trials are available in Supplementary Materials Figure S8.
This improved dewaterability demonstrates a reduction of volume for transport of the solid waste and
reduced moisture content for potential combustion applications.
Figure 6 plots the DTG curves of the hydrochar samples produced at each of the three carbonization
temperatures at 30 min, alongside an Illinois No. 6 coal sample. As can be seen, the hydrochar
sludge samples are considerably more reactive than the coal sample. Their peak DTG temperatures
(highest conversion rates) occur at hundreds of degrees less than the coal sample and at higher
conversion rates. The highest peak mass loss rates occur for all three hydrochars produced from
44
Energies 2020, 13, 2890
the thickened sludge, prior to any anaerobic digestion. Post-digestion, while the shape of the DTG
curves changes, the peak rates are quite similar for both the thickened and dewatered sludge. For all
three samples, the 190 ◦ C carbonized hydrochars display the highest reactivity compared to the other
HTC temperatures. Especially at such mild conditions, HTC does not significantly carbonize the
sample—oftentimes, the original materials microstructure is preserved, whereas higher temperatures
lead to a more complete destruction of the carbon matrix [53,75].
Figure 6. Derivative thermogravimetric curves of oxidation: (a) thickened sludge and hydrochars;
(b) digested sludge and hydrochars; (c) dewatered sludge and hydrochars, produced at 30 min alongside
Illinois No 6. Coal.
Given the relatively high reactivity of any of the sludge hydrochars, it may be difficult to combust
them for electricity generation in current boilers designed for solid fuels such as coal [76]. While the
higher heating values of especially thickened sludge (18–21 MJ kg−1 ) are suitable for such combustion
schemes, their lower ignition and peak reactivity temperatures are considerably lower than that of
most bituminous coals, and therefore may lead to loss of efficiency in boilers [77]. That, combined with
the higher ash content that could result in slagging and fouling, suggests that the sludge hydrochars
may perform better in co-combustion scenarios [78].
It was previously shown that blending biofuels with similar characteristics at ratios less than 20 wt%
with coal mitigates fuel segregation and efficiency loss issues while increasing the share of renewables
in energy generation portfolios [48,49,76]. Recent work in the literature suggests that hydrochars can
be co-combusted with a variety of coals in economically, environmentally, and energetically viable
schemes in existing infrastructure [79–81] and may even improve the emissions profile at optimized
blending ratios [82]. It was recently demonstrated that hydrochars with similar reactivities can be
oxidized with Illinois No. 6 coal at ratios of 10 wt% hydrochar, balance coal, without causing significant
fuel segregation [83]. In summation, the high reactivity and ash content (though high HHV) of the
thickened and dewatered hydrochars temper enthusiasm for their use as a combustible fuel. The low
HHV of the digested sludge—below low-rank coals—makes it difficult to envision a combustion
45
Energies 2020, 13, 2890
scenario where this would be a valuable solid fuel. As such, the potential for using sludge-based
hydrochars as a drop-in solid fuel is likely minimal, though this could be accomplished without the
need for anaerobic digestion of the sludge if used in a co-fired fuel scenario.
4. Conclusions
The present study probed the impact of hydrothermal processing conditions and sewage sludge
withdrawal point on resulting hydrochars. The hydrothermal carbonization (HTC) of sludge proceeds
similarly to many other wet biomasses; as harshness (time and temperature) of the process increases,
the solid yield decreases, the ash (inorganic) content increases, total COD decreases but soluble COD
increases, and solid pH decreases. However, there are distinct differences between the hydrochars
produced from primary (thickened) versus secondary (digested and dewatered) sludge. Thickened
46
Energies 2020, 13, 2890
sludge carbonized at moderate conditions (220 ◦ C, 30 min) produced the most viable solid fuel with
the highest HHV, moderate ash content, and high volatile matter content. However, with O:C and
H:C ratios higher than typical bituminous coals of similar heating content, the thermal reactivity of
the hydrochar was significantly higher than coals typically combusted. This suggests that sludge
hydrochars could be co-fired with coal but are not ideal solid fuels. On the other hand, hydrochars
produced from secondary sludge are more viable as potential soil amendments. The carbonized
digested sludges show relatively neutral pH, low COD, and enhanced phosphorous, along with
enhanced Ca, Mg, and Al concentrations to help mobilize P. Their heavy metal composition is well
below International Biochar Initiative standards, though the elemental oxygen content and lower
volatile matter content warrant future inquiry into this pathway.
References
1. Steichen, R. Council Directive of 21 May 1991 concerning urban waste water treatment (91/271/EEC). Off. J.
Eur. Communities 1991. Available online: https://www.tarimorman.gov.tr/SYGM/Belgeler/ab%20mevzuat%
C4%B1/91-271-EEC.pdf (accessed on 4 June 2020).
2. Bianchini, A.; Bonfiglioli, L.; Pellegrini, M.; Saccani, C. Sewage sludge management in Europe: A critical
analysis of data quality. Int. J. Environ. Waste Manag. 2016, 18, 226. [CrossRef]
3. EUROSTAT. Sewage Sludge Production and Disposal from Urban Wastewater (in Dry Substance (d.s)). 2017.
Available online: https://data.europa.eu/euodp/en/data/dataset/hzWkcfKt5mxEaFijeoA (accessed on
4 June 2020).
4. Zhang, Q.; Hu, J.; Lee, D.; Chang, Y.; Lee, Y. Sludge treatment: Current research trends. Bioresour. Technol.
2017, 243, 1159–1172. [CrossRef]
5. European Commission. Eighth Report on the Implementation Status and the Programmes for Implementation
(as required by Article 17) of Council Directive 91/271/EEC Concerning Urban Waste Water Treatment; European
Commission: Brussels, Belgium, 2016.
6. Yanagida, T.; Fujimoto, S.; Minowa, T. Application of the severity parameter for predicting viscosity during
hydrothermal processing of dewatered sewage sludge for a commercial PFBC plant. Bioresour. Technol. 2010,
101, 2043–2045. [CrossRef]
47
Energies 2020, 13, 2890
7. Campbell, H.W.; Pacific, C.; Technologies, E. Sludge management—Future issues and trends.
Water Sci. Technol. 2000, 41, 1–8. [CrossRef]
8. European Commission. Disposal and Recycling Routes for Sewage Sludge Part 1—Sludge Use Acceptance Report;
European Commission: Brussels, Belgium, 2001; ISBN 9289417986.
9. Rulkens, W. Sewage Sludge as a Biomass Resource for the Production of Energy: Overview and Assessment
of the Various Options. Energy Fuels 2008, 44, 9–15. [CrossRef]
10. Tyagi, V.K.; Lo, S.L. Sludge: A waste or renewable source for energy and resources recovery? Renew. Sustain.
Energy Rev. 2013, 25, 708–728. [CrossRef]
11. Cao, Y.; Pawłowski, A. Sewage sludge-to-energy approaches based on anaerobic digestion and pyrolysis:
Brief overview and energy efficiency assessment. Renew. Sustain. Energy Rev. 2012, 16, 1657–1665. [CrossRef]
12. Merzari, F.; Langone, M.; Andreottola, G.; Fiori, L. Methane production from process water of sewage sludge
hydrothermal carbonization. A review. Valorising sludge through hydrothermal carbonization. Crit. Rev.
Environ. Sci. Technol. 2019, 49, 947–988. [CrossRef]
13. Johansson, K.; Perzon, M.; Fröling, M.; Mossakowska, A.; Svanström, M. Sewage sludge handling with
phosphorus utilization—Life cycle assessment of four alternatives. J. Clean. Prod. 2008, 16, 135–151. [CrossRef]
14. Merzari, F.; Lucian, M.; Volpe, M.; Andreottola, G.; Fiori, L. Hydrothermal carbonization of biomass: Design
of a bench-Scale reactor for evaluating the heat of reaction. Chem. Eng. Trans. 2018, 65, 43–48.
15. Akiya, N.; Savage, P.E. Roles of water for chemical reactions in high-temperature water. Chem. Rev. 2002,
102, 2725–2750. [CrossRef] [PubMed]
16. Kritzer, P. Corrosion in high-temperature and supercritical water and aqueous solutions: A review.
J. Supercrit. Fluids 2004, 29, 1–29. [CrossRef]
17. Lucian, M.; Volpe, M.; Gao, L.; Piro, G.; Goldfarb, J.L.; Fiori, L. Impact of hydrothermal carbonization
conditions on the formation of hydrochars and secondary chars from the organic fraction of municipal solid
waste. Fuel 2018, 233, 257–268. [CrossRef]
18. Lucian, M.; Fiori, L. Hydrothermal carbonization of waste biomass: Process design, modeling, energy
efficiency and cost analysis. Energies 2017, 10, 211. [CrossRef]
19. Volpe, M.; Messineo, A.; Mäkelä, M.; Barr, M.R.; Volpe, R.; Corrado, C.; Fiori, L. Reactivity of cellulose during
hydrothermal carbonization of lignocellulosic biomass. Fuel Process. Technol. 2020, 206, 106456. [CrossRef]
20. Volpe, M.; Fiori, L.; Volpe, R.; Messineo, A. Upgrading of Olive Tree Trimmings Residue as Biofuel by
Hydrothermal Carbonization and Torrefaction: A Comparative Study. Chem. Eng. Trans. 2016, 50, 13–18.
21. Mäkelä, M.; Volpe, M.; Volpe, R.; Fiori, L.; Dahl, O. Spatially resolved spectral determination of polysaccharides
in hydrothermally carbonized biomass. Green Chem. 2018, 20, 1114–1120. [CrossRef]
22. Lucian, M.; Volpe, M.; Fiori, L. Hydrothermal Carbonization Kinetics of Lignocellulosic Agro-Wastes:
Experimental Data and Modeling. Energies 2019, 12, 516. [CrossRef]
23. Zhao, P.; Shen, Y.; Ge, S.; Yoshikawa, K. Energy recycling from sewage sludge by producing solid biofuel
with hydrothermal carbonization. Energy Convers. Manag. 2014, 78, 815–821. [CrossRef]
24. Danso-Boateng, E.; Holdich, R.G.; Wheatley, A.D.; Martin, S.J.; Shama, G. Hydrothermal Carbonization of
Primary Sewage Sludge and Synthetic Faeces: Effect of Reaction Temperature and Time on Filterability.
Environ. Prog. Sustain. Energy 2015, 34, 1279–1290. [CrossRef]
25. Khalil, W.A.S.; Shanableh, A.; Rigby, P.; Kokot, S. Selection of hydrothermal pre-treatment conditions of
waste sludge destruction using multicriteria decision-making. J. Environ. Manag. 2005, 75, 53–64. [CrossRef]
26. Catallo, W.J.; Comeaux, J.L. Reductive hydrothermal treatment of sewage sludge. Waste Manag. 2008, 28,
2213–2219. [CrossRef] [PubMed]
27. Zhao, P.; Chen, H.; Ge, S.; Yoshikawa, K. Effect of the hydrothermal pretreatment for the reduction of NO
emission from sewage sludge combustion. Appl. Energy 2015, 111, 199–205. [CrossRef]
28. Becker, G.C.; Wüst, D.; Köhler, H.; Lautenbach, A.; Kruse, A. Novel approach of phosphate-reclamation as
struvite from sewage sludge by utilising hydrothermal carbonization. J. Environ. Manag. 2019, 238, 119–125.
[CrossRef] [PubMed]
29. Shi, Y.; Luo, G.; Rao, Y.; Chen, H.; Zhang, S. Hydrothermal conversion of dewatered sewage sludge: Focusing
on the transformation mechanism and recovery of phosphorus. Chemosphere 2019, 228, 619–628. [CrossRef]
30. Song, E.; Park, S.; Kim, H. Upgrading Hydrothermal Carbonization (HTC) Hydrochar from Sewage Sludge.
Energies 2019, 12, 2383. [CrossRef]
48
Energies 2020, 13, 2890
31. Danso-Boateng, E.; Shama, G.; Wheatley, A.D.; Martin, S.J.; Holdich, R.G. Hydrothermal carbonisation
of sewage sludge: Effect of process conditions on product characteristics and methane production.
Bioresour. Technol. 2015, 177, 318–327. [CrossRef]
32. Brookman, H.; Gievers, F.; Zelinski, V.; Ohlert, J.; Loewen, A. Influence of Hydrothermal Carbonization on
Composition, Formation and Elimination of Biphenyls, Dioxins and Furans in Sewage Sludge. Energies 2018,
11, 1582. [CrossRef]
33. Breulmann, M.; Schulz, E.; Van Afferden, M.; Müller, R.A. Hydrochars derived from sewage sludge: Effects
of pre-treatment with water on char properties, phytotoxicity and chemical structure. Arch. Agron. Soil Sci.
2017, 64, 860–872. [CrossRef]
34. Liu, T.; Liu, Z.; Zheng, Q.; Lang, Q.; Xia, Y.; Peng, N. Effect of hydrothermal carbonization on migration and
environmental risk of heavy metals in sewage sludge during pyrolysis. Bioresour. Technol. 2018, 247, 282–290.
[CrossRef]
35. Chen, C.; Liu, G.; An, Q.; Lin, L.; Shang, Y.; Wan, C. From wasted sludge to valuable biochar by low
temperature hydrothermal carbonization treatment: Insight into the surface characteristics. J. Clean. Prod.
2020, 263, 1–9. [CrossRef]
36. Zhai, Y.; Peng, C.; Xu, B.; Wang, T.; Li, C. Hydrothermal carbonisation of sewage sludge for char production
with different waste biomass: Effects of reaction temperature and energy recycling. Energy 2017, 127, 167–174.
[CrossRef]
37. Wang, R.; Wang, C.; Zhao, Z.; Jia, J.; Jin, Q. Energy recovery from high-ash municipal sewage sludge by
hydrothermal carbonization: Fuel characteristics of biosolid products. Energy 2019, 186, 115848. [CrossRef]
38. Zheng, X.; Jiang, Z.; Ying, Z.; Song, J.; Chen, W.; Wang, B. Role of feedstock properties and hydrothermal
carbonization conditions on fuel properties of sewage sludge-derived hydrochar using multiple linear
regression technique. Fuel 2020, 271, 1–11. [CrossRef]
39. Xu, Z.-X.; Song, H.; Li, P.-J.; He, Z.-X.; Wang, Q.; Wang, K.; Duan, P.-G. Hydrothermal carbonization of
sewage sludge: Effect of aqueous phase recycling. Chem. Eng. J. 2020, 387, 1–12. [CrossRef]
40. Vardon, D.R.; Sharma, B.K.; Scott, J.; Yu, G.; Wang, Z.; Schideman, L.; Zhang, Y.; Strathmann, T.J. Chemical
properties of biocrude oil from the hydrothermal liquefaction of Spirulina algae, swine manure, and digested
anaerobic sludge. Bioresour. Technol. 2011, 102, 8295–8303. [CrossRef]
41. Hao, X.D.; Li, J.; van Loosdrecht, M.C.M.; Li, T.Y. A sustainability-based evaluation of membrane bioreactors
over conventional activated sludge processes. J. Environ. Chem. Eng. 2018, 6, 2597–2605. [CrossRef]
42. Manser, R.; Gujer, W.; Siegrist, H. Membrane bioreactor versus conventional activated sludge system:
Population dynamics of nitrifiers. Water Sci. Technol. 2005, 52, 417–425. [CrossRef]
43. APHA. Standard Methods for the Examination of Water and Wastewater, 22th ed.; American Public Health
Association: Washington, DC, USA, 2012.
44. Fiori, L.; Basso, D.; Castello, D.; Baratieri, M. Hydrothermal carbonization of biomass: Design of a batch
reactor and preliminary experimental results. Chem. Eng. Trans. 2014, 37, 55–60.
45. Basso, D.; Weiss-Hortala, E.; Patuzzi, F.; Castello, D.; Baratieri, M.; Fiori, L. Hydrothermal carbonization of
off-specification compost: A byproduct of the organic municipal solid waste treatment. Bioresour. Technol.
2015, 182, 217–224. [CrossRef]
46. Basso, D.; Patuzzi, F.; Castello, D.; Baratieri, M.; Rada, C.E.; Weiss-Hortala, E.; Fiori, L. Agro-industrial waste
to solid biofuel through hydrothermal carbonization. Waste Manag. 2016, 47, 114–121. [CrossRef] [PubMed]
47. Hunt, J.E. Argonne Premium Coal Sample Program; Argonne: Lemont, IL, USA, 2007. Available online:
https://publications.anl.gov/anlpubs/2007/04/58856.pdf (accessed on 4 June 2020).
48. Celaya, A.A.M.; Lade, A.T.A.; Goldfarb, J.J.L. Co-combustion of brewer’s spent grains and Illinois No. 6 coal:
Impact of blend ratio on pyrolysis and oxidation behavior. Fuel Process. Technol. 2015, 129, 39–51. [CrossRef]
49. Xue, J.; Chellappa, T.; Ceylan, S.; Goldfarb, J.L. Enhancing biomass + coal Co-firing scenarios via biomass
torrefaction and carbonization: Case study of avocado pit biomass and Illinois No. 6 coal. Renew. Energy
2018, 122, 152–162. [CrossRef]
50. Andreottola, G.; Foladori, P.; Ferrai, M.; Ziglio, G. Respirometria Applicata Alla Depurazione Delle Acque; Principi
e metodi; Department of Civil, Environmental and Mechanical Enginnering, University of Trento: Trento,
Italy, 2002.
49
Energies 2020, 13, 2890
51. Ren, W.; Zhou, Z.; Jiang, L.M.; Hu, D.; Qiu, Z.; Wei, H.; Wang, L. A cost-effective method for the
treatment of reject water from sludge dewatering process using supernatant from sludge lime stabilization.
Sep. Purif. Technol. 2015, 142, 123–128. [CrossRef]
52. Al Momani, F.A.; Örmeci, B. Optimization of polymer dose based on residual polymer concentration in
dewatering supernatant. Water Air Soil Pollut. 2014, 225. [CrossRef]
53. Volpe, M.; Goldfarb, J.L.; Fiori, L. Hydrothermal carbonization of Opuntia ficus indica cladodes: Role of
process parameters on hydrochar properties. Bioresour. Technol. 2018, 247, 310–318. [CrossRef]
54. Jomaa, S.; Shanableh, A.; Khalil, W.; Trebilco, B. Hydrothermal decomposition and oxidation of the organic
component of municipal and industrial waste products. Adv. Environ. Res. 2003, 7, 647–653. [CrossRef]
55. Kocsisová, T.; Juhasz, J.; Cvengroš, J. Hydrolysis of fatty acid esters in subcritical water. Eur. J. Lipid
Sci. Technol. 2006, 108, 652–658. [CrossRef]
56. Díaz, E.; Pintado, L.; Faba, L.; Ordóñez, S.; González-LaFuente, J.M. Effect of sewage sludge composition on
the susceptibility to spontaneous combustion. J. Hazard. Mater. 2019, 361, 267–272. [CrossRef]
57. Berge, N.D.; Ro, K.S.; Mao, J.; Flora, J.R.V.; Chappell, M.A.; Bae, S. Hydrothermal Carbonization of Municipal
Waste Streams. Environ. Sci. Technol. 2011, 45, 5696–5703. [CrossRef]
58. Kruse, A.; Koch, F.; Stelzl, K.; Wüst, D.; Zeller, M. Fate of Nitrogen during Hydrothermal Carbonization.
Energy Fuels 2016, 30, 8037–8042. [CrossRef]
59. Volpe, M.; Fiori, L. From olive waste to solid biofuel through hydrothermal carbonisation: The role
of temperature and solid load on secondary char formation and hydrochar energy properties. J. Anal.
Appl. Pyrolysis 2017, 124, 63–72. [CrossRef]
60. Volpe, M.; Wüst, D.; Merzari, F.; Lucian, M.; Andreottola, G.; Kruse, A.; Fiori, L. One stage olive mill waste
streams valorisation via hydrothermal carbonisation. Waste Manag. 2018, 80, 224–234. [CrossRef] [PubMed]
61. Ekpo, U.; Ross, A.B.; Camargo-Valero, M.A.; Williams, P.T. A comparison of product yields and inorganic
content in process streams following thermal hydrolysis and hydrothermal processing of microalgae, manure
and digestate. Bioresour. Technol. 2016, 200, 951–960. [CrossRef] [PubMed]
62. Ferrentino, R.; Ceccato, R.; Marchetti, V.; Andreottola, G.; Fiori, L. Sewage Sludge Hydrochar: An Option for
Removal of Methylene Blue from Wastewater. Appl. Sci. 2020, 10, 3445. [CrossRef]
63. Zornoza, R.; Moreno-Barriga, F.; Acosta, J.A.; Muñoz, M.A.; Faz, A. Stability, nutrient availability and
hydrophobicity of biochars derived from manure, crop residues, and municipal solid waste for their use as
soil amendments. Chemosphere 2016, 144, 122–130. [CrossRef]
64. Wu, H.; Lai, C.; Zeng, G.; Liang, J.; Chen, J.; Xu, J.; Dai, J.; Li, X.; Liu, J.; Chen, M.; et al. The interactions of
composting and biochar and their implications for soil amendment and pollution remediation: A review.
Crit. Rev. Biotechnol. 2017, 37, 754–764. [CrossRef]
65. Melo, T.M.; Bottlinger, M.; Schulz, E.; Leandro, W.M.; Filho, A.M.D.A.; Wang, H.; Ok, Y.S.; Rinklebe, J. Plant
and soil responses to hydrothermally converted sewage sludge (sewchar). Chemosphere 2018, 206, 338–348.
[CrossRef] [PubMed]
66. Thomsen, P.T.; Zsuzsa, S.; Ahrenfeldt, J.; Henriksen, U.B.; Frandsen, F.J.; Müller-st, D.S. Changes imposed
by pyrolysis, thermal gasi fi cation and incineration on composition and phosphorus fertilizer quality of
municipal sewage sludge. J. Environ. Manag. 2017, 198, 308–318. [CrossRef]
67. Dwivedi, A.D.; Dubey, S.P.; Sillanp, M.; Kwon, Y.N.; Lee, C.; Varma, R.S. Fate of engineered nanoparticles:
Implications in the environment. Coord. Chem. Rev. 2015, 287, 64–78. [CrossRef]
68. Wang, Y.; Westerhoff, P.; Hristovski, K.D. Fate and biological effects of silver, titanium dioxide,
and C60(fullerene) nanomaterials during simulated wastewater treatment processes. J. Hazard. Mater.
2012, 201–202, 16–22. [CrossRef] [PubMed]
69. Dinelli, E.; Lima, A.; Albanese, S.; Birke, M.; Cicchella, D.; Giaccio, L.; Valera, P.; De Vivo, B. Major and trace
elements in tap water from Italy. J. Geochem. Explor. 2012, 112, 54–75. [CrossRef]
70. International Biochar Initiative. Standardized Product Definition and Product Testing Guidelines for Biochar That
Is Used in Soil. 2015. Available online: https://www.biochar-international.org/wp-content/uploads/2018/04/
IBI_Biochar_Standards_V2.1_Final.pdf (accessed on 4 June 2020).
71. Giammanco, S.; Valenza, M.; Pignato, S.; Giammanco, G. Mg, Mn, Fe, and V Concentration in the Ground
Waters of Mount Etna (Sicily). Water Res. 1996, 30, 378–386. [CrossRef]
72. Roccaro, P.; Barone, C.; Mancini, G.; Vagliasindi, F.G.A. Removal of manganese from water supplies intended
for human consumption: A case study. Desalination 2007, 210, 205–214. [CrossRef]
50
Energies 2020, 13, 2890
73. Varrica, D.; Tamburo, E.; Dongarrà, G. Sicilian bottled natural waters: Major and trace inorganic components.
Appl. Geochem. 2013, 34, 102–113. [CrossRef]
74. Roccaro, P.; Vagliasindi, F.G.A. Coprecipitation of vanadium with iron(III) in drinking water: A pilot-scale
study. Desalin. Water Treat. 2015, 55, 799–809. [CrossRef]
75. Kruse, A.; Funke, A.; Titirici, M.-M. Hydrothermal conversion of biomass to fuels and energetic materials.
Curr. Opin. Chem. Biol. 2013, 17, 515–521. [CrossRef]
76. Goldfarb, J.L.; Liu, C. Impact of blend ratio on the co-firing of a commercial torrefied biomass and coal via
analysis of oxidation kinetics. Bioresour. Technol. 2013, 149, 208–215. [CrossRef]
77. Khan, A.A.; de Jong, W.; Jansens, P.J.; Spliethoff, H. Biomass combustion in fluidized bed boilers: Potential
problems and remedies. Fuel Process. Technol. 2009, 90, 21–50. [CrossRef]
78. Haykiri-Acma, H.; Yaman, S.; Kucukbayrak, S. Does carbonization avoid segregation of biomass and lignite
during co-firing? Thermal analysis study. Fuel Process. Technol. 2015, 137, 312–319. [CrossRef]
79. Cartmell, E.; Gostelow, P.; Riddell-black, D.; Simms, N.; Oakey, J.; Morris, J.O.E.; Jeffrey, P.; Howsam, P.
Biosolids-A Fuel or a Waste? An Integrated Appraisal of Five Co-combustion Scenarios with Policy Analysis.
Environ. Sci. Technol. 2006, 40, 649–658. [CrossRef]
80. Xiao, H.; Ma, X.; Liu, K. Co-combustion kinetics of sewage sludge with coal and coal gangue under different
atmospheres. Energy Convers. Manag. 2010, 51, 1976–1980. [CrossRef]
81. Otero, M.; Calvo, L.F.; Gil, M.V.; Garcia, A.I.; Moran, A. Co-combustion of different sewage sludge and coal:
A non-isothermal thermogravimetric kinetic analysis. Bioresour. Technol. 2008, 99, 6311–6319. [CrossRef]
82. Parshetti, G.K.; Liu, Z.; Jain, A.; Srinivasan, M.P.; Balasubramanian, R. Hydrothermal carbonization of
sewage sludge for energy production with coal. Fuel 2013, 111, 201–210. [CrossRef]
83. Gao, L.; Volpe, M.; Lucian, M.; Fiori, L.; Goldfarb, J.L. Does hydrothermal carbonization as a biomass
pretreatment reduce fuel segregation of coal-biomass blends during oxidation? Energy Convers. Manag. 2019,
181, 93–104. [CrossRef]
84. Fregolente, L.G.; Miguel, T.B.A.R.; de Castro Miguel, E.; de Almeida Melo, C.; Moreira, A.B.; Ferreira, O.P.;
Bisinoti, M.C. Toxicity evaluation of process water from hydrothermal carbonization of sugarcane industry
by-products. Environ. Sci. Pollut. Res. 2019, 26, 27579–27589. [CrossRef] [PubMed]
85. Barlindhaug, J.; Ødegaard, H. Thermal hydrolysis for the production of carbon source for denitrification.
Water Sci. Technol. 1996, 34, 371–378. [CrossRef]
86. Friedman, A.A.; Smith, J.E.; DeSantis, J.; Ptak, T.; Ganley, R.C. Characteristics of Residues from Wet Air
Oxidation of Anaerobic Sludges. J. (Water Pollut. Control Fed.) 1988, 60, 1971–1978.
87. Wirth, B.; Reza, T.; Mumme, J. Influence of digestion temperature and organic loading rate on the continuous
anaerobic treatment of process liquor from hydrothermal carbonization of sewage sludge. Bioresour. Technol.
2015, 198, 215–222. [CrossRef]
88. Aragón-briceño, C.; Ross, A.B.; Camargo-valero, M.A. Evaluation and comparison of product yields and
bio-methane potential in sewage digestate following hydrothermal treatment. Appl. Energy 2017, 208,
1357–1369. [CrossRef]
89. Villamil, J.A.; Mohedano, A.F.; Rodriguez, J.J.; de la Rubia, M.A. Valorisation of the liquid fraction from
hydrothermal carbonisation of sewage sludge by anaerobic digestion. J. Chem. Technol. Biotechnol. 2018, 93,
450–456. [CrossRef]
90. Qiao, W.; Peng, C.; Wang, W.; Zhang, Z. Biogas production from supernatant of hydrothermally treated
municipal sludge by upflow anaerobic sludge blanket reactor. Bioresour. Technol. 2011, 21, 9904–9911.
[CrossRef]
91. Appels, L.; Dewil, R.; Baeyens, J.; Degrève, J. Ultrasonically enhanced anaerobic digestion of waste activated
sludge. Int. J. Sustain. Eng. 2008, 1, 94–104. [CrossRef]
92. Vavilin, V.A.; Rytov, S.V.; Lokshina, L.Y. A Description of Hydrolysis Kinetics in Aanaerobic Degradation of
Particulate Organic Matter. Bioresour. Technol. 1996, 56, 229–237. [CrossRef]
93. Nuchdang, S.; Frigon, J.; Roy, C.; Pilon, G.; Phalakornkule, C.; Guiot, S.R. Hydrothermal post-treatment of
digestate to maximize the methane yield from the anaerobic digestion of microalgae. Waste Manag. 2018, 71,
683–688. [CrossRef] [PubMed]
94. Stutzenstein, P.; Weiner, B.; Köhler, R.; Pfeifer, C.; Kopinke, F.D. Wet oxidation of process water from
hydrothermal carbonization of biomass with nitrate as oxidant. Chem. Eng. J. 2018, 339, 1–6. [CrossRef]
51
Energies 2020, 13, 2890
95. Luz, F.C.; Volpe, M.; Fiori, L.; Manni, A.; Cordiner, S.; Mulone, V.; Rocco, V. Spent Coffee Enhanced Biomethane
Potential via an Integrated Hydrothermal Carbonization-Anaerobic Digestion Process. Bioresour. Technol.
2018, 256, 102–109.
96. Singh, A.; Sharma, R.K.; Agrawal, M.; Marshall, F.M. Risk assessment of heavy metal toxicity through
contaminated vegetables from waste water irrigated area of Varanasi, India. Trop. Ecol. 2010, 51, 375–387.
97. Shi, L.; Zhang, G.; Wei, D.; Yan, T.; Xue, X.; Shi, S.; Wei, Q. Preparation and utilization of anaerobic granular
sludge-based biochar for the adsorption of methylene blue from aqueous solutions. J. Mol. Liq. 2014, 198,
334–340. [CrossRef]
98. Stark, K.; Plaza, E.; Hultman, B. Phosphorus release from ash, dried sludge and sludge residue from
supercritical water oxidation by acid or base. Chemosphere 2006, 62, 827–832. [CrossRef]
99. Yoshizaki, S.; Tomida, T. Principle and Process of Heavy Metal Removal from Sewage Sludge.
Environ. Sci. Technol. 2000, 34, 1572–1574. [CrossRef]
100. Hotova, G.; Slovak, V.; Soares, O.S.G.P.; Figueiredo, J.L.; Pereira, M.F.R. Oxygen surface groups analysis of
carbonaceous samples pyrolysed at low temperature. Carbon 2018, 134, 255–263. [CrossRef]
101. Liang, B.; Lehmann, J.; Solomon, D.; Kinyangi, J.; Grossman, J.; Skjemstad, J.O.; Thies, J.; Luiza, F.J.; Petersen, J.;
Neves, E.G. Black Carbon Increases Cation Exchange Capacity in Soils. Soil Sci. Soc. Am. J. 2006, 1719–1730.
[CrossRef]
102. Goldfarb, J.L.; Dou, G.; Salari, M.; Grinstaff, M.W. Biomass-Based Fuels and Activated Carbon Electrode
Materials: An Integrated Approach to Green Energy Systems. ACS Sustain. Chem. Eng. 2017, 5, 3046–3054.
[CrossRef]
103. Isitan, S.; Ceylan, S.; Topcu, Y.; Hintz, C.; Tefft, J.; Chellappa, T.; Guo, J.; Goldfarb, J.L. Product quality
optimization in an integrated biorefinery: Conversion of pistachio nutshell biomass to biofuels and activated
biochars via pyrolysis. Energy Convers. Manag. 2016, 127, 576–588. [CrossRef]
104. Goldfarb, J.L.; Buessing, L.; Gunn, E.; Lever, M.; Billias, A.; Casoliba, E.; Schievano, A.; Adani, F. Novel
Integrated Biorefinery for Olive Mill Waste Management: Utilization of Secondary Waste for Water Treatment.
ACS Sustain. Chem. Eng. 2017, 5, 876–884. [CrossRef]
105. Gopu, C.; Gao, L.; Volpe, M.; Fiori, L.; Goldfarb, J.L. Valorizing municipal solid waste: Waste to energy and
activated carbons for water treatment via pyrolysis. J. Anal. Appl. Pyrolysis 2018, 133, 48–58. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
52
energies
Article
Experimental and Computational Evaluation of
Heavy Metal Cation Adsorption for Molecular Design
of Hydrothermal Char
Louise Delahaye 1,† , John Thomas Hobson 1,† , Matthew Peter Rando 1 , Brenna Sweeney 1 ,
Avery Bernard Brown 1 , Geoffrey Allen Tompsett 1 , Ayten Ates 2 , N. Aaron Deskins 1 and
Michael Thomas Timko 1, *
1 Department of Chemical Engineering, Worcester Polytechnic Institute, Worcester, MA 01609, USA;
lcdelahaye@wpi.edu (L.D.); jthobson@wpi.edu (J.T.H.); mprando@wpi.edu (M.P.R.); bes117@pitt.edu (B.S.);
abbrown@wpi.edu (A.B.B.); gtompsett@wpi.edu (G.A.T.); nadeskins@wpi.edu (N.A.D.)
2 Department of Chemical Engineering, Engineering Faculty, Sivas Cumhuriyet University,
58140 Sivas, Turkey; aytates@gmail.com
* Correspondence: mttimko@wpi.edu
† These authors contributed equally to this work.
Abstract: A model hydrochar was synthesized from glucose at 180 ◦ C and its Cu(II) sorption
capacity was studied experimentally and computationally as an example of molecular-level adsorbent
design. The sorption capacity of the glucose hydrochar was less than detection limits (3 mg g−1 ) and
increased significantly with simple alkali treatments with hydroxide and carbonate salts of K and
Na. Sorption capacity depended on the salt used for alkali treatment, with hydroxides leading to
greater improvement than carbonates and K+ more than Na+ . Subsequent zeta potential and infrared
spectroscopy analysis implicated the importance of electrostatic interactions in Cu(II) sorption to the
hydrochar surface. Computational modeling using Density Functional Theory (DFT) rationalized
the binding as electrostatic interactions with carboxylate groups; similarly, DFT calculations were
consistent with the finding that K+ was more effective than Na+ at activating the hydrochar.
Based on this finding, custom-synthesized hydrochars were synthesized from glucose-acrylic acid
and glucose-vinyl sulfonic acid precursors, with subsequent improvements in Cu(II) adsorption
capacity. The performance of these hydrochars was compared with ion exchange resins, with the
finding that Cu(II)-binding site stoichiometry is superior in the hydrochars compared with the resins,
offering potential for future improvements in hydrochar design.
1. Introduction
According to the World Health Organization (WHO), approximately 785 million people lack access
to clean drinking water, mainly in poor countries [1], but also in some rural and even highly urbanized
areas [2] in the developed world. Heavy metal contamination of the water supply is a persistent
problem that dates back to antiquity [3], and some researchers speculate that lead contamination of
the water supply may have played a role in the downfall of the Roman empire [3]. More recently,
Fernández-Luqueño et al. [4] summarized the health effects of heavy metals, listing their contributions
to disease ranging from cancer to lung failure. Despite widespread acknowledgement of these negative
human health outcomes, providing uniform access to drinking water free of heavy metal contamination
has proven remarkably difficult to achieve, as evidenced by recent widely reported examples [2].
Different technologies have been developed to remove heavy metals from water [5] including
precipitation [5–7], sedimentation [8], flotation [8,9], membrane processes [8,10–12], electrochemical
53
Energies 2020, 13, 4203
processes [13–15], adsorption [16,17], and ion exchange [18,19]. Most of these technologies are
either costly, wasteful, reliant on non-renewable resources, energy inefficient, or unable to achieve
sufficiently low levels of metal concentration on their own. Of the available options, adsorption and
the kindred technique of ion exchange are energy efficient and capable of achieving suitably low
metal concentrations [20]. However, commercial adsorbents and ion exchange resins are derived
from non-renewable resources, either petrochemicals or coal, meaning that their use has associated
negative environmental impacts, creating a tradeoff between clean drinking water and mitigating
climate change [21]. Accordingly, development of renewable, cost-effective, and high-capacity metal
adsorbates has potential to greatly expand access to clean drinking water while minimizing other
negative environmental impacts [22,23].
Recently, pyrolysis biochar has emerged as a renewable sorbent for heavy metal removal [24,25].
Unlike most activated carbons, which are produced from coal, pyrolysis biochar is produced from
biomass or agricultural wastes [26]. In some cases, heavy metal capacity on biochar can reach
40 mg g−1 , which is comparable with activated carbons (20–80 mg g−1 ) [25] or ion exchange resin
capacity (20–30 mg g−1 ) [25]. Surface precipitation and electrostatic interactions are thought to be
the key adsorption mechanisms onto biochar, and abundant oxygenated functional groups (OFG)
are associated with effective cation adsorption [27,28]. Unfortunately, pyrolysis at typical conditions
(>400 ◦ C) tends to be ineffective for OFG formation, meaning that biochar must usually be activated to
increase its adsorption capacity [29]. Various activation procedures can increase the sorption capacity
of biochar, but only by adding cost, energy use, or waste generation [30]. Lastly, pyrolysis is performed
in the vapor phase and requires a dry feedstock, negatively impacting the process energy balance for
utilization of abundant wet wastes [31,32].
Hydrothermal carbonization (HTC) [33–35] of carbon-rich feeds, including carbohydrates, biomass,
and food waste, at moderate temperature (130–250 ◦ C) and autogenous pressures is a versatile,
low-energy, and renewable way to produce carbon-rich materials with abundant OFGs that are
known as hydrothermal chars (hydrochars) [36,37]. Possibly due to the abundance of OFGs present in
hydrochars, they have greater metal sorption capacity than most pyrolysis biochars, making them
especially attractive for drinking water purification [5,38,39]. For example, Regmi et al. [33] reported a
hydrochar with greater Cu(II) sorption capacity than conventional activated carbon (4.8 compared
with 1.8 mg g−1 ) [39]. Similarly, HTC compatibility with wet feeds eliminates the need for drying,
benefiting process-level energy balance especially for agricultural and food waste streams [40,41].
Although HTC is a promising technology, process costs and uncertainties must be reduced to
de-risk further investment [42]. Similarly, HTC can benefit by maximizing its value, which in the
case of sorbent manufacture, can be accomplished by maximizing hydrochar sorption capacity [16].
Accordingly, a persistent mystery in the field of hydrochar sorption is the high capacity that the material
has for metal cations despite its relatively low (<10 m2 g−1 ) measurable surface area. For this reason,
many reports describe methods to increase hydrochar surface area [29,37,43]. Unfortunately, hydrochar
activation to increase surface area again produces wastes and requires energy; moreover, the resulting
capacity of the activated material often decreases on a per area basis; for example, Jain et al. [44]
reported that the phenol capacity of activated carbon produced from hydrochar decreased with
increasing surface area, from 0.16 to 0.13 mg g−1 [16].
While pyrolytic treatment and various chemicals can increase hydrochar sorption capacity,
immersion in an alkali solution at room temperature reportedly increases OFG abundance and heavy
metal adsorption capacity with minimal energy requirements while generating minimal amounts of
waste [28,33,37]. The mechanism of the alkali-treatment promotion effect is not clear since mild alkali
treatment does not increase hydrochar surface area [45], and the conditions are not sufficient to make
or break covalent bonds. Moreover, alkali treatment is sometimes reported as a necessary step for
hydrochar to exhibit any heavy metal sorption capacity [29,37], while for others, alkali treatment is not
reported [33]. Understanding alkali treatment is, therefore, one goal of this work.
54
Energies 2020, 13, 4203
Complicating analysis further, hydrochar OFG abundance, and sorption capacity vary depending
on the properties of the feed and reaction conditions. Guo et al. [46,47] have shown that, starting with a
lignocellulosic raw material, a hydrochar can be produced with greater OFG abundance than produced
from carbohydrate precursors. For example, HTC from wood (initially consisting of 52.7 wt% C)
produces a hydrochar with 77 wt% C [48], whereas the carbon content of hydrochar produced from
glucose typically contains 53–62 wt% C [49]. Hydrochar OFG abundance tends to decrease with
increasing reaction conditions such as reaction time, temperature, and water-to-biomass ratio [46–48].
For many feeds [50–54], HTC can be performed over the temperature range from 160–180 ◦ C and
at reaction times <12 h to promote formation of OFGs. More aggressive conditions (≥200 ◦ C) are
preferred for other applications [55].
The relationship between feed properties, reaction conditions, and sorption capacities provides
an opportunity to synthesize a hydrochar tailor-made for a specific application, such as heavy metal
adsorption. For example, Demir-Cakan et al. [56] reported that co-HTC of glucose and acrylic acid
produced a hydrochar with exceptionally high OFG abundance; the resulting materials exhibited
sorption capacities up to 350 mg g−1 for Pb(II) and 90 mg g−1 for Cd(II). In comparison, Xue et al. [16]
reported peanut-hull based hydrochar sorption capacity for Cd(II) as 12.38 mg g−1 .
Ideally, hydrochars could be designed for a specific application at the molecular level.
Computational methods have proven effective at understanding sorption mechanisms and thereby
enabling the molecular level design of metal-organic frameworks for sorption of perfluoroalkyl
substances [57], nanopores for CO2 adsorption [58], and ion exchange resins and activated carbon
for heavy metal sorption [59,60]. Density Functional Theory (DFT) is an especially valuable tool for
studying the geometry and energetics of sorbate binding to the active site, provided that molecular
structures are known that can be used as targets of rational design. Unfortunately, the structural
models of hydrochar have only recently converged [61–63], making attempts at molecular level design
or computational modeling of sorbate-sorbent binding difficult until now.
Recent work by Brown et al. [61] reconciled several disparate models proposed for the structure of
hydrochar synthesized from glucose. Previous models inferred from infrared spectroscopy and Raman
microscopy [62,64] indicated that hydrochar structures resembled activated carbon and consisted of
fused aromatic cores, comprised of many aromatic rings and with OFGs present primarily as side chains.
In contrast, solid-state Nuclear Magnetic Resonance (NMR) [65] and Near-Edge X-ray Absorption
Fine Structure (NEXAFS) [63] indicate a structure consisting primarily of individual furan and arene
groups, polymerized via short alkyl chains, and decorated with OFGs. Brown et al. [61] recognized that
previously reported Raman spectra of hydrochar contained artifacts due to laser-induced pyrolysis of
the hydrochar material, causing it to collapse into a condensed aromatic structure. DFT simulation of
hydrochar Raman vibrations [61] then indicated that artifact-free Raman spectra were indeed consistent
with the furan-arene polymer previously inferred from NMR and NEXAFS [62–64,66]. This paves the
way for a molecular-level study of metal binding to hydrochar as a furan-rich polymer, thereby enabling
rational design of hydrochar.
The objective of this work was molecular-level design of a hydrochar adsorbent using both
experiments and simulation. To focus on generalizable mechanisms, we studied a model hydrochar
synthesized from glucose for sorption of a model heavy metal, Cu(II) cations. Sorption capacity
was studied before and after alkali treatment and compared with capacities measured for several
activated carbon materials. Similarly, the hydrochar was characterized for OFG type and density
using Fourier transforms-infrared spectroscopy (FT-IR), solid-state titration, and zeta potential
measurement. Metal-hydrochar binding interactions and geometries were evaluated using DFT
simulations. Chars with different types and/or densities of OFGs and other metal binding groups
were custom synthesized for comparison with glucose hydrochar. The experimental and simulation
results described here establish a new method for the rational design of hydrochar sorbents at the
molecular level.
55
Energies 2020, 13, 4203
2.1. Materials
All reagents were analytical grade, including: D-(+)-glucose (≥99.5%-Sigma Aldrich), acrylic acid
monomer (≥99.0%-TCI chemicals), vinyl sulfonic acid (≥99.0%-TCI Chemicals), hydrochloric acid
(0.1 M, ≥99.0%-Acros), sodium hydroxide (0.1 M, ≥95.0%), and the anhydrous salts sodium carbonate
(≥99.5%-Sigma Aldrich), sodium hydroxide (≥97.0%-EM Science), potassium carbonate (≥99.0%-Alfar
Aesar), and potassium hydroxide (≥85%-Sigma Aldrich). The carbon materials were donations from
Norit (since acquired by Cabot) and MeadWestvaco. Specific activated carbon samples included
wood-based carbons, Norit Darco® KB-G, Norit Darco® KB-WJ and Nuchar® (MeadWestvaco), and a
peat-based carbon, Norit® SX-1. Amberlyst® 15 was used in its hydrogen form and AG® 50 W-X4 in its
hydrogen form was purchased from Bio-Rad. Cu(NO3 )2 · 2.5 H2 O and elemental standards, 5% HNO3 ,
10 μg/mL were acquired from PerkinElmer. Deionized (DI) water was purified to a minimum resistivity
of 17.9 MΩ cm prior to use.
56
Energies 2020, 13, 4203
Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS) was performed on powder
samples using a Thermo-Fisher FT-IR 6700 with DRIFTS accessory, described previously in the
literature [69]. The spectral resolution was 2 cm−1 and all samples were purged with N2 gas for 2 min
before analysis to exclude atmospheric CO2 and H2 O from the sample space. A background spectrum
was obtained prior to each measurement and results were obtained by scanning 1024 times and taking
their average. Spectra were analyzed using MagicPlot software and plotted by normalization with
the baseline.
Zeta potential measurements were determined using a zeta meter (Malvern Zetasizer-Nano-Z)
that has been previously described elsewhere [70]. For each test, 0.005 g of the solid sample was
suspended in 100 cm3 of de-ionized water containing 0.1 N NaCl followed by homogenization for
2 h in an ultrasonic bath. After ultrasonication, the aqueous suspension was equilibrated at different
pH values for 30 min. Zeta potential results are reported as the average and standard deviation of
three measurements.
The combined densities of strong and weak acid groups were determined using the Boehm
titration method, described previously in the literature [29,71,72]. In brief, a carbon sample (0.5 g) was
placed in NaHCO3 solution (20 mL, 0.1 N), agitated for 48 h, and the carbon was removed by filtration.
The resulting filtrate was degassed for at least 30 min using N2 to remove CO2 and was then titrated to
determine the acid site density of carbon-rich materials [73].
Each adsorption measurement was performed at least in duplicate and ICP concentration
measurements were performed in triplicate. Average values are reported here. Control runs were
performed in the absence of sorbent and the loss to the vial was equivalent to <1 mg g−1 of sorbent.
57
Energies 2020, 13, 4203
configurations were studied to confirm that the final geometry captured a global minimum, rather than
a local one.
Table 1. Adsorption Capacity and Surface Area of Hydrochar and Activated Carbon. BET =
Brunauer–Emmett–Teller model.
Material Post Treatment Adsorption Capacity (mg Cu g−1 ) BET Surface Area (m2 g−1 )
None <3
Na2 CO3 20.0 ± 0.4
Glucose Hydrochar K2 CO3 29 ±5 4±2
NaOH 35 ±4
KOH 40 ±4
None 19 ±4
Norit® SX1 KOH 21 ±6
800 ± 100
None 29 ±3
Nuchar® KOH 11 ±7
1700 ± 200
Activated Carbons
None 9 ±0.1
Darco® KB-G KOH 5 ±0.8
1500 ± 100
Darco® None 6 ±0.2
1600 ± 100
KB-WJ KOH 21 ±4
Sun et al. [37] reported that alkali treatment increased the sorption capacity of hydrochar by
2–3 times, motivating the study of alkali treatment in the current study. As shown in Table 1,
alkali treatment greatly increased the Cu(II) capacity, by at least an order of magnitude compared with
the original glucose hydrochar. Several different bases were evaluated, with the finding that strong
bases (hydroxides) outperformed weak ones (carbonates) and that bases featuring the potassium cation
outperformed ones possessing sodium.
We compared the capacity of alkali-activated glucose hydrochar with several different activated
carbons (Table 1), selected to cover a range of properties [83]. Interestingly, the activated carbons
exhibited much greater sorption capacity than glucose hydrochar without activation, but less capacity
than their alkali-activated forms. Alkali treatment was evaluated for two of the activated carbons,
and it was found that the treatment either had no effect (Norit® SX1) or even negative effect (Nuchar® )
on sorption capacity. The different response to alkali treatment observed for activated carbon and
hydrochar clearly points to differences in the mechanism that must be understood for molecular-level
hydrochar design.
To understand the adsorption results presented in Table 1, surface areas were measured using N2
sorption and the BET isotherm fitting method. Consistent with previous reports [29,37,43], the measured
BET surface area of hydrochar was <10 m2 g−1 . Alkali treatment had no effect on the measured
58
Energies 2020, 13, 4203
hydrochar surface area, allowing us to reject the hypothesis that the effect of the treatment was to
open up the hydrochar pore structure [82,84]. Similarly, consistent with previous reports [83], the BET
surface areas of the activated carbons were >800 m2 g−1 and not affected by the dilute alkali treatment.
Accordingly, while the capacity of glucose hydrochar is comparable to activated carbon on a mass
basis, on a surface area basis the capacity is orders of magnitudes greater. This observation clearly
points to a specific hydrochar-sorbate interaction that can be engineered to maximize adsorption.
The strongest common sorbate-hydrochar interaction is electrostatic [85], which can be understood
as the interaction between the positively charged metal cation and negatively charged functional
groups on the hydrochar surface. Accordingly, as a way to understand and quantify hydrochar surface
charge, we measured hydrochar zeta potential before and after alkali activation and over a wide
range of pH, from 2–12. Figure 1 presents the results, showing that zeta potential of alkali activated
hydrochar was much more negative in the pH range of interest (pH < 7) than the parent hydrochar.
Under strongly alkali pH, the zeta potential of the parent and the alkali treated material are the same
to within the limits of experimental uncertainty, which is consistent with expectations given that the
alkali treatment is simply immersion in an alkali solution with pH > 9.
Zeta potential measurements support electrostatic interaction as the primary basis of cation
sorption to the hydrochar, providing a valuable clue for rational design. For further comparison,
we measured the zeta potential of one of the aforementioned activated carbons (Norit® SX-1) and
include these data in Figure 1. The zeta potential of the activated carbon was much less negative than
glucose hydrochar, even before alkali treatment. Again, this points to a qualitatively different sorption
mechanism for activated carbon compared to hydrochar, with cation sorption to activated carbon likely
occurring due to cation-π interactions, which appear to be less important than electrostatic interactions
for cation binding to hydrochar [86].
Rational sorbent design requires understanding the molecular binding sites. Figure 1 clearly
implicates the importance of groups that ionize on alkali treatment, which naturally suggests carboxylic
acids, acid anhydrides, and strongly acidic aromatic alcohols, such as phenol [16,24,73]. The fact that
strong bases were more activating than weak ones (Table 1) seems consistent with de-protonation of
59
Energies 2020, 13, 4203
weak acids but does not provide sufficient molecular detail for sorbent design. Accordingly, glucose
hydrochar was analyzed using FT-IR for identification of ionizable OFGs. Figure 2 provides FT-IR
spectra divided into the fingerprint region, 1000–2000 cm−1 (Figure 2a) and the C−H and O−H
stretching region, 2400–4000 cm−1 (Figure 2b). Prior to alkali treatment, glucose hydrochar exhibits
bands attributable to carbonyl (1720 cm−1 ) and hydroxyl (3200–3400 cm−1 ) groups associated with
carboxylic acids. Other features at 2900, 1600, and 1200 cm−1 are attributable to C−H stretches,
arenes/furans breathing modes, and C−O stretches, respectively [65] These spectroscopic attributions
are broadly consistent with the structural models previously inferred from NMR [65], Raman [61],
and NEXAFS [62–64,66]. Interestingly, the C−H and O−H stretches are sufficiently differentiated to
suggest that carboxylic acid exists in its free, rather than dimerized, state [87].
(a) (b)
Figure 2. FT-IR spectra of glucose hydrochar as synthesized and after activation by various bases.
(a) the fingerpint region (1000–2000 cm−1 ) (b) the O−H and C−H stretch region (2500–4000 cm−1 ).
After alkali treatment, the intensities of the hydroxyl band at 3200–3400 cm−1 and carbonyl band
at 1720 cm−1 become much less intense. Simultaneously with these changes, the intensity of the band
at 1600 cm−1 increases and the feature broadens noticeably. The C−O stretch present at approximately
1200 cm−1 becomes less intense and broader after alkali treatment. The effects are more noticeable for
treatment with the hydroxides than the carbonates, consistent with their relative basicities and with
the observed effects on sorption capacity, noted in Table 1.
All of the aforementioned changes observed in the FT-IR spectra of glucose hydrochar after alkali
treatment are attributable to deprotonation of carboxylic acid groups to form carboxylates [37,88].
Specifically, deprotonation involves a shift of the main carbonyl band from approximately 1700 to
about 1600 cm−1 ; [84] a reduction of intensity of the C−O stretch at 1200 cm−1 ; and a reduction of the
intensity of the O−H stretch at 3200–3400 cm−1 . The last of these is consistent with partial removal of
the H atoms involved with O−H stretches, as expected for de-protonation. The carboxylate feature at
1600 cm−1 overlaps with the furan/arene breathing mode that is characteristic of hydrochar [61,63,89].
60
Energies 2020, 13, 4203
The fact that alkali treated hydrochar still exhibits an O−H stretching band is consistent either with
incomplete de-protonation of acid groups or with the presence of multiple forms of O−H in the
structure (i.e., alcohol groups that are not sufficiently acidic to be deprotonated).
Since the carboxylic acid groups present in glucose hydrochar appear to be primarily in their
protonated forms (Figure 1) and since the pH of the HTC reaction mixture is about 3 [90], the pKa of
these acid groups must be greater than approximately 3–otherwise, they would be present in hydrochar
in their deprotonated forms. Alkali treatment then deprotonates these groups, resulting in formation
of the alkali carboxylate. Because hydrochar is a complex material and because localized induction
and steric effects can influence pKa [91], carboxylic acids present in hydrochar likely possess a range of
pKa’s. In fact, this assertion is supported from the broad zeta potential curve observed for glucose
hydrochar and shown in Figure 1. Treatment with carbonates may therefore deprotonate only the
strongest carboxylic acids present in hydrochar, while treatment with hydroxides deprotonates both
strong and weak carboxylic acids.
We considered the possibility of alternative ionizable groups, aside from carboxylic acid.
Treatment with hydroxide would partially deprotonate any strongly acidic alcohol groups (e.g., phenols)
present in the hydrochar structure; however, the FT-IR spectra show no direct evidence to support
the formation of phenolate ions, nor do reported structural models suggest the presence of
phenol in hydrochar [24,61,63]. Accordingly, metal-carboxylate binding appears to be the primary
cation adsorption mechanism underlying glucose hydrochar sorption, providing a clear target for
molecular simulation.
Hydrochar is thought to be composed of furan/arene polymers connected by alkyl spacers.
Mild alkali treatment is insufficient to break or form covalent bonds present in this structure [82],
which is consistent with the negligible change in surface area associated with alkali treatment
(see Table 1). That stated, Mihajlovic et al. [92] proposed that hydrolytic degradation of OFGs can
sometimes occur during alkali treatment, and re-arrangement of the hydrochar structure from its
hypothesized form would complicate attempts to model the binding site. Accordingly, we searched for
evidence of bond breaking in the spectra shown in Figure 2. Inspection of the FT-IR spectra suggests
that important hydrochar features [28,29] at 1020 cm−1 (C−OH alcohol and/alkyl-substituted ethers
groups), 1600 cm−1 (furan/arene breathing modes), and 2900–3000 cm−1 (C−H stretch) are not affected
by alkali treatment, consistent with the main effect of alkali treatment being confined to deprotonation
rather than making and/or breaking of covalent bonds. This observation permits use of published
hydrochar structural models to recreate the local environment of the metal-carboxylate binding site for
DFT simulation.
61
Energies 2020, 13, 4203
of carboxylate groups inferred from sorption capacity measurements presented here. The local
environment experienced by a metal cation during adsorption also includes water solvent molecules.
Here, we recreated the water solvation effect using an implicit cavity model of the appropriate dielectric
constant (taken as 78). Future work can improve the accuracy of our calculations by including explicit
water molecules in the simulation.
D
E
F
We then simulated a series of possible cation-carboxylate structures, starting with the hydrochar
model shown in Figure 3a. The focus of these calculations was to answer the aforementioned questions
that focus on elucidation of trends, rather than quantitative energy estimates. We then simulated cation
binding, shown stoichiometrically in Figure 3b, by replacing either H+ , K+ , or Na+ with the Cu(II)
cation to form the final structure shown in Figure 3c.
Consistent with experimental observations (Table 1), we find that replacing H+ with Cu(II) is
energetically unfavorable, whereas replacing K+ and Na+ is energetically favorable. The simulated
energies are consistent with the observation that glucose hydrochar requires alkali treatment prior to
activation. Moreover, DFT simulations predict that replacing K+ is energetically more favorable than
replacing Na+ by 10.72 kJ mol−1 , which is consistent with the observation that KOH is a more effective
62
Energies 2020, 13, 4203
activating salt than NaOH and K2 CO3 is more effective than Na2 CO3 . That stated, the calculated
energy difference between K+ and Na+ substitution is relatively modest, which is again consistent with
experimental observation. Note that for these reactions the cations in solution may not be properly
modeled by implicit solvation, which is why for instance the replacement of a hydrogen by Cu(II) is so
endothermic. Nonetheless the trends in cation exchange are captured by the DFT calculations.
Figure 3c shows the optimized geometry of a Cu-hydrochar structure. Here, the Cu-carboxylate
bond length is approximately 1.85 Å, slightly longer than that associated with the distance between the
proton and carboxylate group in carboxylic acid. The longer bond is consistent with the size of the
Cu(II) ion compared with the proton [37].
The DFT simulations summarized in Figure 3 explain that alkali treatment removes the proton to
activate the sorption capacity of glucose hydrochar. Physically, the proton is more tightly bonded to
the carboxylate group than the metal cations, owing to the differences in ionic radii and the strong
effect of ion-ion distance on the strength of electrostatic interactions [94]. Similarly, the differences
observed between potassium and sodium can be ascribed to their relative ionic radii.
Interestingly, alkali treatment is not always reported as a necessary step for observation of
hydrochar sorption capacity. This may be due to differences in the reaction mixture pH for different
precursors and/or the presence of alkali salts in many hydrochar starting materials [95,96]. Accordingly,
subtle differences in the reaction mixture and the composition of the precursor may decide whether
or not alkali treatment is required to activate a given hydrochar for metal adsorption. Alternatively,
the alkali step may not be uniformly reported, even when it is required. We recommend more consistent
reporting of alkali treatment and reaction mixture pH in future work in this area.
63
Energies 2020, 13, 4203
Table 2. Adsorption capacity and surface area of custom-synthesized hydrochar and ion exchange resins.
Material Post Treatment Adsorption Capacity (mg Cu g−1 ) BET Surface Area (m2 g−1 )
None <3
Acrylic Na2 CO3 30 ±7
acid-hydrochar K2 CO3 26 ±6 5–10
(AA-hydrochar) NaOH 50 ±4
Hydrochars
KOH 45 ±7
Vinyl sulfonic acid
None <2
(VSA-hydrochar)
<1
NaOH 34 ±5
KOH 51 ±3
Amberlyst® None 128 ±4 53 [97]
Resins
AG® 50W-X4 None 109 ±7 <1
The effect of AA and glucose co-processing to produce hydrochar was consistent with our
expectations, but consistency does not imply confirmation and we considered alternative hypotheses.
Table 2 shows that the surface area of AA-hydrochar was similar to the glucose hydrochar, eliminating
surface area changes as a major difference between these materials. To understand further, we studied
the OFGs of AA-hydrochar using FT-IR. Figure 4 provides the FT-IR spectra obtained for AA-hydrochar
before and after KOH treatment. The FT-IR of glucose hydrochar is included in Figure 4 for direct
comparison to show that the AA-hydrochar spectrum exhibits much more intense bands associated
with carboxylic acids at 1720 (carbonyl) and 1200 cm−1 (C-O stretch) than glucose hydrochar. In fact,
the carboxylic acid bands dominate the AA-hydrochar spectrum and appear as the most prominent
features. The band at 1600 cm−1 , which is characteristic of furans and arenes, appears only as a minor,
though distinct, feature in the AA-hydrochar spectrum. In comparison, the furan/arene band is one
of the most prominent features in the glucose hydrochar spectrum. Similarly, after alkali treatment,
the carbonyl band shifts to approximately 1550 cm−1 and becomes the most prominent feature in the
AA-hydrochar spectrum. Correspondingly, the C-O stretch feature shifts and broadens. Taken together,
these observations clearly indicate that AA-hydrochar has abundant carboxylic acid groups that
deprotonate after alkali treatment.
Figure 4. FT-IR spectra of AA-hydrochar before and after KOH treatment. G-hydrochar is synthesized
entirely of glucose precursor, shown before alkali treatment. AA-hydrochar synthesized from co-feed
of acrylic acid and glucose, shown before and after alkali treatment using KOH.
64
Energies 2020, 13, 4203
The performance of hydrochar sorbents is often compared with activated carbon [25,39]. However,
our findings indicate that hydrochar adsorption is mediated by metal-carboxylate binding interactions
that are more similar to what occurs on an ion exchange resin, rather than activated carbon. Accordingly,
we measured Cu(II) sorption capacity of two commercial ion exchange resins, Amberlyst® -15 and
AG® 50W-X4. Capacity results for these resins are provided in Table 2. Interestingly, these resins far
outperform activated carbon (Table 1) and outperform by about a factor of two the AA-hydrochar.
Moreover, the ion exchange resins did not require alkali activation, unlike the hydrochars. Since the
resins outperform hydrochar, even AA-hydrochar, we sought to understand the differences between
the resins and the hydrochar as part of our rational design approach.
An obvious potential difference between the resins and the hydrochar is surface area. Table 2
provides N2 sorption-based BET surface areas for Amberlyst® -15 and AG® 50W-X4. Interestingly,
Amberlyst® -15 exhibits much greater surface area than any of the hydrochars, which could explain its
superior performance. However, the surface area measured for AG® 50W-X4 was less than any of the
other materials (<1 m2 g−1 , the instrument detection limit), meaning that surface area considerations
alone cannot explain the performance of the resins–at least not the surface area measured by N2
sorption and estimated by BET analysis of the isotherm. In fact, the swelling behavior of ion exchange
resins has been studied carefully in water and other solvents [98–100]; swelling in the presence of water
likely opens the pore structure of AG® 50W-X4 (and possibly the other sorbents), accounting for its
sorption capacity despite negligible N2 sorption surface area. Understanding the effects of hydrochar
swelling on surface area available for cation sorption is an area that should be studied in the future.
The binding site in both Amberlyst® -15 and AG® 50W-X4 is a sulfonate group [101], whereas the
findings presented here indicate that carboxylate groups are mainly responsible for binding in glucose
hydrochar and especially AA-hydrochar. The sulfonic acid group is at least 1000× stronger than the
carboxylic acid group, meaning that this difference could explain sorption behavior and the need for
alkali activation. Accordingly, we modified the acrylic acid synthesis procedure for incorporation
of a sulfonate group into the hydrochar structure by co-processing glucose and vinyl sulfonic acid.
Like acrylic acid, vinyl sulfonic acid possesses a polymerizable double bond that can be incorporated
in the hydrochar-alkylated backbone. Unlike acrylic acid, though, vinyl sulfonic acid can introduce
a sulfonate group into the hydrochar instead of the carboxylic acid introduced by acrylic acid.
Accordingly, we term this new char vinyl sulfonic acid-hydrochar, or VSA-hydrochar.
Table 2 provides the Cu(II) cation sorption capacity of VSA-hydrochar. Interestingly, despite the
strength of the vinyl sulfonic acid precursor (pKa < 1 compared with 4.35 for acrylic acid) [98,102],
we observed negligible Cu(II) sorption capacity for VSA-hydrochar before treatment with alkali.
After treatment with KOH, the Cu(II) capacity of VSA-hydrochar increased substantially to
51 ± 3 mg g−1 . Interestingly, NaOH was much less effective at increasing Cu(II) sorption capacity than
KOH, consistent with the aforementioned trend observed and simulated for carboxylate binding.
As before, the measured BET surface area of VSA-hydrochar was in the same range as the
other hydrochars and <10 m2 g−1 . Similarly, Figure 5 shows the FT-IR spectrum of VSA-hydrochar.
Unlike carboxylic acid and carboxylate groups that have intense and well-differentiated vibrational
bands, sulfonic acid and sulfonate give rise to weak and broad bands that are not easily differentiated
from other features [103]. That stated, the FT-IR VSA-hydrochar spectrum contains bands in the range
expected for sulfonic acid (1100–1300 cm−1 ). The carboxylate/carboxylic acid bands are less intense in
VSA-hydrochar than glucose hydrochar, indicating substitution of the weak acid in AA-hydrochar for
the strong acid in VSA-hydrochar.
Cu(II)-sulfonate structures were simulated using DFT methods, similar to those previously
presented for carboxylate binding. Figure 6a shows the sulfonic acid- hydrochar geometry,
which consisted of two furan groups bonded to a sulfonic acid group. As before, binding was
simulated as an exchange of Cu(II) for H+ , K+ , and Na+ . Despite the strength of the sulfonic acid,
DFT calculations found that replacing H+ with Cu(II) was thermodynamically unfavorable, consistent
with the need to activated VSA-hydrochar with alkali. Figure 6b summarizes this result. Similarly,
65
Energies 2020, 13, 4203
the distance of the Cu−O bonded to sulfonate (shown in Figure 6c) is 1.95 Å, somewhat greater than the
Cu−O bond in carboxylate hydrochar (1.85 Å). As before, the cations in solution may not be properly
modeled by implicit solvation, which is why some energies may be so large, despite DFT identifying
the trends in cation exchange.
Figure 5. FT-IR spectrum of VSA-hydrochar, before and after KOH treatment. Vertical lines mark the
carboxylic acid and carboxylate vibration bands. The region where sulfonic acid vibrations appear is
indicated. The spectrum of glucose hydrochar is reproduced from Figure 4 as a point of reference.
Surface area, FT-IR, and DFT simulations provide further evidence of cation-sulfonate binding
in the VSA-hydrochar, but do not explain why the performance of neither VSA-hydrochar nor
AA-hydrochar can match the commercial ion exchange resins. As a final hypothesis, we quantified
the density of surface acids present on the various sorbents, with the expectation that differences
in the density of surface acids might explain observed differences in sorption capacity. For these
experiments, hydrochars were first treated with strong acid (HCl) to protonate fully all available acid
groups. Then, the acid group density was measured of the protonated sorbent using Boehm titration
methods [29,71,72].
Table 3 summarizes the carboxylic acid site density measurements. As expected, the density
of acid functional groups on the glucose hydrochar is much greater than on the activated carbons
considered here, consistent with the different adsorption mechanisms for the two materials (primarily
electrostatic vs. primarily π-cation). The ion exchange resins have much greater acid concentrations
than any of the other sorbents, consistent with their superior performance and indicating that the AA-
and VSA-hydrochars function as designed, albeit with fewer acid binding groups than are available on
the ion exchange resins tested here. Nonetheless, the Cu(II) adsorption performance of the designer
hydrochars is comparable to the ion exchange resins (to within a factor of two) and superior to activated
carbon, meaning that strategies to increase acid functional group density can be effective for synthesis
of task-specific hydrochar sorbents.
Table 3 provides qualitative evidence of the importance of acid group density on sorption
performance and permits analysis of a critical parameter: the binding stoichiometry of the metal-acid
complex formed during adsorption. Binding stoichiometry is important for quantifying sorbent
performance since the ideal absorbent will possess high density of binding sites and utilize them as
efficiently as possible. Simultaneously achieving high binding site density and binding site utilization
may not be possible, since densely spaced binding sites may promote bidentate binding instead of
66
Energies 2020, 13, 4203
monodentate binding, which is less efficient binding site utilization. We analyzed the sorption and
acid site density data to evaluate these effects in hydrochar, activated carbon, and ion exchange resins.
D
E
F
Figure 6. Sulfonate-containing hydrochar structures optimized using DFT. The structure in (a) is
the sulfonated hydrochar molecule. (b) Depicts the reactants which are the interactions between the
sulfonate group and either hydrogen, potassium, or sodium, respectively. Structure (c) shows binding
of Cu(II) to the sulfonated hydrochar molecule. Adsorption energies are provided as shown. Legend:
carbon; hydrogen; oxygen; sulfur; potassium; sodium; copper.
67
Energies 2020, 13, 4203
To use Table 3 data to understand stoichiometry, we plotted Cu(II) sorption capacity as a function of
measured acid group density, converting both to molar quantities, as shown in Figure 7. For comparison,
lines of constant ion-binding site stoichiometry (two Cu ions per acid, 1:1, and 1:2) are shown. Data for
the activated carbons cluster around the origin and fall entirely off the stoichiometric trend lines,
as expected given that the sorption mechanism to activated carbons is likely cation-π interactions and
is, therefore, independent of acid site density. In contrast, sorption for the hydrochars falls between the
1:1 and 1:2 stoichiometry lines, indicating that—on average—each acid group binds approximately
0.75 Cu ions. This again is further quantitative evidence of the importance of electrostatic interactions
for binding to hydrochar.
Figure 7. Measured Cu(II) adsorption capacity plotted as a function of measured carboxylate group
density. Lines of constant cation-anion stoichiometry are shown as dotted lines at 2 Cu(II) ions per
binding site (2:1), 1 Cu(II) ions per binding site, and 1 Cu(II) ion for every two binding sites. Legend:
depicts activated carbon, ion exchange resins, hydrochar.
The stoichiometry inferred from Figure 7 shows that the designer hydrochars outperform
Amberlyst® -15 on a per acid site basis. This is an important finding since increasing binding site
utilization efficiency is an effective means of increasing sorption capacity, along with increasing
the density of binding sites themselves. Interestingly, AG® 50W-X4 far exceeds all other sorbents
on effectiveness per acid site, with nearly two Cu ions associated with every acid site (Figure 6).
The difference between AG® 50W-X4 and Amberlyst® -15 is noteworthy as both sorbents are described
in the literature as polymerized styrene backbones with periodic sulfonic acid group substitution [101].
The difference in their performance must be due either to (1) the ability of the sorbent to hold charge,
which could be saturated for Amberlyst® -15 limiting its sorption capacity, (2) differences in acid site
accessibility in the swollen resins and hydrochars, or (3) differences in the spatial proximity of the
acid binding groups in the different sorbents. The performance of AG® 50W-X4 suggests further
engineering of the hydrochar structure to optimize sorption capacity.
To understand the origins of stoichiometry between Cu(II) ions and carboxylate or sulfonate
groups, we performed simulations to compare monodentate with bidentate binding of Cu(II) to
carboxylate and sulfonate groups. To make the calculation accessible using DFT, we simplified the
structure previously used in Figures 3 and 6 to remove the furan groups. Figure 8a shows the optimized
geometry for the monodentate binding structures, and Figure 8b shows the optimized geometry for
the bidentate binding structures. As expected, bidentate binding is much more energetically favorable
than binding to a single acid functional group. For the sulfonate site, bidentate binding is more
68
Energies 2020, 13, 4203
stable by 201 kJ mol−1 , and for the carboxylate site, bidentate binding is more stable by 126 kJ mol−1 .
These values indicate a clear thermodynamic preference for bidentate binding. As measured by the
Cu−O distance, the Cu(II) ion is roughly equidistant between the two sulfonate groups; as a result,
the Cu−O distance in bidentate binding complex is greater than found in the geometry optimized
for single Cu-acid stoichiometry (shown previously in Figure 6). This clearly shows that Cu(II) (and
presumably other double charged cations) will prefer bidentate binding, when such an option is
available. Since bidentate binding is a less efficient use of sites than monodentate binding, rational
design of hydrochars should attempt to achieve uniform acid spacing to minimize acid-acid interaction
and the ability of cations to bind simultaneously to multiple acid sites.
Figure 8. Optimized geometries of monodentate (a) and bidentate (b) binding of Cu(II) to hydrochar
carboxylate and sulfonate groups simulated using DFT. Oxygen-Cu(II) distances are shown for reference.
Legend: carbon; hydrogen; oxygen; sulfur; copper.
When functional group precursors with polymerizable double bonds are co-fed to the HTC
reactor with glucose, the functional group bearing molecules will polymerize primarily with each
other, rather than with groups present in the hydrochar. Because they are formed by co-feeding
glucose and vinyl groups, VSA-hydrochar is not engineered to achieve the desired spacing, which may
explain why it falls short of the desired 1:1 Cu-binding site stoichiometry. More uniform spacing of
the binding groups has potential to improve binding site utilization by forcing binding to occur via
the preferred monodentate arrangement rather than via the thermodynamically preferred bidentate
geometry. Furthermore, utilization of vinyl sulfonic acid as a source of binding groups detracts
from the renewable and green characteristics of hydrochar. Accordingly, future work in this area
should seek to utilize feeds that are naturally abundant in anionic binding sites and/or functional
groups that are converted into anionic binding sites during HTC. Questions of binding site access
and cooperative effects should be addressed for hydrochars synthesized from renewable or waste
resources, using similar methods as shown here for rational sorbent design.
As a final analysis, we evaluated cation-π binding to the furan backbone itself in the absence
of acid groups, as a comparison with the arene backbone present in commercial exchange resins.
By providing a secondary stabilizing interaction, optimizing the cation-π binding interaction can
potentially improve the utilization efficiency of the anionic binding sites—a desired goal as explained
previously. In particular, we were interested to understand the effect of locating the cation between
nearby rings as compared with interacting with a single ring individually—in the absence of
69
Energies 2020, 13, 4203
anionic binding groups, such as sulfonate or carboxylate. Figure 9 provides the results of these
calculations; Figure 9a,b depict arene binding and Figure 9c,d depict furan binding, respectively.
In both cases, locating the Cu(II) between two nearby aromatic rings (either furan or arene) is more
stable than interaction with a single aromatic ring. Interestingly, the energy difference is greater for
furan-cation interactions (Figure 9c,d) than arene-cation interactions (Figure 9a,b), by approximately
23 kJ mol−1 . Accordingly, a final design consideration for custom-synthesis of hydrochar sorbents is
inclusion of geometries, which permit formation of furan “pockets” for optimized cation-π interaction.
When combined with electrostatic interactions, cation-π interactions can provide a secondary stabilizing
force to optimize hydrochar sorption capacity.
Figure 9. Optimized geometries calculated for arene, (a,b), and furan, (c,d), cation-π binding, in the
absence of anionic binding groups. Legend: carbon; hydrogen; oxygen; sulfur; copper.
Differences in energy are shown directly in the Figure.
Figures 3 and 6–9 describe a combined experimental and simulation approach for rational design
of hydrochar sorbents to exploit electrostatic interactions between anionic functional groups and metal
cations. Maximizing the effectiveness of each functional group can be achieved by spacing them
uniformly throughout the material, thus emphasizing monodentate binding over bidentate binding.
Presumably, highly effective hydrochar sorbents, such as those reported by Demir-Cakan et al. [56],
who first demonstrated the acrylic acid co-HTC approach, and Xue et al. [16], who activated peanut
hull hydrochar using hydrogen peroxide, already exploit these principles. Likewise, the accuracy
of the computational approach in particular will benefit as hydrochar structure is further resolved,
especially for materials produced from precursors other than glucose. The result of the rational
design approach will be hydrochars with maximized value; thus, making them as competitive as
possible with sorbents obtained from non-renewable resources. Although not within the scope of
this work, computational modeling should be appropriate for guiding selection of conditions for
hydrochar regeneration, for example by using alkali solutions to remove the heavy metal adsorbates.
Applying our adsorption capacity results directly to metals other than Cu(II) is not recommended;
however, the combined experimental and computational approach should be amenable to any metal
cation of interest. Similar analysis can be applied in the future to understand the adsorption of organic
substances to hydrochar, as organic pollutants will exhibit different hydrochar interactions than metal
cations [104].
70
Energies 2020, 13, 4203
4. Conclusions
Glucose hydrochar was studied as a model renewable sorbent for heavy metals, using Cu(II) as a test
case for custom-designing a hydrochar sorbent. Glucose hydrochar required alkali activation to exhibit
Cu(II) sorption capacity; a strong base (hydroxide) was more effective than a weak base (carbonate)
and K+ counter ions were more effective than Na+ for activation. In comparison, activated carbon
sorption was less than observed for activated glucose hydrochar, despite significant differences in
their measured BET surface areas (<10 m2 g−1 compared with >800 m2 g−1 ). Similarly, activated
carbon did not require alkali treatment to promote cation sorption, consistent with entirely different
sorption mechanisms for these two common sorbents. Zeta potential measurements indicated that
Cu(II) sorption to hydrochar was due to electrostatic interactions and FT-IR analysis implicated a key
role for carboxylate groups. DFT simulations provided further information on Cu-hydrochar binding
to the carboxylate site, suggesting beneficial synergy with nearby furan groups.
These results were used as the basis for molecular level design of two hydrochars bearing either
carboxylate or sulfonate groups. The designer hydrochars exhibited approximately 50 m2 g−1 Cu(II)
sorption capacity, consistent with the role of carboxylate and sulfonate groups in cation binding.
Nonetheless, the capacity of the hydrochar sorbents failed to yield the expected increase in sorption
capacity. In accordance with the electrostatic binding mechanism, sorption capacity of two ion
exchange resins, Amberlyst® -15 and AG® 50W-X4, was studied for comparison with hydrochar.
The ion exchange resins outperformed hydrochar on a per mass basis. Interestingly, the hydrochars
bound approximately 0.75 Cu ions per acid site, whereas Amberlyst® -15 bound only 0.5 Cu ions per acid
site. The superior performance of Amberlyst® -15 compared with hydrochar was therefore attributable
entirely to differences in acid site density. AG® 50W-X4, on the other hand, bound nearly 2 Cu ions per
acid site, which accounted for its superior performance compared with hydrochar. DFT simulations
confirmed that bidentate binding is preferred whenever possible, meaning, uniform spacing of the
binding groups will maximize their binding efficiency on a per site basis. Similarly, cooperative
effects from nearby aromatic groups (either furan or arene) can promote cation-π interactions that
improve binding and acid site utilization. The combination of experimental investigation and computer
simulation provide a clear starting point for molecular-level design of hydrochar for use as sorbents,
that can be used in future work.
Author Contributions: Conceptualization, M.T.T. and N.A.D.; Methodology, A.B.B., and G.A.T.; Validation, L.D.
and B.S.; Formal Analysis L.D. and M.P.R.; Investigation J.T.H., L.D., A.A., and M.P.R.; Data Curation, L.D.; Writing
(original draft), Writing (review and editing), N.A.D., M.T.T., A.B.B., A.A, J.T.H. and G.A.T.; Visualization, B.S.,
L.D., M.T.T.; Supervision, M.T.T., N.A.D.; Project Administration, M.T.T., A.B.B., N.A.D.; Funding Acquisition,
M.T.T. All authors have read and agreed to the published version of the manuscript.
Funding: The work was funded by the U.S. National Science Foundation (ENG/CBET 1605916).
Acknowledgments: Rediet Tegegne and Maksim Tyufekchiev supported experimental efforts by synthesizing
and titrating several of the hydrochars.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Christian-Smith, J.; Gleick, P.H.; Cooley, H. Twenty-first century US water policy. In 21st Century US Water
Policy; Oxford University Press: Oxford, UK, 2012.
2. Bellinger, D.C. Lead contamination in Flint—An abject failure to protect public health. N. Engl. J. Med. 2016,
374, 1101–1103. [CrossRef]
3. Retief, F.P.; Cilliers, L. Lead poisoning in ancient Rome. Acta Theol. 2006, 26, 147–164. [CrossRef]
4. Fernandez-Luqueno, F.; López-Valdez, F.; Gamero-Melo, P.; Luna-Suárez, S.; Aguilera-González, E.N.;
Martínez, A.I.; García-Guillermo, M.; Hernández-Martínez, G.; Herrera-Mendoza, R.; Álvarez-Garza, M.A.
Heavy metal pollution in drinking water-a global risk for human health: A review. Afr. J. Environ. Sci. Technol.
2013, 7, 567–584.
71
Energies 2020, 13, 4203
5. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92,
407–418. [CrossRef] [PubMed]
6. Wen, T.; Zhao, Y.; Zhang, T.; Xiong, B.; Hu, H.; Zhang, Q.; Song, S. Effect of anions species on copper
removal from wastewater by using mechanically activated calcium carbonate. Chemosphere 2019, 230, 127–135.
[CrossRef] [PubMed]
7. Charerntanyarak, L. Heavy metals removal by chemical coagulation and precipitation. Water Sci. Technol.
1999, 39, 135–138. [CrossRef]
8. Blöcher, C.; Dorda, J.; Mavrov, V.; Chmiel, H.; Lazaridis, N.; Matis, K. Hybrid flotation—Membrane filtration
process for the removal of heavy metal ions from wastewater. Water Res. 2003, 37, 4018–4026. [CrossRef]
9. Aldrich, C.; Feng, D. Removal of heavy metals from wastewater effluents by biosorptive flotation. Miner. Eng.
2000, 13, 1129–1138. [CrossRef]
10. Al-Rashdi, B.; Johnson, D.; Hilal, N. Removal of heavy metal ions by nanofiltration. Desalination 2013, 315,
2–17. [CrossRef]
11. Mukherjee, R.; Bhunia, P.; De, S. Impact of graphene oxide on removal of heavy metals using mixed matrix
membrane. Chem. Eng. J. 2016, 292, 284–297. [CrossRef]
12. Oehmen, A.; Viegas, R.; Velizarov, S.; Reis, M.A.; Crespo, J.G. Removal of heavy metals from drinking water
supplies through the ion exchange membrane bioreactor. Desalination 2006, 199, 405–407. [CrossRef]
13. Wang, A.; Zhou, K.; Zhang, X.; Zhou, D.; Peng, C.; Chen, W. Arsenic removal from highly-acidic wastewater
with high arsenic content by copper-chloride synergistic reduction. Chemosphere 2020, 238, 124675. [CrossRef]
[PubMed]
14. Tran, T.-K.; Chiu, K.-F.; Lin, C.-Y.; Leu, H.-J. Electrochemical treatment of wastewater: Selectivity of the heavy
metals removal process. Int. J. Hydrogen Energy 2017, 42, 27741–27748. [CrossRef]
15. Hunsom, M.; Pruksathorn, K.; Damronglerd, S.; Vergnes, H.; Duverneuil, P. Electrochemical treatment of
heavy metals (Cu2+, Cr6+, Ni2+) from industrial effluent and modeling of copper reduction. Water Res.
2005, 39, 610–616. [CrossRef] [PubMed]
16. Xue, Y.; Gao, B.; Yao, Y.; Inyang, M.; Zhang, M.; Zimmerman, A.R.; Ro, K.S. Hydrogen peroxide modification
enhances the ability of biochar (hydrochar) produced from hydrothermal carbonization of peanut hull to
remove aqueous heavy metals: Batch and column tests. Chem. Eng. J. 2012, 200, 673–680. [CrossRef]
17. Ihsanullah, D.; Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A.
Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption
applications. Sep. Purif. Technol. 2016, 157, 141. [CrossRef]
18. Bakkaloglu, I.; Butter, T.; Evison, L.; Holland, F.; Hancockt, I. Screening of various types biomass for removal
and recovery of heavy metals (Zn, Cu, Ni) by biosorption, sedimentation and desorption. Water Sci. Technol.
1998, 38, 269–277. [CrossRef]
19. Rengaraj, S.; Yeon, J.-W.; Kim, Y.; Jung, Y.; Ha, Y.-K.; Kim, W.-H. Adsorption characteristics of Cu (II) onto ion
exchange resins 252H and 1500H: Kinetics, isotherms and error analysis. J. Hazard. Mater 2007, 143, 469–477.
[CrossRef]
20. Lin, S.H.; Lai, S.L.; Leu, H.G. Removal of heavy metals from aqueous solution by chelating resin in a
multistage adsorption process. J. Hazard. Mater 2000, 76, 139–153. [CrossRef]
21. Pourhashem, G.; Adler, P.R.; Spatari, S. Time effects of climate change mitigation strategies for second
generation biofuels and co-products with temporary carbon storage. J. Clean. Prod. 2016, 112 Pt 4, 2642.
[CrossRef]
22. Bulut, Y. Removal of heavy metals from aqueous solution by sawdust adsorption. J. Environ. Sci. 2007, 19,
160–166. [CrossRef]
23. Javaid, A.; Bajwa, R.; Shafique, U.; Anwar, J. Removal of heavy metals by adsorption on Pleurotus ostreatus.
Biomass Bioenergy 2011, 35, 1675–1682. [CrossRef]
24. Kılıc, M.; Kırbıyık, C.; Çepelioğullar, Ö.; Pütün, A.E. Adsorption of heavy metal ions from aqueous solutions
by bio-char, a by-product of pyrolysis. Appl. Surf. Sci. 2013, 283, 856–862. [CrossRef]
25. Inyang, M.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.R.; Pullammanappallil, P.; Cao, X. Removal of heavy
metals from aqueous solution by biochars derived from anaerobically digested biomass. Bioresour. Technol.
2012, 110, 50–56. [CrossRef]
26. Tripathi, M.; Sahu, J.N.; Ganesan, P. Effect of process parameters on production of biochar from biomass
waste through pyrolysis: A review. Renew. Sustain. Energy Rev. 2016, 55, 467–481. [CrossRef]
72
Energies 2020, 13, 4203
27. Wiedner, K.; Rumpel, C.; Steiner, C.; Pozzi, A.; Maas, R.; Glaser, B. Chemical evaluation of chars produced
by thermochemical conversion (gasification, pyrolysis and hydrothermal carbonization) of agro-industrial
biomass on a commercial scale. Biomass Bioenergy 2013, 59, 264–278. [CrossRef]
28. Sun, Y.; Gao, B.; Yao, Y.; Fang, J.; Zhang, M.; Zhou, Y.; Chen, H.; Yang, L. Effects of feedstock type, production
method, and pyrolysis temperature on biochar and hydrochar properties. Chem. Eng. J. 2014, 240, 574–578.
[CrossRef]
29. Sevilla, M.; Ferrero, G.A.; Fuertes, A.B. Beyond KOH activation for the synthesis of superactivated carbons
from hydrochar. Carbon 2017, 114, 50–58. [CrossRef]
30. Wahi, R.; Zuhaidi, N.; Yusof, Y.; Jamel, J.; Kanakaraju, D.; Ngaini, Z. Chemically treated microwave-derived
biochar: An overview. Biomass Bioenergy 2017, 107, 411–421. [CrossRef]
31. Verma, D.; Fortunati, E.; Jain, S.; Zhang, X. Biomass, Biopolymer-Based Materials, and Bioenergy: Construction,
Biomedical, and Other Industrial Applications; Woodhead Publishing: Cambridge, UK, 2019.
32. Zhang, J.; Zhang, X. 15-The thermochemical conversion of biomass into biofuels. In Biomass, Biopolymer-Based
Materials, and Bioenergy; Verma, D., Fortunati, E., Jain, S., Zhang, X., Eds.; Woodhead Publishing: Cambridge,
UK, 2019; pp. 327–368.
33. Regmi, P.; Moscoso, J.L.G.; Kumar, S.; Cao, X.; Mao, J.; Schafran, G. Removal of copper and cadmium from
aqueous solution using switchgrass biochar produced via hydrothermal carbonization process. J. Environ.
Manag. 2012, 109, 61–69. [CrossRef]
34. Marfíl, A.P.; Ocampo-Pérez, R.; Collins-Martínez, V.H.; Flores-Vélez, L.M.; Gonzalez-Garcia, R.;
Medellín-Castillo, N.A.; Labrada-Delgado, G.J. Synthesis and characterization of hydrochar from industrial
Capsicum annuum seeds and its application for the adsorptive removal of methylene blue from water.
Environ. Res. 2020, 184, 109334. [CrossRef] [PubMed]
35. Funke, A.; Ziegler, F. Hydrothermal carbonization of biomass: A summary and discussion of chemical
mechanisms for process engineering. Biofuels Bioprod. Biorefining 2010, 4, 160–177. [CrossRef]
36. Titirici, M.-M. Sustainable Carbon Materials from Hydrothermal Processes; John Wiley & Sons: Hoboken, NJ,
USA, 2013.
37. Sun, K.; Tang, J.; Gong, Y.; Zhang, H. Characterization of potassium hydroxide (KOH) modified hydrochars
from different feedstocks for enhanced removal of heavy metals from water. Environ. Sci. Pollut. Res. 2015,
22, 16640–16651. [CrossRef] [PubMed]
38. Poo, K.-M.; Son, E.-B.; Chang, J.-S.; Ren, X.; Choi, Y.-J.; Chae, K.-J. Biochars derived from wasted marine
macro-algae (Saccharina japonica and Sargassum fusiforme) and their potential for heavy metal removal in
aqueous solution. J. Environ. Manag. 2018, 206, 364–372. [CrossRef] [PubMed]
39. Karnib, M.; Kabbani, A.; Holail, H.; Olama, Z. Heavy Metals Removal Using Activated Carbon, Silica and Silica
Activated Carbon Composite; Elsevier Ltd.: Amsterdam, The Netherlands, 2014; Volume 50, pp. 113–120.
40. Oliveira, I.; Blöhse, D.; Ramke, H.-G. Hydrothermal carbonization of agricultural residues. Bioresour. Technol.
2013, 142, 138–146. [CrossRef]
41. Kruse, A.; Funke, A.; Titirici, M.-M. Hydrothermal conversion of biomass to fuels and energetic materials.
Curr. Opin. Chem. Biol. 2013, 17, 515–521. [CrossRef]
42. Saqib, N.U.; Sharma, H.B.; Baroutian, S.; Dubey, B.; Sarmah, A.K. Valorisation of food waste via hydrothermal
carbonisation and techno-economic feasibility assessment. Sci. Total Environ. 2019, 690, 261–276. [CrossRef]
43. Yu, X.; Liu, S.; Lin, G.; Yang, Y.; Zhang, S.; Zhao, H.; Zheng, C.; Gao, X. Promotion effect of KOH surface
etching on sucrose-based hydrochar for acetone adsorption. Appl. Surf. Sci. 2019, 496, 143617. [CrossRef]
44. Jain, A.; Balasubramanian, R.; Srinivasan, M.P. Tuning hydrochar properties for enhanced mesopore
development in activated carbon by hydrothermal carbonization. Microporous Mesoporous Mater. 2015, 203,
178–185. [CrossRef]
45. Sevilla, M.; Fuertes, A.B. A green approach to high-performance supercapacitor electrodes: The chemical
activation of hydrochar with potassium bicarbonate. Chem. Sus. Chem. 2016, 9, 1880–1888. [CrossRef]
46. Guo, S.; Dong, X.; Liu, K.; Yu, H.; Zhu, C. Chemical, energetic, and structural characteristics of hydrothermal
carbonization solid products for lawn grass. Bio. Resour. 2015, 10, 4613–4625. [CrossRef]
47. Guo, S.; Dong, X.; Wu, T.; Shi, F.; Zhu, C. Characteristic evolution of hydrochar from hydrothermal
carbonization of corn stalk. J. Anal. Appl. Pyrolysis 2015, 116, 1–9. [CrossRef]
48. Guo, S.; Dong, X.; Wu, T.; Zhu, C. Influence of reaction conditions and feedstock on hydrochar properties.
Energy Convers. Manag. 2016, 123, 95–103. [CrossRef]
73
Energies 2020, 13, 4203
49. Baccile, N.; Falco, C.; Titirici, M.-M. Characterization of biomass and its derived char using 13 C-solid state
nuclear magnetic resonance. Green Chem. 2014, 16, 4839–4869. [CrossRef]
50. Aragón-Briceño, C.; Ross, A.B.; Camargo-Valero, M.A. Evaluation and comparison of product yields and
bio-methane potential in sewage digestate following hydrothermal treatment. Appl. Energy 2017, 208,
1357–1369. [CrossRef]
51. Aragón-Briceño, C.I.; Grasham, O.; Ross, A.B.; Dupont, V.; Camargo-Valero, M.A. Hydrothermal carbonization
of sewage digestate at wastewater treatment works: Influence of solid loading on characteristics of hydrochar,
process water and plant energetics. Renew. Energy 2020, 157, 959–973. [CrossRef]
52. Jackowski, M.; Niedzwiecki, L.; Lech, M.; Wnukowski, M.; Arora, A.; Tkaczuk-Serafin, M.; Baranowski, M.;
Krochmalny, K.; Veetil, V.K.; Seruga, P. HTC of Wet Residues of the Brewing Process: Comprehensive
Characterization of Produced Beer, Spent Grain and Valorized Residues. Energies 2020, 13, 2058. [CrossRef]
53. Urbanowska, A.; Kabsch-Korbutowicz, M.; Wnukowski, M.; Seruga, P.; Baranowski, M.; Pawlak-Kruczek, H.;
Serafin-Tkaczuk, M.; Krochmalny, K.; Niedzwiecki, L. Treatment of Liquid By-Products of Hydrothermal
Carbonization (HTC) of Agricultural Digestate Using Membrane Separation. Energies 2020, 13, 262. [CrossRef]
54. Wilk, M.; Magdziarz, A.; Jayaraman, K.; Szymańska-Chargot, M.; Gökalp, I. Hydrothermal carbonization
characteristics of sewage sludge and lignocellulosic biomass. A comparative study. Biomass Bioenergy 2019,
120, 166–175. [CrossRef]
55. Liu, Z.; Quek, A.; Kent Hoekman, S.; Srinivasan, M.P.; Balasubramanian, R. Thermogravimetric investigation
of hydrochar-lignite co-combustion. Bioresour. Technol. 2012, 123, 646–652. [CrossRef]
56. Demir-Cakan, R.; Baccile, N.; Antonietti, M.; Titirici, M.M. Carboxylate-rich carbonaceous materials via
one-step hydrothermal carbonization of glucose in the presence of acrylic acid. Chem. Mater. 2009, 21,
484–490. [CrossRef]
57. Liu, K.; Zhang, S.; Hu, X.; Zhang, K.; Roy, A.; Yu, G. Understanding the adsorption of PFOA on MIL-101
(Cr)-based anionic-exchange metal–organic frameworks: Comparing DFT calculations with aqueous sorption
experiments. Environ. Sci. Technol. 2015, 49, 8657–8665. [CrossRef] [PubMed]
58. Vishnyakov, A.; Ravikovitch, P.I.; Neimark, A.V. Molecular Level Models for CO2 Sorption in Nanopores.
Langmuir 1999, 15, 8736–8742. [CrossRef]
59. Gao, J.; Liu, F.; Ling, P.; Lei, J.; Li, L.; Li, C.; Li, A. High efficient removal of Cu (II) by a chelating resin
from strong acidic solutions: Complex formation and DFT certification. Chem. Eng. J. 2013, 222, 240–247.
[CrossRef]
60. Gupta, N.K.; Sengupta, A.; Boda, A.; Adya, V.C.; Ali, S.M. Oxidation state selective sorption behavior
of plutonium using N, N -dialkylamide functionalized carbon nanotubes: Experimental study and DFT
calculation. Rsc. Adv. 2016, 6, 78692–78701. [CrossRef]
61. Brown, A.B.; McKeogh, B.J.; Tompsett, G.A.; Lewis, R.; Deskins, N.A.; Timko, M.T. Structural analysis
of hydrothermal char and its models by density functional theory simulation of vibrational spectroscopy.
Carbon 2017, 125, 614–629. [CrossRef]
62. Chuntanapum, A.; Matsumura, Y. Formation of tarry material from 5-HMF in subcritical and supercritical
water. Ind. Eng. Chem. Res. 2009, 48, 9837–9846. [CrossRef]
63. Latham, K.G.; Simone, M.I.; Dose, W.M.; Allen, J.A.; Donne, S.W. Synchrotron based NEXAFS study on
nitrogen doped hydrothermal carbon: Insights into surface functionalities and formation mechanisms.
Carbon 2017, 114, 566–578. [CrossRef]
64. Sevilla, M.; Fuertes, A.B. Chemical and Structural Properties of Carbonaceous Products Obtained by
Hydrothermal Carbonization of Saccharides. Chem. A Eur. J. 2009, 15, 4195–4203. [CrossRef]
65. Falco, C.; Perez Caballero, F.; Babonneau, F.; Gervais, C.; Laurent, G.; Titirici, M.-M.; Baccile, N. Hydrothermal
Carbon from Biomass: Structural Differences between Hydrothermal and Pyrolyzed Carbons via 13C Solid
State NMR. Langmuir 2011, 27, 14460–14471. [CrossRef]
66. Baccile, N.; Laurent, G.; Coelho, C.; Babonneau, F.; Zhao, L.; Titirici, M.-M. Structural Insights on
Nitrogen-Containing Hydrothermal Carbon Using Solid-State Magic Angle Spinning 13C and 15N Nuclear
Magnetic Resonance. J. Phys. Chem. C 2011, 115, 8976–8982. [CrossRef]
67. Dieguez-Alonso, A.; Funke, A.; Anca-Couce, A.; Rombolà, A.G.; Ojeda, G.; Bachmann, J.; Behrendt, F.
Towards biochar and hydrochar engineering—Influence of process conditions on surface physical and
chemical properties, thermal stability, nutrient availability, toxicity and wettability. Energies 2018, 11, 496.
[CrossRef]
74
Energies 2020, 13, 4203
68. Naderi, M. Chapter Fourteen-Surface Area: Brunauer–Emmett–Teller (BET). In Progress in Filtration and
Separation; Tarleton, S., Ed.; Academic Press: Oxford, UK, 2015; pp. 585–608.
69. Sheng, K.; Zhang, S.; Liu, J.; Shuang, E.; Jin, C.; Xu, Z.; Zhang, X. Hydrothermal carbonization of cellulose and
xylan into hydrochars and application on glucose isomerization. J. Clean. Prod. 2019, 237, 117831. [CrossRef]
70. Clogston, J.D.; Patri, A.K. Zeta Potential Measurement. In Characterization of Nanoparticles Intended for Drug
Delivery; McNeil, S.E., Ed.; Humana Press: Totowa, NJ, USA, 2011; pp. 63–70.
71. Yu, G.X.; Jin, M.; Sun, J.; Zhou, X.L.; Chen, L.F.; Wang, J.A. Oxidative modifications of rice hull-based carbons
for dibenzothiophene adsorptive removal. Catal. Today 2013, 212, 31–37. [CrossRef]
72. Seredych, M.; Lison, J.; Jans, U.; Bandosz, T.J. Textural and chemical factors affecting adsorption capacity of
activated carbon in highly efficient desulfurization of diesel fuel. Carbon 2009, 47, 2491–2500. [CrossRef]
73. Boehm, H.P. Surface oxides on carbon and their analysis: A critical assessment. Carbon 2002, 40, 145–149.
[CrossRef]
74. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.;
Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09. 2009. Available online: https://gaussian.com/g09citation/
(accessed on 1 May 2020).
75. Schmidt, J.R.; Polik, W.F. WebMO Enterprise, Version 20.0; WebMO LLC: Holland, MI, USA. Available online:
https://www.webmo.net (accessed on 20 May 2020).
76. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev.
A 1988, 38, 3098. [CrossRef]
77. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional
of the electron density. Phys. Rev. B 1988, 37, 785. [CrossRef]
78. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005,
105, 2999–3094. [CrossRef]
79. Kunin, R.; Meitzner, E.F.; Oline, J.A.; Fisher, S.A.; Frisch, N. Characterization of amberlyst 15 macroreticular
sulfonic acid cation exchange resin. Ind. Eng. Chem. Prod. Res. Dev. 1962, 1, 140–144. [CrossRef]
80. Xia, Y.; Liu, H.; Guo, Y.; Liu, Z.; Jiao, W. Immobilization of heavy metals in contaminated soils by modified
hydrochar: Efficiency, risk assessment and potential mechanisms. Sci. Total Environ. 2019, 685, 1201–1208.
[CrossRef] [PubMed]
81. Han, L.; Sun, H.; Ro, K.S.; Sun, K.; Libra, J.A.; Xing, B. Removal of antimony (III) and cadmium (II) from
aqueous solution using animal manure-derived hydrochars and pyrochars. Bioresour. Technol. 2017, 234,
77–85. [CrossRef] [PubMed]
82. Shi, Y.; Zhang, T.; Ren, H.; Kruse, A.; Cui, R. Polyethylene imine modified hydrochar adsorption for chromium
(VI) and nickel (II) removal from aqueous solution. Bioresour. Technol. 2018, 247, 370–379. [CrossRef]
[PubMed]
83. Timko, M.T.; Wang, J.A.; Burgess, J.; Kracke, P.; Gonzalez, L.; Jaye, C.; Fischer, D.A. Roles of surface chemistry
and structural defects of activated carbons in the oxidative desulfurization of benzothiophenes. Fuel 2016,
163, 223–231. [CrossRef]
84. Petrović, J.T.; Stojanović, M.D.; Milojković, J.V.; Petrović, M.S.; Šoštarić, T.D.; Laušević, M.D.; Mihajlović, M.L.
Alkali modified hydrochar of grape pomace as a perspective adsorbent of Pb2+ from aqueous solution.
J. Environ. Manag. 2016, 182, 292–300. [CrossRef]
85. Israelachvili, J.N. Intermolecular and Surface Forces; Academic press: Cambridge, MA, USA, 2011.
86. Tran, H.; You, S.-J.; Chao, H.-P. Insight into adsorption mechanism of cationic dye onto agricultural
residues-derived hydrochars: Negligible role of π-π interaction. Korean J. Chem. Eng. 2017, 34, 1708–1720.
[CrossRef]
87. Doan, V.; Köppe, R.; Kasai, P.H. Dimerization of Carboxylic Acids and Salts: An IR Study in Perfluoropolyether
Media. J. Am. Chem. Soc. 1997, 119, 9810–9815. [CrossRef]
88. Semerciöz, A.S.; Göğüş, F.; Celekli, A.; Bozkurt, H. Development of carbonaceous material from grapefruit
peel with microwave implemented-low temperature hydrothermal carbonization technique for the adsorption
of Cu (II). J. Clean. Prod. 2017, 165, 599–610. [CrossRef]
89. Elaigwu, S.E.; Greenway, G.M. Microwave-assisted and conventional hydrothermal carbonization of
lignocellulosic waste material: Comparison of the chemical and structural properties of the hydrochars.
J. Anal. Appl. Pyrolysis 2016, 118, 1–8. [CrossRef]
75
Energies 2020, 13, 4203
90. Liang, J.; Liu, Y.; Zhang, J. Effect of Solution pH on the Carbon Microsphere Synthesized by Hydrothermal
Carbonization. Procedia Environ. Sci. 2011, 11, 1322–1327. [CrossRef]
91. Feng, Z.; Odelius, K.; Rajarao, G.K.; Hakkarainen, M. Microwave carbonized cellulose for trace pharmaceutical
adsorption. Chem. Eng. J. 2018, 346, 557–566. [CrossRef]
92. Mihajlović, M.; Petrović, J.; Kragović, M.; Stojanović, M.; Milojković, J.; Lopičić, Z.; Koprivica, M. Effect of
KOH activation on hydrochar surface: FT-IR analysis. RAD Assoc. J. 2017, 2, 65–67.
93. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.-M.; Fühner, C.; Bens, O.;
Kern, J. Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry, processes
and applications of wet and dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
94. Xia, Y.; Yang, T.; Zhu, N.; Li, D.; Chen, Z.; Lang, Q.; Liu, Z.; Jiao, W. Enhanced adsorption of Pb(II) onto
modified hydrochar: Modeling and mechanism analysis. Bioresour. Technol. 2019, 288, 121593. [CrossRef]
95. Fornes, F.; Belda, R.M.; Lidón, A. Analysis of two biochars and one hydrochar from different feedstock: Focus
set on environmental, nutritional and horticultural considerations. J. Clean. Prod. 2015, 86, 40–48. [CrossRef]
96. Fang, J.; Gao, B.; Chen, J.; Zimmerman, A.R. Hydrochars derived from plant biomass under various
conditions: Characterization and potential applications and impacts. Chem. Eng. J. 2015, 267, 253–259.
[CrossRef]
97. Dupont, AMBERLYST™ 15DRY Polymeric Catalyst, product Data Sheet. 2019. Available online: https://www.
dupont.com/content/dam/dupont/amer/us/en/water-solutions/public/documents/en/45-D00927-en.pdf
(accessed on 1 December 2019).
98. Nandan, D.; Gupta, A. Solvent sorption isotherms, swelling pressures, and free energies of swelling of
polystyrenesulfonic acid type cation exchanger in water and methanol. J. Phys. Chem. 1977, 81, 1174–1179.
[CrossRef]
99. Davies, C.; Yeoman, G. Swelling equilibria with some cation exchange resins. Trans. Faraday Soc. 1953, 49,
968–974. [CrossRef]
100. Marcus, Y.; Naveh, J. Anion exchange of metal complexes. XVII. Selective swelling of the exchanger in mixed
aqueous-organic solvents. J. Phys. Chem. 1969, 73, 591–596. [CrossRef]
101. Howery, D.G.; Shore, L.; Kohn, B.H. Proton magnetic resonance studies of the structure of water in Dowex
50 W. J. Phys. Chem. 1972, 76, 578–581. [CrossRef]
102. Mao, Y.; Fung, B.M. A Study of the Adsorption of Acrylic Acid and Maleic Acid from Aqueous Solutions
onto Alumina. J. Colloid Interface Sci. 1997, 191, 216–221. [CrossRef] [PubMed]
103. Sun, B.; Zhao, Y.; Wu, J.-G.; Yang, Q.-C.; Guang-Xian, X. Crystal structure and FT-IR study of cesium
4-methylbenzenesulfonate. J. Mol. Struct. 1998, 471, 63–66. [CrossRef]
104. Zhang, X.; Gao, B.; Fang, J.; Zou, W.; Dong, L.; Cao, C.; Zhang, J.; Li, Y.; Wang, H. Chemically activated
hydrochar as an effective adsorbent for volatile organic compounds (VOCs). Chemosphere 2019, 218, 680–686.
[CrossRef] [PubMed]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
76
energies
Article
Sewage Sludge Valorization via Hydrothermal
Carbonization: Optimizing Dewaterability and
Phosphorus Release
Taina Lühmann and Benjamin Wirth *
Biorefineries Department, Deutsches Biomasseforschungszentrum Gemeinnützige GmbH, Torgauer Straße 116,
04347 Leipzig, Germany; taina.luehmann@ru.nl
* Correspondence: benjamin.wirth@dbfz.de; Tel.: +49-341-2434-449
Abstract: As the use of sewage sludge as a fertilizer in agriculture is increasingly restricted in the
European Union, other ways to utilize this waste stream need to be developed. Sewage sludge
is an ideal input material for the process of hydrothermal carbonization, as it can convert wet
biomass into a solid energy carrier with increased mechanical dewaterability. Digested sewage sludge
was hydrothermally carbonized at 160–200 ◦ C for 30–60 min with initial pH levels of 1.93–8.08 to
determine optimal reaction conditions for enhanced dewaterability and phosphorus release into
the liquid phase. Design of experiments was used to develop response surface models, which can
be applied to optimize the process conditions. For optimal dewaterability and phosphorus release,
low initial pH values (pH 1.93) and mild temperatures around 170 ◦ C are favorable. Because holding
time had no statistically relevant effect, a dependency of reaction time was investigated. Though it
did not yield substantially different results, it could be included in investigations of short reaction
times prospectively. Low reaction temperatures and short holding times are desirable considering
economic reasons for scale-up, while the high acid consumption necessary to achieve these results
is unfavorable.
Keywords: dry matter; moisture; dewatering; sewage sludge; hydrothermal treatment; phosphorus;
RSM; DoE; optimization; regression model
1. Introduction
Sewage sludge is widely used in agriculture because high contents of phosphorus, nitrogen,
and potassium, as well as organic carbon make it a favorable organic fertilizer [1,2]. Because of
concerning substances, such as heavy metals, organic residues, microplastics, and various pathogens,
stronger regulations regarding the use as fertilizer have been passed. This has been combined with a
requirement to recover phosphorus from sewage sludge in countries such as Germany [3,4], so that
new ways for valorization are sought after. After landfilling became restricted in the European
Union (EU) in 1999, incineration has become the preferred solution in the EU-15 countries [5]. During
incineration, sewage sludge can be mono-combusted in specifically designed furnaces, co-combusted in
existing waste incineration plants or added in small quantities in cement kilns [6]. Thermal dewatering,
an energy intensive process, is the major drawback of incineration [6].
Hydrothermal carbonization (HTC) is a process that can convert sewage sludge into a more
easily mechanically dewaterable substance by breaking up the cell structure, thereby releasing bound
water [7–10]. Essentially, it is a process that converts biomass in hot pressurized water yielding an
upgraded solid biofuel [11,12]. Typical reaction conditions reported in literature are 180 to 250 ◦ C for
1 to 12 h holding time [11] at an elevated pressure due to water saturation pressure and additional
pressure by gaseous compounds formed during the reaction. To measure the dewaterability of
77
Energies 2020, 13, 4417
sewage sludge, different approaches such as filtration and mechanical dewatering were examined.
Danso-Boateng et al. [8] used a filtration cell and calculated the specific cake resistance as a measure
for dewaterability. They found that the reaction temperature was the defining factor for dewaterability,
especially when compared to holding time. When Escala et al. [9] mechanically dewatered hydrochar
(HC) from stabilized sewage sludge that was acidified, they were able to reach dry matter contents
of 52 ± 5.5%, compared to the dry matter content of sewage sludge of 30%. Besides improved
dewaterability, another advantage of applying HTC to sewage sludge is the opportunity for phosphorus
recovery, the most valuable component of wastewater. Distribution of phosphorus in liquid and solid
phase as well as chemical composition has been increasingly studied [13–16]. Shi et al. [14] investigated
the effect of initial pH and reaction temperature on the distribution of phosphorus. They found that
higher temperatures caused phosphorus to be more present in the solid phase while adding large
amounts of acid could shift phosphorus into the liquid phase.
The aim of this study is to investigate the dewaterability and phosphorus release of hydrothermally
treated sewage sludge as minimum required reaction conditions are yet to be determined. A method
to investigate the optimal reaction conditions in a defined reaction space is facilitated by a Design
of Experiments (DoE)/Response Surface Model (RSM) approach. For HTC, this has been already
utilized frequently [13,17–20] and for sewage sludge dewatering after HTC by Danso-Boateng et al. [8].
While they tested the filterability as an indicator for dewatering, the objective of this study was to take
a more practical approach by investigating mechanical dewatering using a screw press at lab-scale.
This experimental set-up was implemented to reflect a possible industrial application. Sewage sludge
was used as received (without any pre-treatment) from a wastewater treatment plant (WWTP) to
identify if mild HTC conditions are suitable for valorization.
Although the process parameters reaction temperature, holding time and initial pH all play a
role on phosphorus release and dewaterability, they have only been investigated individually so far.
To also consider interaction effects, a DoE approach was taken. This allowed for the development of
regression models to predict the above-mentioned parameters as well as typical HTC characteristics.
Reaction conditions that are close to industrial applicability, for which low reaction temperatures and
short holding times are of interest, were investigated.
Table 1. Properties of the tested samples with standard deviations given in parentheses.
The calculated higher heating value (HHV) is slightly lower than the actual measured one as high
ash contents of a sample can severely impact the results of various calculation formulas [21]. However,
78
Energies 2020, 13, 4417
the correlation of ultimate analysis data and HHV used here is reported to fit measured HHVs of HCs
very well and the value for DSS was added as a comparison.
Figure 1. Design space of the face-centered central composite design (FCCD), red dots indicate
triplicates, green dots duplicates, and blue dots single repetitions.
In preparation for the pH adjustment, the amount of acid needed for the targeted pH levels
was determined according to DIN EN 15933 [23]. After homogenizing a mixture of 10 g of sludge
and 100 mL of deionized water on a vibrating plate for 10 min at 130 rpm, concentrated sulfuric
acid (18 M) was added whilst stirring until the desired pH levels were reached and held constant for
30 min. The amount of acid needed for the actual amount of sludge could then be extrapolated from
these results.
For each experiment, around 350 g of DSS were weighed out and the calculated amount of sulfuric
acid (18 M) was added. Of that mixture, 290 g were used for the HTC experiment and 10 g for pH
measurement of the acidified sample. The measured pH was later used for regression modeling instead
of the targeted pH. Except for the addition of acid, the DSS was not further diluted with water but
utilized as received from the WWTP. The experiments were conducted in a 500 mL high-pressure
stirred-tank reactor made of stainless steel (BR-500, Berghof Products + Instruments GmbH, Eningen
unter Achalm, Germany). The stirrer was set to 100 rpm and the reactor was heated up electrically
79
Energies 2020, 13, 4417
with a heating rate of 2 K min−1 . After reaching the desired temperature, the reactor temperature was
kept constant over the holding time. Then, the heater of the reactor was switched off and the reactor
was allowed to cool down to room temperature by the environment.
After the process, the HC-water-slurry was separated with a screw press with an integrated scale.
The screw press was tightened until a pressure of 300 kg (which corresponds to 2.2 bar) was applied
and retightened when the weight dropped by 10 kg. Pressing was stopped when there was no liquid
flow for more than one minute. After the pressing process was completed, the HC was divided into
two homogeneous samples according to DIN EN ISO 14780 [24] by coning and quartering. The HC
was dried and the process water was stored at 4 ◦ C until further use.
2.3. Analytics
In all solid samples (DSS and HC), dry matter content, elemental composition (C, H, N, S) and ash
content was analyzed. Additionally, HHV of DSS was measured and the concentrations of phosphorus
analyzed in DSS as well as process water.
Dry matter content was determined by placing the samples in a drying oven at 105 ◦ C overnight
referring to DIN EN ISO 18134-3 [25]. Elemental carbon, hydrogen, nitrogen, and sulfur (C, H, N, S)
were measured according to DIN EN 15104 [26] with an elemental analyzer (vario MACRO CUBE,
Elementar Analysensysteme GmbH, Langenselbold, Germany). Ash content was determined by
incineration at 550 ◦ C according to DIN EN 14775 [27] and the oxygen content was calculated by
difference. HHV was determined by combustion with oxygen in a bomb calorimeter (Parr Instrument
GmbH, Moline, IL, USA), according to DIN EN ISO 18125 [28]. Phosphorus was determined according
to DIN EN ISO 11885 [29] with a spectrometer (ICP-OES Icap 6300, Thermo Fisher Scientific, Waltham,
MA, USA). Every sample was analyzed three times.
2.4. Calculations
Solid yields based on dry matter (DM) were calculated according to Equation (1) with respect to
the mass of DSS put into the reactor and the mass of HC after the reaction:
mchar × DMchar
Solid yield (%db ) = × 100% (1)
minput × DMinput
For the calculation of carbon yields, Equation (1) was supplemented by the carbon contents of
DSS and HC and calculated as follows:
mchar × DMchar × Cchar
Carbon yield (%db ) = × 100% (2)
minput × DMinput × Cinput
Higher heating value (HHV) was calculated from the elemental analysis according to Channiwala
and Parikh [30]:
HHV MJ kgdb −1 =
(3)
0.3491 × C + 1.1783 × H − 0.1005 × S − 0.1034 × O − 0.0051 × N − 0.0211 × Ash
The calculated HHV from Equation (3) can be then used to determine the energy yield:
To determine the share of phosphorus in the liquid phase (Pliquid ), the concentrations of phosphorus
(cP ) in the DSS and HC are put into relation as follows:
mliquid × cP,liquid
Pliquid (%) = × 100% (5)
minput × DMinput × cP,input
80
Energies 2020, 13, 4417
in which T denotes the reaction temperature, t the holding time after the reaction temperature was
reached and pH the initial pH.
Equation (6) can also be expressed as:
y = Xβ + ε (7)
where y is a vector of the measured responses (e.g., dry matter content), X is a matrix containing
information about the levels of the variables at which the responses were obtained, β is the vector of
regression coefficients, and ε is a vector of random errors. With the aim to find the values of the β’s
that minimize the sum of squares of ε, the least squares estimates are calculated by:
ŷ = X β̂ (9)
and the difference between measured responses (y) and fitted values (ŷ) is given in the vector of
residuals:
e = y − ŷ (10)
Analysis of variance was used to refine the regression model by testing the significance of each
of the terms in Equation (6) by conducting an F-test. The F-value of a term (e.g., temperature) was
calculated by comparing the mean squares of the evaluated term (MSTerm ) and the remaining residuals
(MSResidual ):
MSTerm SSTerm/dfTerm
F0 = = (11)
MSResidual SSResidual/dfResidual
with SS denoting the sum of squares and df the degrees of freedom. F0 was tested against FStat (α, dfTerm ,
dfResidual ) on a significance level of α = 0.1 and if F0 > Fstat , it was concluded that there are significant
effects caused by the evaluated term. After the model was refined to include only statistically relevant
terms, it was further improved by testing to exclude outliers (results from one or more experiments)
from the model. For each term, a t-value was calculated for each run, according to Weisberg [32]:
ê i
ti = (12)
MSResidual,i 1 + xi (X i Xi )−1 xi
The designation i denotes that data from the ith run was excluded during the calculation and x is
essentially a vector containing the first row of X. The t-value was compared with tStat (α/2, n – p − 1),
where n denotes the total number of runs and p the number of parameters in the model. If |ti | < tStat ,
the run was excluded from regression modelling.
81
Energies 2020, 13, 4417
To give the reader an impression of the regression model quality, the metrics predictive and
adjusted R2 are provided. R2 , the coefficient of determination, expresses how much variation around
the mean is explained by the model and is calculated as follows:
SSResidual
R2 = 1 − (13)
SSModel + SSResidual
where SSModel denotes the sum of squares of the model, which is calculated by adding up all SSTerm
of the terms included in the model. If the model captures all variation around the mean, R2 equals
one, and if the model cannot account for any variation, R2 equals zero. R2 is closest to one when all
terms are still in the model and can thus mislead to include terms in the model that do not contribute
a statistically significant effect. An adjusted R2 (R2 adj ) accounts for this and, therefore, decreases as
the number of model terms increases if the additional terms do not improve the model. Additionally,
predictive R2 (R2 pred ) expresses the predictive ability of the model. It takes into account the variation
that arises when excluding one run from the model and using it to predict this value [31].
Regression modeling and data plotting were performed with the software packages Design Expert
12 (Stat-Ease, Inc., Minneapolis, MN, USA) and Origin 2020 (OriginLab Corp., Northampton, MA, USA).
Table 2. Coded regression model coefficients for hydrochar (HC) and process water properties.
The displayed models were further refined by the exclusion of outliers by conducting a t-test
(Equation (12)). Because of the experimental design (FCCD) and the repetitions of corner and central
points, outliers could be excluded without having to limit the order of the model, which led to
regression models with good predictive and adjusted R2 in most cases. The coded regression model
coefficients indicate by how much the response is expected to change when the input parameter is
changed by one unit. By default, the coded units range from −1 for the lowest level and +1 for the
highest level. In this study, the change by one unit relates to a temperature change of 20 ◦ C, holding
82
Energies 2020, 13, 4417
time change of 30 min or a pH change of 3.1. The comparison of the coefficient’s magnitude within a
response allows for an estimation of the relative impact that each input parameter has.
Table 2 shows that reaction temperature and initial pH have a statistically significant (at least
p < 0.05) influence on the output parameters in all models while holding time is only included in two
models. As the absolute values of the coefficients of reaction temperature and initial pH are of similar
magnitude for HC characteristics, it can be assumed that the effect of changing the temperature by
20 ◦ C is similar to a pH change of 3.1. Even though holding time is included for carbon content, its
value is the smallest, which relates to a small influence on the actual model. These trends are different
for the investigated process water properties where pH has 12–21 times higher effects compared to
temperature. In addition, the final pH is equally influenced by reaction temperature as by holding
time. In comparison with other RSM publications, Álvarez-Murillo et al. [19] observed a similar effect
of temperature on the solid yield of olive stone (−5.56 × xT whereby a temperature change of one
unit corresponds to 30 ◦ C). Holding time was included in the model, because their tested maximum
duration of 10 h yielded a statistically relevant effect. Mäkelä et al. [17] encountered a much stronger
influence on the solid yield (daf) treating industrial mixed sludge from a pulp and paper mill (−27 × xT
whereas a temperature change of one unit corresponds to 40 ◦ C). This can be attributed to the higher
temperatures that were investigated (180–260 ◦ C) and the exclusion of the observed high ash contents
for the calculation of the solid yields. Hence, greater differences in yield can be observed. A more in
depth analysis of the investigated output parameters is included in the following sections.
83
Energies 2020, 13, 4417
Table 3. Coded regression model coefficients for hydrochar (HC) and process water properties with
t140 reaction time instead of holding time.
The refined data analysis with reaction time instead of holding time did not yield significantly
different results. This shows that for batch experiments, variation of holding time does not have a
significant influence on most output parameters. It could be considered to be kept constant in future
DoEs, unless parameters such as carbon content, HHV, and final pH are of major interest and holding
or reaction time will be varied in a much wider range. However, holding time is clearly important
when developing continuous processes as it will alter the size of the equipment for a certain targeted
mass flow. This has to be addressed with corresponding equipment in lab-scale to allow for a smaller
difference between holding time and time above a certain reaction temperature.
Figure 2. van-Krevelen-diagram of selected hydrochars (HC) and digested sewage sludge (DSS) at
different process conditions.
84
Energies 2020, 13, 4417
By visualizing the atomic H/C versus O/C ratios of DSS and HCs in Figure 2, insights into the
governing reactions can be gained. Generally, carbonization decreases both H/C and O/C ratios and,
depending on how the ratios change in relation, an assumption about the major governing reaction can
be made. In this case, the atomic ratios of the HCs primarily follow dehydration, which confirms the
finding of Wang and Li [33] that reactions at temperatures below 180 ◦ C are governed by dehydration.
When comparing the HCs that were both produced at 160 ◦ C and pH 7, but at differing holding times
of 30 and 90 min, it becomes apparent that varying the holding time between these time intervals does
not have a noticeable effect. Contrarily, the variation of pH or temperature leads to a more severe
conversion of the feedstock. The pH variation resulted in HCs with lower H/C ratios than caused by
dehydration reactions only.
Considering the lower nitrogen contents of pH-adjusted HCs (Table 4), it seems like the pH
adjustment favored deamination, which is besides decarboxylation the major conversion mechanism
for proteins [22]. This causes the additional removal of hydrogen, resulting in a lower H/C ratio.
When analyzing the influence of temperature variation, HC produced at 200 ◦ C possesses almost the
same atomic ratios as the HC produced at the harshest combined conditions (200 ◦ C, 90 min, pH 2).
It can be seen from the data that the initial calculated HHV based on dry matter of DSS is
slightly higher (13.4 MJ kgdb −1 , Table 1) than the HHV of some of the produced HCs (12.7–15.0 MJ
kgdb −1 ). This is due to the accumulation of ash in HC (increasing ash content from 40.9%db in DSS
to 39.2–53.4%db in HC). Calculating HHV on a dry and ash free basis (daf) results in the expected
increase when comparing the resulting chars (26.0–30.9 MJ kgdaf −1 ) with the initial DSS (22.6 MJ
kgdaf −1 ). The HC characteristics solid, carbon, and energy yield were calculated according to the
described approach (Equations (1), (2), and (4), respectively). They follow the expected trajectory that
more severe reactions result in lower yields. Conducting an analysis of variance for the respective
yields showed for reaction temperature (p < 0.01), initial pH (p < 0.001), and their interaction (p < 0.05)
to have statistically significant effects on the yields. The regression models in terms of actual factors
(T in ◦ C, t in min and pH) are as follows:
In contrast to the majority of studies, holding time was found not to be statistically relevant in
these experiments. This is because others investigated a much longer time span [18,19]. Therefore,
it can be concluded that for solid, carbon, and energy yield, a variation of holding time between 30 and
90 min does not cause statistically different results, and the time could be minimized. Danso-Boateng
et al. [34] developed a regression model for the solid yield of primary sewage sludge (SY = 118.49 −
0.26xT − 0.04xt , actual factors, range of 140–200 ◦ C and 15–240 min) and it compares quite well with
Equation (14) when set at a neutral pH of 7 (SY = 135.62 − 0.30xT , actual factors). Though holding time
is included, the main effect comes from temperature, as found in this study.
The regression model of the solid yield is displayed as a contour plot in Figure 3 as a representative
for the other yields, as they follow the same trend. It shows that mild conditions (low reaction
85
Energies 2020, 13, 4417
temperature and neutral pH) lead to higher solid, carbon, and energy yields. While Mäkelä et al. [18]
found no effect of the addition of acid on the yields when investigating industrial mixed sludge from a
pulp and paper mill, these experiments show a clear effect. This is likely to occur because a wider range
of pH with a more concentrated acid was tested. Unlike Reza et al. [35], who found that mass yield
was more affected by reaction temperature than initial pH when converting wheat straw, the opposite
is indicated here with pH having 1.5 times the effect compared to that of reaction temperature.
Figure 3. Regression model contour plot showing the influence of initial pH and reaction temperature
on the solid yield (%db ).
The dry matter content after mechanical dewatering was used as a measure of the increase in
sludge dewaterability. Analysis of variance has shown that only the influence of reaction temperature
(p < 0.001) and initial pH (p < 0.001) was statistically relevant (see Table 2). Furthermore, their interaction
(p < 0.001) as well as the quadratic term of temperature (p < 0.01) are included in the model due to
their high significance level, resulting in the following equation with actual factors:
Dry matter content (%) = −91.48 + 1.64xT − 7.74xpH + 0.036xT × xpH − 0.0047xT 2 (17)
The resulting response contour, plotted in Figure 4, shows the general trend that low pH and
higher temperatures favor dewaterability. Initial pH has a clear and strong influence, as a more acidic
initial pH yields increasing dry matter contents. A deviation from the expected trend exists at low
pH levels, where an increase in reaction temperature does not result in ever increasing dry matter
contents. This trend is caused by the quadratic term xT 2 in the regression model, which causes the
slight curvature. However, when determining to what extent this term contributes to the model by
comparing the SS of the term xT 2 to the total SS, it becomes clear that it contributes only with 5% to the
model and should not be emphasized.
The regression model of the dry matter content is in agreement with Danso-Boateng et al. [8],
Gao et al. [36], and Wang et al. [37], who also found a strong influence of temperature on dewaterability,
but not of holding time. When the threshold time of 30 min is exceeded, there is no longer any
holding time dependency according to Wang et al. [37]. At the selected temperatures, the typical
HTC reactions already take place sufficiently and another increase in temperature will yield even
faster reactions. Because dehydration and decarboxylation cause a decrease in hydrophilic functional
groups, such as hydroxyl and carboxyl groups [38], the product is more easily dewatered already by a
simple mechanical press. The effect of acidic pH on improved dewaterability can be explained by the
enhancement of dehydration reactions, caused by sulfuric acid [39], which adds to the temperature
effect. Dewaterability is additionally improved by the floc structure disintegration, which releases the
bound water as free water [22]. Escala et al. [9], who conducted a comparable experiment, achieved
similar dry matter contents. They were able to reach a dry matter content of 52 ± 5.5% after applying a
86
Energies 2020, 13, 4417
pressure of 40 bar to HC produced at 205 ◦ C and 24 min. This is higher than the 45.6 ± 2.24% achieved
in this study at comparable reaction conditions (200 ◦ C, 30 min, pH 2 and 7) and is very likely due to a
much lower dewatering pressure.
Figure 4. Regression model contour plot showing the influence of initial pH and reaction temperature
on the dry matter content (%) after mechanical dewatering.
Phosphorus share (%) = 204.32 − 0.42xT − 47.22xpH + 0.059xT × xpH + 2.63xT 2 (19)
A lower initial pH generally leads to a lower pH value after the reaction (Figure 5a). This lower
pH value favors the solution of phosphorus from the solid and the phosphorus concentration in the
process water increases. The influence of the pH value is particularly evident in a range of the initial
pH of 2–5 (Figure 5b).
Figure 5. Regression model contour plots showing (a) the dependence of process water pH and (b) the
dependence of liquid phase phosphorus release (% of total P) on reaction temperature and initial pH.
87
Energies 2020, 13, 4417
At neutral pH, not much phosphorus will be dissolved by varying holding time and reaction
temperature. As others already found, most of the phosphorus is retained in the solid phase [9,13,22].
However, the variation of the initial pH has a major influence on the transfer of phosphorus into the
liquid phase. When Shi et al. [14] reached an initial pH of 0.24 and applied HTC conditions of 170 ◦ C
and 30 min, they were able to transfer 83% of phosphorus into the liquid phase. The results of this
study confirm their findings, as the maximum phosphorus share achieved in these experiments is 78%
at 160 ◦ C and an initial pH of 1.93. An even higher share of 88% was achieved in an experiment at
160 ◦ C and an initial pH of 2.25 but was excluded as an outlier from the model as a result from the
t-test according to Equation (12). It is, thus, possible to transfer phosphorus into the liquid phase to
be potentially recovered by precipitation, e.g., as magnesium ammonium phosphate (struvite) [14],
though this comes at the cost of high acid consumption. Additionally, precipitation of phosphorus in
HTC process water brings additional challenges due to the huge variety of other components found in
the process water.
The future upscaling of this process requires considerations regarding holding time, mixing
behavior and heat transfer, especially because a continuous process is desired. While the results of
this study show that the holding time can be minimized to 30 min for maximum dewaterability and
phosphorus release in a batch process, this time needs to be assessed anew for the application in a
continuous process. During the heating time in a batch process, the conversion of sewage sludge
already begins and, therefore, the holding time of a batch cannot be directly transferred to a continuous
process. More severe reaction conditions (higher reaction temperature and longer holding time) would
be necessary, if energy densification and a further increase in HHV are desired [17].
4. Conclusions
The results of this study provide useful information for the design and optimization of HTC
systems for sewage sludge treatment. It aimed at finding the optimal HTC reaction conditions for a
maximized mechanical dewaterability and phosphorus release in the liquid phase. In the investigated
design space (160–200 ◦ C, 30–90 min, pH 1.93–8.09), low initial pH levels are desirable and holding
time can be minimized as it had no statistically relevant effect. The influence of reaction temperature
is not completely clear. While low temperatures are somewhat more advantageous for phosphorus
release, medium temperatures are favorable for dewaterability.
It was investigated if using reaction time as an input parameter instead of holding time would
cause different results. Though only minor changes were observed, some models were more refined.
However, with regard to future experimental plans using batch reactors, aiming at identifying the
impact of HTC process parameters on several product characteristics, it could be considered to exclude
holding time from the parameter variation. Residence time is of course a main factor in the design of
larger plants, but can only be addressed sufficiently in lab-scale using equipment allowing continuous
operation to minimize actual heating and cooling times.
A numerical optimization to maximize mechanical dewaterability and phosphorus release yielded
achievable dry matter contents of 48.6% and a phosphorus share of 70.3% in the liquid phase at 170 ◦ C
and an initial pH of 1.93. This would be accompanied by an HHV of 13.95 MJ kgdb −1 .
Author Contributions: Conceptualization, T.L. and B.W.; Formal analysis, T.L.; Funding acquisition, B.W.;
Investigation, T.L.; Methodology, T.L. and B.W.; Project administration, T.L.; Supervision, B.W.; Validation, T.L.
and B.W.; Visualization, T.L.; Writing—original draft, T.L.; Writing—review and editing, B.W. All authors have
read and agreed to the published version of the manuscript.
Funding: This research received funding by the Federal Ministry of Food and Agriculture of Germany (BMEL)
and by the Sächsische Aufbaubank (SAB) financed by the European Regional Development Fund (ERDF) under
grant no. 100283030.
Acknowledgments: The authors would like to thank all of the laboratory and technical staff at DBFZ for conducting
the experiments and analyses. Furthermore, the authors would like to thank L. Prokot for her contributions to the
experimental part of this study.
Conflicts of Interest: The authors declare no conflict of interest.
88
Energies 2020, 13, 4417
Appendix A
Table A1. Experimental design (temperature, holding time, and pH) as well as reaction time and
measured pH.
Figure A1. Regression model contour plots showing the dependence of carbon content (%daf ) on
reaction temperature and (a) holding time and (b) reaction time.
89
Energies 2020, 13, 4417
Figure A2. Regression model contour plots showing the dependence of the higher heating value (MJ
kgTS −1 ) on reaction temperature and initial pH based on a model referring to (a) holding time and (b)
reaction time (contours shown at the respective time at the corresponding center point).
Figure A3. Regression model contour plots showing the dependence of the process water pH (-) on
initial pH and (a) holding time and (b) reaction time.
References
1. Sterritt, R.M.; Lester, J.N. The value of sewage sludge to agriculture and effects of the agricultural use of
sludges contaminated with toxic elements: A review. Sci. Total Environ. 1980, 16, 55–90. [CrossRef]
2. Beni, C.; Servadio, P.; Marconi, S.; Neri, U.; Aromolo, R.; Diana, G. Anaerobic Digestate Administration:
Effect on Soil Physical and Mechanical Behavior. Commun. Soil Sci. Plant Anal. 2012, 43, 821–834. [CrossRef]
3. Hudcová, H.; Vymazal, J.; Rozkošný, M. Present restrictions of sewage sludge application in agriculture
within the European Union. Soil Water Res. 2019, 14, 104–120. [CrossRef]
4. BMU. Verordnung über die Verwertung von Klärschlamm, Klärschlammgemisch und
Klärschlammkompost-AbfKlärV. BGBl. I S. 3465. 2017. Available online: https://www.bmu.de/
en/law/sewage-sludge-ordinance/ (accessed on 21 August 2020).
5. Raheem, A.; Sikarwar, V.S.; He, J.; Dastyar, W.; Dionysiou, D.D.; Wang, W.; Zhao, M. Opportunities and
challenges in sustainable treatment and resource reuse of sewage sludge: A review. Chem. Eng. J. 2018, 337,
616–641. [CrossRef]
90
Energies 2020, 13, 4417
6. Werther, J.; Ogada, T. Sewage sludge combustion. Prog. Energy Combust. Sci. 1999, 25, 55–116. [CrossRef]
7. Wang, T.; Zhai, Y.; Zhu, Y.; Li, C.; Zeng, G. A review of the hydrothermal carbonization of biomass waste
for hydrochar formation: Process conditions, fundamentals, and physicochemical properties. Renew. Sust.
Energ. Rev. 2018, 90, 223–247. [CrossRef]
8. Danso-Boateng, E.; Holdich, R.G.; Wheatley, A.D.; Martin, S.J.; Shama, G. Hydrothermal carbonization
of primary sewage sludge and synthetic faeces: Effect of reaction temperature and time on filterability.
Environ. Prog. Sustain. Energy 2015, 34, 1279–1290. [CrossRef]
9. Escala, M.; Zumbühl, T.; Koller, C.; Junge, R.; Krebs, R. Hydrothermal Carbonization as an Energy-Efficient
Alternative to Established Drying Technologies for Sewage Sludge: A Feasibility Study on a Laboratory
Scale. Energy Fuels 2012, 27, 454–460. [CrossRef]
10. Meng, D.; Jiang, Z.; Yoshikawa, K.; Mu, H. The effect of operation parameters on the hydrothermal drying
treatment. Renew. Energy 2012, 42, 90–94. [CrossRef]
11. Libra, J.; Ro, K.S.; Kammann, C.I.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.-M.; Fühner, C.; Bens, O.;
Kern, J.; et al. Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry,
processes and applications of wet and dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
12. Reza, M.T.; Andert, J.; Wirth, B.; Busch, D.; Pielert, J.; Lynam, J.G.; Mumme, J. Hydrothermal Carbonization
of Biomass for Energy and Crop Production. Appl. Bioenergy 2014, 1, 11–29. [CrossRef]
13. Heilmann, S.M.; Molde, J.S.; Timler, J.G.; Wood, B.M.; Mikula, A.L.; Vozhdayev, G.V.; Colosky, E.C.;
Spokas, K.A.; Valentas, K.J. Phosphorus reclamation through hydrothermal carbonization of animal manures.
Environ. Sci. Technol. 2014, 48, 10323–10329. [CrossRef] [PubMed]
14. Shi, Y.; Luo, G.; Rao, Y.; Chen, H.; Zhang, S. Hydrothermal conversion of dewatered sewage sludge: Focusing
on the transformation mechanism and recovery of phosphorus. Chemosphere 2019, 228, 619–628. [CrossRef]
15. Wang, T.; Zhai, Y.; Zhu, Y.; Peng, C.; Wang, T.; Xu, B.; Li, C.; Zeng, G. Feedwater pH affects phosphorus
transformation during hydrothermal carbonization of sewage sludge. Bioresour. Technol. 2017, 245, 182–187.
[CrossRef] [PubMed]
16. Ovsyannikova, E.; Arauzo, P.J.; Becker, G.C.; Kruse, A. Experimental and thermodynamic studies of
phosphate behavior during the hydrothermal carbonization of sewage sludge. Sci. Total Environ. 2019, 692,
147–156. [CrossRef] [PubMed]
17. Mäkelä, M.; Benavente, V.; Fullana, A. Hydrothermal carbonization of lignocellulosic biomass: Effect of
process conditions on hydrochar properties. Appl. Energy 2015, 155, 576–584. [CrossRef]
18. Mäkelä, M.; Benavente, V.; Fullana, A. Hydrothermal carbonization of industrial mixed sludge from a pulp
and paper mill. Bioresour. Technol. 2016, 200, 444–450. [CrossRef]
19. Álvarez-Murillo, A.; Román, S.; Ledesma, B.; Sabio, E. Study of variables in energy densification of olive
stone by hydrothermal carbonization. J. Anal. Appl. Pyrolysis 2015, 113, 307–314. [CrossRef]
20. Román, S.; Libra, J.; Berge, N.D.; Sabio, E.; Ro, K.S.; Li, L.; Ledesma, B.; Álvarez, A.; Bae, S. Hydrothermal
Carbonization: Modeling, Final Properties Design and Applications: A Review. Energies 2018, 11, 216.
[CrossRef]
21. Kieseler, S.; Neubauer, Y.; Zobel, N. Ultimate and Proximate Correlations for Estimating the Higher Heating
Value of Hydrothermal Solids. Energy Fuels 2013, 27, 908–918. [CrossRef]
22. Wang, L.; Chang, Y.; Li, A. Hydrothermal carbonization for energy-efficient processing of sewage sludge:
A review. Renew. Sustain. Energ. Rev. 2019, 108, 423–440. [CrossRef]
23. Deutsches Institut für Normung e. V. Sludge, Treated Biowaste and Soil—Determination of pH (EN 15933:2012);
German Version; Beuth Verlag GmbH: Berlin, Germany, 2012; DIN EN 15933:2012-11.
24. Deutsches Institut für Normung e. V. Solid Biofuels—Sample Preparation (ISO 14780:2017); German Version;
Beuth Verlag GmbH: Berlin, Germany, 2017; DIN EN ISO 14780:2017-08.
25. Deutsches Institut für Normung e. V. Solid Biofuels—Determination of Moisture Content—Oven Dry Method—
Part 3: Moisture in General Analysis Sample (ISO 18134-3:2015); German Version; Beuth Verlag GmbH: Berlin,
Germany, 2015; DIN EN ISO 18134-3:2015-12.
26. Deutsches Institut für Normung e. V. Solid biofuels—Determination of Total Content of Carbon, Hydrogen and
Nitrogen—Instrumental Methods (EN 15104:2011); German Version; Beuth Verlag GmbH: Berlin, Germany,
2011; DIN EN 15104:2011-04.
27. Deutsches Institut für Normung e. V. Solid Biofuels—Determination of Ash Content (EN 14775:2009); German
Version; Verlag GmbH: Berlin, Germany, 2012; DIN EN 14775:2012-11.
91
Energies 2020, 13, 4417
28. Deutsches Institut für Normung e. V. Solid Biofuels—Determination of Calorific Value (ISO 18125:2017); German
Version; Verlag GmbH: Berlin, Germany, 2017; DIN EN ISO 18125:2017-08.
29. Deutsches Institut für Normung e. V. Water quality—Determination of Selected Elements by Inductively Coupled
Plasma Optical Emission Spectrometry (ICP-OES) (ISO 11885:2007); German Version; Beuth Verlag GmbH:
Berlin, Germany, 2009; DIN EN ISO 11885:2009-09.
30. Channiwala, S.A.; Parikh, P.P. A unified correlation for estimating HHV of solid, liquid and gaseous fuels.
Fuel 2002, 81, 1051–1063. [CrossRef]
31. Montgomery, D.C. Design and Analysis of Experiments, 8th ed.; John Wiley & Sons Inc.: Singapore, 2013;
ISBN 978-1-118-09793-9.
32. Weisberg, S. Applied Linear Regression, 3rd ed.; Wiley-Interscience: Hoboken, NJ, USA, 2005;
ISBN 9780471663799.
33. Wang, L.; Li, A. Hydrothermal treatment coupled with mechanical expression at increased temperature for
excess sludge dewatering: The dewatering performance and the characteristics of products. Water Res. 2015,
68, 291–303. [CrossRef] [PubMed]
34. Danso-Boateng, E.; Shama, G.; Wheatley, A.D.; Martin, S.J.; Holdich, R.G. Hydrothermal carbonisation of
sewage sludge: Effect of process conditions on product characteristics and methane production. Bioresour.
Technol. 2015, 177, 318–327. [CrossRef]
35. Reza, M.T.; Rottler, E.; Herklotz, L.; Wirth, B. Hydrothermal carbonization (HTC) of wheat straw: Influence
of feedwater pH prepared by acetic acid and potassium hydroxide. Bioresour. Technol. 2015, 182, 336–344.
[CrossRef]
36. Gao, N.; Li, Z.; Quan, C.; Miscolczi, N.; Egedy, A. A new method combining hydrothermal carbonization
and mechanical compression in-situ for sewage sludge dewatering: Bench-scale verification. J. Anal. Appl.
Pyrolysis 2019, 139, 187–195. [CrossRef]
37. Wang, L.; Zhang, L.; Li, A. Hydrothermal treatment coupled with mechanical expression at increased
temperature for excess sludge dewatering: Influence of operating conditions and the process energetics.
Water Res. 2014, 65, 85–97. [CrossRef]
38. Funke, A.; Ziegler, F. Hydrothermal carbonization of biomass: A summary and discussion of chemical
mechanisms for process engineering. Biofuels Bioprod. Biorefining 2010, 4, 160–177. [CrossRef]
39. Peterson, A.A.; Vogel, F.; Lachance, R.P.; Fröling, M.; Michael, J.; Antal, J.R.; Tester, J.W. Thermochemical
biofuel production in hydrothermal media: A review of sub- and supercritical water technologies.
Energy Environ. Sci. 2008, 1, 32–65. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
92
energies
Article
Comparative Studies on Water- and Vapor-Based
Hydrothermal Carbonization: Process Analysis
Kyoung S. Ro 1, *, Judy A. Libra 2, * and Andrés Alvarez-Murillo 3
1 USDA-ARS Coastal Plains Soil, Water & Plant Research Center, Florence, SC 29501, USA
2 Leibniz Institute for Agricultural Engineering and Bioeconomy, 14469 Potsdam, Germany
3 Department of Applied Physics, School of Industrial Engineering, University of Extremadura, 06006 Badajoz,
Spain; andalvarez@unex.es
* Correspondence: kyoung.ro@usda.gov (K.S.R.); jlibra@atb-potsdam.de (J.A.L.);
Tel.: +1-843-669-5203 (K.S.R.); +49-331-5699-866 (J.A.L.)
Abstract: Hydrothermal carbonization (HTC) reactor systems used to convert wet organic wastes
into value-added hydrochar are generally classified in the literature as liquid water-based (HTC) or
vapor-based (VTC). However, the distinction between the two is often ambiguous. In this paper,
we present a methodological approach to analyze process conditions for hydrothermal systems.
First, we theoretically developed models for predicting reactor pressure, volume fraction of liquid
water and water distribution between phases as a function of temperature. The reactor pressure
model predicted the measured pressure reasonably well. We also demonstrated the importance of
predicting the condition at which the reactor system enters the subcooled compression liquid region
to avoid the danger of explosion. To help understand water–feedstock interactions, we defined a
new solid content parameter %S(T) based on the liquid water in physical contact with feedstock,
which changes with temperature due to changes in the water distribution. Using these models,
we then compared the process conditions of seven different HTC/VTC cases reported in the literature.
This study illustrates that a large range of conditions need to be considered before applying the label
VTC or HTC. These tools can help in designing experiments to compare systems and understand
results in future HTC research.
1. Introduction
The process of hydrothermal carbonization (HTC) is used to carbonize organic residues and
wastes for diverse applications ranging from fuels to soil amendments. In HTC, subcritical water is
used as a solvent and reactant to transform a wide variety of organic feedstocks to solid carbonaceous
products (hydrochar), which usually contain higher carbon contents, heating values and degrees of
aromaticity than the original feedstocks [1–4]. Diverse types of reactors have been used, ranging from
batch [3,5–12], semi-batch [2,9,10] to continuous reactors [13], with and without mixing, using direct
heating through steam injection or through the reactor walls with controlled heating rates, or indirectly
in muffle ovens. The pressurized reaction system usually consists of all three phases (gas, liquid, solid)
and is heated to temperatures from 160 to 280 ◦ C with pressures between 0.6 and 6.4 MPa due to
the water vapor and gases produced in the reactions. Water is initially introduced into the reaction
system via the moisture content of the feedstock and/or through the addition of water as a liquid or
as steam. Common HTC process variations are differentiated according to how the water initially
contacts the feedstock. When the feedstock is immersed in bulk liquid water, it is often called HTC.
Where the feedstock is in direct contact with only steam, it is often called steam HTC, vapor HTC or
93
Energies 2020, 13, 5733
vapothermal carbonization (VTC) [2,7,10,11]. Here, feedstock can be held in baskets away from liquid
water, transported in and out on conveyor belts with little to no post-processing dewatering steps
necessary. However, the dividing line between the two process types often does not remain sharp over
the operating time, since the distribution of water between the liquid and vapor phases will change
as temperature rises in batch systems, or as more steam is added in semi-batch systems. Usually we
cannot see inside high-pressure reaction systems to gain visual insights, so the extent of the phase
change is often unknown. Process comparisons between HTC vs. VTC based on the state of the initial
water phase may be misleading. The phase changes may play an important role in causing structural
variations in the hydrochars produced, their biodegradability, stability, and functionality in various
applications [5].
Only a few studies comparing hydrochars produced by HTC vs. VTC have been reported in the
literature. Cao et al. [5] found that biomass was more carbonized under liquid water in HTC than
through steam in VTC. They determined that more aromatic and less alkyl groups were formed in the
sugar beet and bark hydrochars made from HTC than in those from VTC under the same operating
conditions (200 ◦ C for 3 h). The hydrochars made from HTC also were less biodegradable than those
from VTC as indicated by the lower values of the ratio BOD/COD for HTC-hydrochars. The lower
biodegradability of hydrochar from HTC was probably due to its higher aromaticity produced during
the reactions taking place in liquid water. In this case, the hydrochar made from HTC should be
more stable in the environment than that made from VTC, which would have consequences for its
use, e.g., as a soil amendment. On the other hand, comparisons between hydrochars from HTC vs.
VTC for use as fuels have produced mixed results on how the phase of the reaction medium affects the
important energy parameters: solid yield, higher heating value (HHV) and energy yield [10,11,14].
Comparing HTC vs. VTC for two feedstocks, digestate and straw, at 230 ◦ C, 6 h, Funke et al. [11] found
no clear trend in HHV: VTC-chars had a higher HHV than HTC-chars from digestate, but for straw,
the order was reversed. However, VTC produced higher solid yields and, therefore, higher energy
yields than HTC for both feedstocks. In contrast, Shafie et al. and Yeoh et al. reported mostly lower
solid yields and higher HHV from the VTC process compared to HTC process (both at 220 ◦ C, 1 h;
two feedstocks [10,14], three feedstocks [14]). Again, as in [11], the combination of these two trends in
the energy yield showed that the VTC process was more efficient for energy yield. We suspect that the
reason for these conflicting trends in HHV and solid yield arise from the fact that a clear picture of
how the water was distributed between gas and liquid phases was not given, nor to which extent the
feedstocks were exposed to the liquid water phase. For instance, the VTC experimental setup used
by Yeoh et al. (two concentric chambers with water filled in the outside chamber and the biomass in
the inner chamber) was assumed to avoid the liquid water directly contacting the biomass feedstock.
Yet it is not clear from their description that the liquid water was contained in the outside chamber
throughout the reaction, since liquid water expands at higher temperature, possibly causing overflow
into the biomass chamber. It appears that the transition between HTC/VTC reported in the literature is
fuzzy at best because it can be changed by small variations in the same reactor system. How much
water is present in each phase depends not only on the reactor temperature and pressure, but also on
the total amount of water in the system relative to the volume of the reactor system.
As more knowledge is gained on the beneficial applications of hydrochars and HTC for waste
and residue processing, more work on reactor designs for diverse settings (ranging from high to
low tech systems) will be carried out. The variety of process variations may increase, with process
configurations and conditions utilizing the unique transport properties of each medium, e.g., the higher
thermal conductivity of bulk liquid water or the higher diffusivity of steam to penetrate the porous
structure of the feedstock [7,15]. These changes in transport properties can affect reactions and product
characteristics [15]. Therefore, knowledge on what influences the distribution of water between the
reaction phases is essential for the production of the desired hydrochar quality. Especially if we want
to replicate process conditions in various reactor types and scales to produce a desired hydrochar
94
Energies 2020, 13, 5733
quality, we must be able to predict the distribution of water between the vapor and liquid phases at the
design HTC reaction conditions.
Furthermore, reactor designs must consider how the HTC reactor pressure will change in response
to operating conditions to ensure process safety. All reactors must be able to withstand the high
temperatures and pressures that can develop during the process. As a rigid HTC reactor partially filled
with water and feedstock is heated, the increase in the saturated water vapor and gases produced by
the chemical reactions cause the reactor pressure to rise. At the same time, the density of the bulk
liquid water decreases and consequently the volume of liquid water increases, decreasing the volume
of the reactor headspace. When the liquid volume in the HTC reactor completely fills the headspace,
it can no longer expand if the reactor temperature is increased further. The reactor water then enters
a subcooled liquid compression region. In this region, pressure increases very rapidly with small
increases in reactor temperature. To avoid the reactor pressure exceeding the tensile strength of reactor
material, it is very important that the reactor system has a working safety disk or valve that can release
pressure at a preset value. Without the use of proper rupture disks, the reactor can explode. Therefore,
in order to maintain safe operating conditions, we need to predict the reactor pressure at the chosen
process conditions. This requires understanding the relationship between the HTC reactor conditions
(temperature, water volume, feedstock) and pressure.
The aim of this work is to present a methodological approach to analyze process conditions
for hydrothermal systems in the framework of the hydrothermal carbonization reactions. In the
paper, we first theoretically develop models for predicting reactor pressure, the distribution of water
between phases, and the liquid water volume fractions as a function of reactor temperatures. Then,
the evaluation is expanded to water and feedstock. Finally, using these new models, we analyze and
compare process conditions for VTC and HTC systems reported in the literature.
95
Energies 2020, 13, 5733
Figure 1. Temperature-volume (T-v) diagram for water showing the common operating region for
vapor and hydrothermal carbonization reactions.
where
MH2O = total mass of liquid and vapor water in the reactor (kg);
xL = mass fraction of liquid water;
xV = mass fraction of vapor water (or steam quality);
xL + xV = 1.
As the HTC reactor is heated beyond the boiling temperature, the reactor volume is mostly filled
with liquid water and steam, and the following relationship can be developed assuming both liquid
water and steam are in equilibrium (i.e., for the saturated liquid–vapor region; Figure 1).
where
VR = reactor volume (m3 );
vL = specific volume of saturated liquid water (m3 /kg);
vV = specific volume of saturated steam (m3 /kg).
Combining Equations (1) and (2), the mass fraction of vapor water (xv ) can be calculated by
knowing the thermophysical properties of water at those conditions, the mass of water in the reactor
and the reactor volume:
vR − v L
xV = (3)
vV − vL
where
vR = VR /MH2O , overall specific volume of reactor water and steam mixture (m3 /kg).
As the liquid-steam mixture in the reactor is heated, the volume of liquid water expands due to
the decrease in water density ρL. Using Equation (3) along with values for saturated vapor and liquid
96
Energies 2020, 13, 5733
specific volumes [16], the fraction of liquid-water occupying the reactor volume VFw can be estimated
at the HTC reaction temperature:
Vw ( 1 − xV ) v L
VFw = = (4)
VR vR
where
Vw = volume of liquid water in the reactor at temperature T (m3 );
VFw = volume fraction of liquid water in the reactor at temperature T (-).
As long as the reactor volume is larger than the bulk liquid water volume (i.e., VFw < 1), we can
assume the liquid and vapor water phases are in equilibrium and the autogenic pressure can be
estimated from the saturation properties of water using saturated steam tables [16–18].
If the temperature is further increased so that the liquid volume completely fills the reactor due
to the decrease in its density (i.e., VFw = 1, and xL = 1), the liquid water will enter the subcooled
liquid compression region. This region can be seen in the T-v phase diagram, left of the saturated
vapor curve (Figure 1). There is no longer any headspace in the reactor and the water density in
the reactor system at this point (also called overall reactor water density) becomes constant and can
be calculated from D = MH2O /VR . As the rigid reactor walls are suppressing the tendency of the
liquid volume to increase in response to the decrease in liquid water density, the reactor pressure
increases rapidly as the water expands with the increase in temperature. When VFw > 1 calculated
from Equation (4) (i.e., physically impossible unless the reactor explodes), the reactor pressure in this
range can be estimated with liquid compressibility factor for subcooled water:
P = ZL × D × RT/MWH2O (5)
where
P = reactor pressure (MPa);
ZL = liquid compressibility factor for subcooled water (-);
D = overall reactor water density, MH2O /VR (kg/m3 );
R = universal gas constant (8.31451 × 10−3 m3 -MPa/kmol-K);
T = reactor temperature (K);
MWH2O = molecular weight of water (kg/kmol).
To illustrate the danger of a potential reactor explosion if the liquid fills the reactor completely,
example calculations to estimate the reactor pressure at three common HTC temperatures using
Equation (5) are reported in Table 1. The values of liquid compressibility factor of the subcooled water
reported by Lemmon et al. (2018) were used. A value for D was chosen that is slightly higher than the
saturated liquid water density at 200 ◦ C. This simulates the reactor pressure for the case when the
liquid water fills the reactor completely at 200 ◦ C. A further increase in T to 250 ◦ C will rapidly increase
P from 2 to 81.6 MPa, a pressure that many HTC reactors are not made to withstand. For instance,
maximum allowable pressures for common laboratory reactors range from 13.3 to 34.5 MPa [19].
In contrast, if there is less liquid water added and more headspace in the reactor so that the liquid
water-vapor equilibrium can exist at all operating temperatures, the pressure increase would follow
the saturation pressure, increasing only from 1.6 to 4.0 MPa.
In order to avoid the subcooled compressible region, some manufacturers of pressure equipment
recommend calculating the maximum allowable water mass using a safety factor and the ratio of
ρL or its inverse vL at the desired T to that at room temperature [20]. It is also important to note
that the actual HTC pressure will be higher than that from the pure water because of gas production
(predominantly CO2 ) from HTC reactions.
97
Energies 2020, 13, 5733
Table 1. Example calculations for the effect of increasing the hydrothermal carbonization (HTC)
temperature on reactor pressure for D = 865 kg/m3 (values for saturated water properties taken from
Lemmon et al. (2018).
98
Energies 2020, 13, 5733
is shown in Figure 2 for a reactor filled with water only. Values for VFw were estimated at various
temperatures and VFo using Equation (4). For a reactor initially filled with water at 90% (i.e., VFo = 0.9),
the liquid volume expands to the reactor volume (i.e., VFw = 1) when the reactor temperature reaches
165 ◦ C. Fortunately, this critical temperature, at which VFw =1, increases rapidly as VFo is decreased,
e.g., 305 ◦ C for VFo = 0.7 and 365 for VFo = 0.5, so that process conditions can be chosen to remain
well below the critical temperature. When the reactor is initially filled with water to less than half its
volume (i.e., VFo < 0.5), the liquid does not fill the reactor even when the temperature approaches the
critical point of water around 374 ◦ C. Interestingly, for experiments with low values of VFo common to
VTC operating conditions, VFw can actually decrease with temperature. When the reactor is initially
filled with a very low volume of water, such as VFo = 0.1, the liquid volume decreases to zero at
T = 340 ◦ C. This happens when there is so much headspace that the liquid water completely vaporizes,
i.e., the molecular collision frequency of H2 O molecules in the headspace is so small that condensation
does not happen in this high headspace situation.
9)R
9)R
R
7 &
Figure 2. Change in volume fraction of reactor filled with liquid water VFw (Vw /VR ) as a function of
temperature for various initial liquid water volume fractions (VFo ).
99
Energies 2020, 13, 5733
line at temperatures up to 349 ◦ C (Figure 3). However, for VFo = 0.67 at temperatures higher than
349 ◦ C, the pressure increased much more rapidly than the saturation vapor pressure, indicating that
the water entered the subcooled compression region. According to the predicted pressure line for the
estimated VFo , it should have entered the subcooled region at a lower temperature of 320 ◦ C. Instead,
the observed pressure followed the predicted pressure line of VFo = 0.6 (Figure 3). We suspect that this
discrepancy can be attributed to the fact that the actual reactor volume was larger than the estimated
volume, (e.g., 1.1 L instead of 1 L for 0.67 L water initially filled). These results show that the approach
is adequate to estimate VFw at various reactor temperatures in practical applications, however, if the
actual reactor volume is not known accurately, there will be some deviation from the predicted values.
3 03D
Figure 3. Comparison of measured and estimated reactor pressure at different initial liquid water
volume fractions (VFo ).
Another important question to consider in deciding upon operating conditions and evaluating
experimental results is: How does a higher initial pressure affect the pressure development and phase
distribution of water in the reactor? For example, some experimenters pressurize the system initially
using an inert gas such as N2 or Ar. The answer in short is that the addition of pressure to the reactor
headspace does not change the behavior of water. If enough water and time are available, water will
vaporize to the gas phase to reach the saturation water vapor pressure at which liquid water and
vapor water are in equilibrium. This pressure is a function of temperature only and independent of
the presence of other gases. The added inert gas does not change the relationships for VFw and the
distribution of water between the liquid and vapor phases. However, the total reactor pressure will be
higher in the reactor initially filled with N2 or Ar than that without the initial inert gases. The total
pressure P(total) can be estimated by summing the partial pressures of all individual non-reacting gases
as stated in Dalton’s law. The increase in the partial pressure for each component with temperature
can be calculated independently and added together. This can be seen in Figure 4 for two experimental
runs in an 18.75-L Parr reactor in which water was heated to 220 ◦ C (VFo = 0.63): one starting at
atmospheric pressure and the second one with N2 addition to achieve an initial pressure of 1.4 MPa.
The measured values from the nonpressurized run (Po = 0.1 MPa) are compared to the saturation water
vapor pressure P(sat) from [2] in the lower curve. For the run at Po = 1.4 MPa, the partial pressure
increase for N2 P(N2 ) was estimated using the ideal gas law, combined with Equation (4) to calculate
the changes in headspace volume (1-VFw ) as temperature increases.
Comparison of the measured and theoretical values shows clearly that the contribution of the
saturated water vapor to the total pressure is not affected by the initial addition of N2 gas. The small
deviation between the calculated pressure and the measured can be due to inaccuracies in the pressure
measurement or in estimating the reactor volume, and the assumption that N2 behaves as an ideal
gas with no solubility in the liquid. Nevertheless, the difference does not mask that the fact that the
addition of pressure to the reactor headspace does not change the behavior of water.
100
Energies 2020, 13, 5733
ϳ
ϲ
W;ƚŽƚĂůͿ
ϱ
W;EϮͿ
ϰ
W;DWĂͿ
W;ƐĂƚͿ
ϯ
W;džƉ͘ĂƚĂͿ
Ϯ
W;ƐĂƚ͕džƉ͘ĂƚĂͿ
ϭ
Ϭ
Ϭ ϭϬϬ ϮϬϬ ϯϬϬ
d;ΣͿ
Figure 4. Comparison of the theoretical total pressure P(total) calculated from the partial pressures
for N2 P(N2 ) and saturated water vapor P(sat) to the measured values for pressure P and P(sat) in the
reactor for VFo = 0.63.
3.2.1. Estimating the Distribution of Water between Phases as a Function of Temperature and Its Effect
on Solid Content
In their comparison of hydrochars from VTC and from HTC systems, Cao et al. (2013) postulated
that the amount of liquid water in contact with the feedstock in the reaction system may determine the
degree of carbonization and influence which reactions take place and their sequence [5]. However,
they did not quantify how much liquid water was in contact with the feedstock in their reaction
systems. This is a common problem in most of the literature on HTC/VTC systems. Often the label
used for the system is defined by the initial conditions. For instance, when the feedstock is initially
completely submerged in bulk liquid water, it is commonly called an HTC system. Whereas, when dry
or wet feedstock is placed separately from the bulk liquid water, it is called a VTC system. However,
the volume of liquid water and the distribution of water between the liquid and vapor phase change
with temperature, which can change the amount of water contacting the feedstock. In addition,
the feedstock characteristics such as moisture content, particle size, bulk density, as well as structural
changes during the reaction can affect how water interacts with the feedstock. In VTC, carbonization
reactions can take place between a wet feedstock and water within its cells or present as a film on its
surface [10,11]. Ref. [11] Even with completely dried feedstock, the feedstock can be wetted during the
process by absorbing water vapor or water vapor condensing on its surface.
The parameters often used to describe the relationship between water and feedstock in a reaction
system do not differentiate between the bulk liquid water added to the process and the liquid water
in contact with the feedstock. The nominal solid content at the start of the run is usually reported in
published studies as:
Mbiomass
%So = × 100 (6)
MH2O + Mbiomass T=To
where
%So = nominal solid content;
Mbiomass = initial feedstock dry mass;
MH2O = total mass of water in the reactor.
A similar parameter R which describes the initial ratio of feedstock dry mass to total mass of
water is also often used. These parameters only describe the initial conditions based on the initial
filling masses of water and feedstock, but do not provide critical information on the extent to which
101
Energies 2020, 13, 5733
feedstock is exposed to liquid water in the HTC or VTC systems to promote important hydrothermal
carbonization reactions. In order to provide useful information on the degree of physical contact
between the feedstock and liquid water throughout the process, we propose reporting the following
solid content parameter:
Mbiomass
%S(T ) = × 100 (7)
mH2O + Mbiomass T=T
where
%S(T) = actual solid content based on liquid water in contact with feedstock;
mH2O = mass of liquid water in contact with feedstock;
T = reactor temperature.
With these new definitions, one can systematically distinguish various HTC/VTC process
conditions in terms of fraction of liquid water physically in contact with feedstock. For HTC
systems, where the feedstock is assumed to be completely submerged in the bulk liquid water over
the whole reaction time, mH2O = xL . MH2O . Using Equations (3) and (4), these assumptions can
be checked for the reaction temperature and the solid content values adjusted with Equation (7).
For example, the change in the distribution of water between the two phases can be seen in Figure 5a.
For temperatures below 250 ◦ C and VFo larger than 0.3, less than 4% of the water will be vaporized.
The expansion of VFw, as seen in Figure 2, should offset the small loss of liquid water to the vapor phase
and submerged feedstocks should remain submerged at these conditions. Therefore, the actual solid
content will be approximately the same as the nominal solid content at the initial reactor temperature
To (i.e., %S(T) = %So ). Only for systems with VFo closer to 0.1, more common to VTC systems,
will approximately 20% of the water be present as vapor at 250 ◦ C.
ďͿ
9)R
9)R
9)R
9)R
9)R
7 R&
Figure 5. Changes in process parameters at various initial VFo as the operating temperature increases
for (a) mass fractions of water in liquid xL and vapor xV , (b) the ratio of actual to nominal solid content
in the reactor system, %S(T) at T to %So at the initial temperature.
102
Energies 2020, 13, 5733
In VTC systems, wetted or completely dried feedstock can be suspended without any physical
contact with bulk liquid water. The bulk liquid water can be placed either at the bottom of the reactor
or in a separate interconnected chamber, or steam can be injected to heat the reactor. The value
reported for %So for such systems often includes the bulk water. However, this can be misleading,
especially for dried feedstock, where the actual initial solid content %S(To ) = 1 because m2O = 0.
Although %S(To ) = 1 initially, %S(T) will become less than one over time because water vapor will be
volatilized from the physically separated bulk liquid as the VTC reactor temperature increases and will
condense on the surface of the dry feedstock. The extent to which the vaporized water condenses onto
the feedstock depends on the kinetics of condensation and vaporization at the reaction temperature,
but %S(T) will rarely reach %So . The condensed water will promote typical hydrolysis and other
important carbonization reactions as in HTC systems. For VTC systems with initially wetted feedstock,
bulk liquid water may or may not be added to the reactor. If no bulk liquid water is added similar to
that of Funke et al. [11], %S(To ) = %So , where the moisture content (MC) determines the value of the
initial solid content. As the temperature increases, liquid water is lost to vaporization, reducing the
water content of the feedstock. %S(T) can then be calculated with Equation (7) and:
Mbiomass × MC
mH2O = xL × MH2O = xL × (8)
1 − MC
Using these equations for the new solid content parameters, the ratios of actual to nominal solid
content are plotted against the reactor temperatures in Figure 5b for various initial volume fractions
of liquid water. For HTC systems with VFo larger than 0.3 and temperatures below 250 ◦ C, there is
little difference between the two values. At 250 ◦ C, only a 4% increase is seen in %S(T), and after the
reactor is half-filled (i.e., VFo > 0.5), no differences are noticeable. In contrast, for systems with low
values of VFo (e.g., VFo = 0.1), %S(T) becomes 20% higher than %So . Less liquid water is in contact
with the solids to participate in reactions. For VTC systems with wet feedstocks where the liquid
water is associated in or on the feedstock, the transfer of water to the vapor would change the %S(T)
in the reactor significantly. For wetted feedstock suspended over bulk liquid water, the situation is
more complicated, since water can vaporize from the wet feedstock or bulk liquid water, mass transfer
within and in-between feedstock, and the kinetics of condensation and vaporization all play roles in
the location of the liquid water. This is beyond the scope of this paper. Moreover, the implications of
the reduction in the mass of bulk liquid water on the potential of reducing physical contact between
feedstock with a bulk volume larger than that of liquid water and subsequent carbonization reactions
need to be further studied.
3.2.2. Estimating VFw and Pressure under Process Conditions with Feedstock
Up until now. we have analyzed the changes in volume fractions due to changes in the physical
properties of water. The addition of feedstock to the reactor can reduce the headspace volume available
to accommodate the expansion of liquid water. To adjust the volume fractions for the presence of
feedstock, the reactor volume must be reduced by the volume occupied by the feedstock. To strictly
determine this volume, we need to know the true density of the material, i.e., the ratio between the
feedstock mass and its volume excluding the cavities, pores and gaps in the material where water and
air could be trapped. In addition, the loss of solid mass and structure during HTC reactions would
have to be taken into consideration. For practical purposes, simple estimates of the initial volume of
feedstock can be made with liquid displacement methods and used to adjust the calculation of VFw .
HTC reactions with the feedstock can also change the gas composition and pressure of the
headspace. The composition and amount of the produced gases is closely tied to the process conditions
and feedstock material. In general, most HTC reactions with biomass produce predominantly CO2
(~>80%) with minor percentages of N2 , H2 S, O2 , CH4 , H2 , etc., in the gas besides water vapor. Explosive
gas mixtures are not expected in HTC, unlike hydrothermal gasification where reactor temperatures
are near or above the critical temperature of water (i.e., ~374 ◦ C) and produce approximately 1:1 of CH4
103
Energies 2020, 13, 5733
and CO2 . However, an in-depth analysis about gas production and compositions is beyond the scope
of this paper. The impact of the product gas on the total reactor pressure and its partition between
water and gas phases need to be further investigated as gas solubility changes with temperature and
pressure. The results of this study provide a theoretical framework for further experimental and
modeling research on this aspect of HTC.
3.3. Comparison of Process Conditions for Hydrothermal Treatment (HTC and VTC) Reported in the Literature
The results from published HTC/VTC studies that have been made at various scales, ranging
from 1 L to 10 m3 , and with different modes of operation, e.g., batch and semi-batch with respect to the
steam, are analyzed in this section. As summarized in the introduction, few studies comparing HTC
and VTC systems have been published and some results are contradictory. The goal here is to identify
the effect of process conditions on the distribution of the water between the phases to understand what
is behind the labels—HTC and VTC—and develop criteria on how to label systems, either HTC or
VTC. This is necessary especially for cases in which we want to replicate process conditions used to
produce a desired hydrochar quality in other reactor types and/or scales. The results are structured
into seven cases for the discussion here and an overview of the process conditions and feedstocks is
given in Table 2.
Table 2. Overview of the process conditions and feedstock for the seven cases with VTC/HTC processes.
Case Reactor Feedstock Water in System Literature
T VR MC Mbiomass MH2O vR
Type Heating Mixed Type
(◦ C) (L) (%) (gDM) (g) (m3 /kg)
bark 7 23.5 101.6 0.010
1 Batch HTC 200 1 heating band No
sugar beet 7 37.6 162.9 0.006
Cao et al. [5]
Semi- batch steam, condensate bark 40–70 n.r. n.r. 0.013 *
2 200 70 No
VTC removal sugar beet 40–70 n.r. n.r. 0.012 *
start start
Semi- batch
steam, no condensate 1998.1 0.005
3 VTC 200 10,000 Yes MSW 53 1771.9 Safril et al. [9]
removal end end
3398.1 0.0029
wheat straw 0 450 8550 0.00219
4 Batch HTC 230 18.75 heating band No
digestate 0 630 11,970 0.00157
Funke et al. [11]
wheat straw 25 450 1350 0.01390
5 Batch VTC 230 18.75 heating band No
digestate 25 630 1890 0.00995
bagasse 67.2 16.6–28.8 1495.5 0.003
6 Batch HTC 220 4.6 heating band No
lime peel 78 14.4–44.0 1495.5 0.003
Shafie et al. [10]
Semi- batch heating band, bagasse 67.2 n.r. n.r. 0.041–0.053 @
7 220 4.6 No
VTC condensate separated lime peel 78 n.r. n.r. 0.022–0.044 @
n.r.—not reported. MC – moisture content. * Assumed 50% reactor volume filled with bark or sugar beet feedstock
(MC = 55%) suspended in baskets with bulk density of 0.267 kg/L for bark [21] and 0.298 kg/L for sugar beet pulp [22].
@ Assumed the same amount of feedstock as in Case 6 with biomass bulk density of 0.616 kg/L for bagasse [23] and
In Cases 1 and 2, a comparison of batch HTC (Case 1) with semi-batch VTC (Case 2) using the
same feedstocks was made in different reactor systems [5]. For Case 1, the feedstock was dried, ground,
and water added to the 1 L reactor, submerging the feedstock. In Case 2, the wet feedstock was
suspended in baskets and steam (1.6 MPa) was injected to heat the 70 L reactor. As water condensed
over the heating up and holding time, it was removed except for that which remained on the feedstock.
Therefore, there was no increase in the mass of condensed water in the reactor over time (Revatec
GmbH, DE 10 2009 010 233.7). The feedstock was reported to have a moisture content between 40 and
70%. For the calculations here, the mass of water in the system was estimated from the overall specific
volume of saturated water vapor at 200 ◦ C in the 70 L reactor and an average moisture content of
the feedstock (55%) during the HTC reaction. As feedstock loading information is not available,
we assumed 50% of the reactor volume was filled with bark or sugar beet pulp suspended in baskets
104
Energies 2020, 13, 5733
inside the reactor. The mass of water was estimated using bulk densities of bark and sugar beet pulp
reported in the literature [21,22].
In Case 3, a commercial-scale hybrid system with municipal solid waste (MSW) in a 10 m3 reactor
system was used [9]. The semi-batch system was fed saturated steam (2.5 MPa) to heat the feedstock
and held at 180–230 ◦ C for 30 min. The system started similar to a VTC system with only wet feedstock,
but as the condensed steam was mixed with the feedstock over time, the system became more similar
to HTC. A mid-range temperature of 200 ◦ C was assumed for further analyses here. In Cases 4 and
5, Funke et al. compared VTC and HTC for two feedstocks in the same batch reactor system [11].
For VTC, the dried feedstock was first soaked in water and then suspended in a basket without any
additional liquid water added to the reactor. For HTC, the dried feedstock was submerged in water.
In Cases 6 and 7, Shafie et al. used bagasse with MC of 67.16% (cut into less than 10 mm) and lime peel
with MC of 78.04% (size as received) as feedstock [10]. For HTC, the feedstock was fully submerged
inside the water in a reactor. For VTC, saturated steam was supplied to the reactor with the feedstock
suspended in order to avoid contact with condensed liquid water accumulated at the bottom of the
reactor. As feedstock loading information was not reported for the VTC experiments, we assumed
the mass of feedstock was the same as that used in HTC experiments, along with bulk densities from
the literature (bagasse [23]; lime peel (value for lemon peel was used [24]). The process conditions
and feedstock for each case are summarized in Table 2, along with the respective vR, overall specific
volume of reactor liquid water and steam mixture.
3.3.1. Change in Process Conditions in Batch HTC (Cases 1, 4, 6) or VTC Processes (Case 5)
In batch reactors, the solids and liquids are introduced at the beginning of the run and the reactor
is sealed before heating starts. The initial VFo and %So can be easily calculated and are usually reported
(Table 3). Feedstock initially submerged in water can unequivocally be called HTC when VFw at the
holding temperature remains as large or larger than VFo . This is true for all batch HTC cases (1, 4,
6) analyzed here. Each case includes results for two feedstocks under slightly differing conditions.
In Case 4, with a relatively large amount of initial water (VFo = 0.46 and 0.64, for wheat straw and
digestate, respectively), the expansion of water at 230 ◦ C causes VFw to increase by approximately 20%.
As only 0.5 to 1.4% (m/m) of the initial liquid water is transferred to the vapor phase, there is little to
no change in %S(T). Similarly, very little increase in solid content is observed in Case 6. In contrast,
Case 1 at 200 ◦ C has a low degree of initial water filling for both feedstocks (i.e., VFo = 0.1 and 0.16),
and VFw is very similar to VFo . Between 4 and 7% of the water is transferred to the vapor, causing a
corresponding increase in the value of %S(T). The values for %S(T) ranged from 1.0 to 19.9% for all
batch HTC cases, ensuring adequate contact with liquid water to promote HTC reactions. Despite
the loss of liquid water due to vaporization, the filling volume (VFw ) slightly increases because the
volume of water expands with the reactor temperature, guaranteeing that the feedstock is completely
submerged in the liquid water throughout the reaction period. Therefore, hydrothermal reactions will
take place between the feedstock and liquid water and the process can be called batch HTC in Cases 1,
4, and 6.
If the solids are suspended in the reactor in baskets or on trays so that they are not submerged in
water, the process is commonly called VTC (Cases 5). If the feedstock has a high moisture content such
as the dried feedstock soaked in water (Case 5, MC = 75% or %So = %S(To ) = 25%), or it is made up of
intact microorganisms or fresh plant material, the actual %S(T) slightly increases compared to %So
(27.6 or 28.9 vs. 25%) due to the small loss of liquid water in the feedstock to vapor (Table 3).
105
Energies 2020, 13, 5733
Table 3. Overview of process conditions and water distribution for the seven cases of VTC/HTC.
Case Reactor Feedstock vR Mass Fraction in Vapor xv VFo VFw %So %S(T)
Type Type (m3 /kg) (-) (-) (-) (%) (%)
bark 0.010 0.069 0.1 0.11 18.8% 19.9%
1 Batch HTC
sugar beet 0.006 0.04 0.16 0.18 18.8% 19.4%
bark 0.013 * 0.051 * n.r. - 45%# 45% #
2 Semi-batch VTC
sugar beet 0.012 * 0.046 * n.r. - 45%# 45% #
start 0.005 start 0.03 0.2 47.0%
3 Semi-batch VTC MSW
end 0.0029 end 0.0142 0.34 34.6%
wheat straw 0.00219 0.014 0.46 0.54 5.0% 5.1%
4 Batch HTC
digestate 0.00157 0.0051 0.64 0.77 5.0% 5.0%
wheat straw 0.01390 0.1804 0.07 0.07 25.0% 28.9%
5 Batch VTC
digestate 0.00995 0.1239 0.1 0.11 25.0% 27.6%
bagasse 0.003 0.03 0.33 0.378 1.1–1.9% 1.1–2.0%
6 Batch HTC
lime peel 0.003 0.03 0.33 0.378 1.0–2.9% 1.0–3.0%
bagasse 0.041–0.053 @ 0.473–0.609 @ n.r. - 32.8% 32.8%
7 Semi-batch VTC
lime peel 0.022–0.044 @ 0.251–0.510 @ n.r. - 22.0% 22.0%
n.r.—not reported. * Assumed 50% of reactor volume filled with bark or sugar beet suspended in baskets. # Averaged
value for feedstock. @ Assumed the same amount of feedstock as in Case 6.
3.3.2. Change in Process Conditions for Semi-Batch VTC Process with Steam Injection (Cases 2 and
7 with Condensate Removal or Separation, and Case 3 without Condensate Removal)
For reactors in Cases 2 and 3 with a semi-batch mode of operation where saturated steam is
introduced over time to first heat the reactor and then to maintain the desired operating temperature,
the calculations for how much mass of the water is present as liquid or vapor are not as straightforward.
The steam condenses as it heats the feedstock to the targeted operating temperature, and more will
condense over the targeted holding time. As steam is introduced, the reactor pressure will remain
constant at the saturation pressure if there are no reaction products entering the vapor phase. However,
gases are normally produced by the hydrothermal reactions and the pressure rises as the gases, mainly
CO2 , enter the headspace.
For systems with condensate removal, as in Case 2, or condensate separation, as in Case 7,
the mass of bulk liquid water in contact with feedstock comes from moisture already present within
the feedstock and water condensation on the surface of feedstock. Wet feedstocks will retain most
of their moisture. For dried feedstock, the majority of the water in the system will be in the vapor,
with some steam condensing on the feedstock surface, especially in the heating phase. Assuming that
the amount of steam condensed on the feedstock surface is negligible, the overall specific volume vR is
mostly that of the saturated water vapor and the moisture content of the feedstock. In such systems,
the process can be labeled VTC without much ambiguity. The amount of liquid water that can react
with the feedstock for VTC systems mainly depends on the moisture content of the original feedstock
and the condensed water on the feedstock surface. It is very difficult to quantify this amount of water.
For these two cases, %S(T) was assumed to remain the same as the initial value. The condensed water
is sometimes flashed off at the end of the run (e.g., for energy recovery (Revatech, 2012)), so that the
solids come out about as wet as they went in. This is helpful in reducing dewatering requirements,
but this hinders easily assessing how much water was in contact with the feedstock.
In systems without condensate removal, as in Case 3, the continuous injection of steam will build
up the total mass of water in the system, with the majority present in the form of liquid water. The VTC
process then approaches the HTC process. In this hybrid VTC–HTC commercial-scale unit, the reaction
system is well-mixed and liquid water is mixed into the feedstock, gradually lowering the value of
%S(T) in the reactor from the initial %So value 47% to 34.6%. The volume fraction of vapor water (xV )
changes somewhat. Starting with 3% of the water present as vapor, it reduces to 1.4% at the end of the
run (Table 3). In general, it is important to measure the mass of steam introduced in systems without
106
Energies 2020, 13, 5733
condensate removal, so that the mass of accumulated condensed water can be monitored as a safety
precaution. Steam injection must stop before VFw approaches 1 to avoid rupture of the reactor.
It is interesting to note that the solid content %S(T) of feedstock for all seven cases was less than
45% at the reaction temperature. The solid content for HTC systems ranged from 1.0 to 19.9%, while
that for VTC systems from 27.6 to 45%. It means that 55% or more of the total water mass was present
as liquid water and had direct physical contact with feedstock promoting carbonization reactions.
According to Cao et al., the lower the %S(T), the more the product was carbonized. The highest solids
content for VTC was 45% in Case 2 because of the water already present within the raw feedstock even
though additional water was not supplied. This leads to questions on what will happen if we conduct
VTC with completely dried feedstock, such as: Is it possible to carbonize the dried feedstock with
steam alone? For such reaction systems, the initial value of %S(To ) equals one. The value of %S(T) will
become less than one as some of the steam condenses on the surface of feedstock promoting the HTC
reactions. In such a system, the extent of carbonization will be determined by the extent of the wetting
of the feedstock by steam. More detailed study is needed to understand the relationship between the
degree of wetting by steam and carbonization.
Figure 6. Comparison of the process conditions for the seven cases in the T–v diagram for water.
107
Energies 2020, 13, 5733
Furthermore, the diagram helps visualize the safety aspects. It is easy to see that the target
conditions in Cases 1, 2, 5, and 7 are well away from entering the subcooled liquid compression region,
where pressure increases rapidly with an increase in reactor temperature. For the semi-batch system
in Case 3, the overall specific volume vR decreases from 0.005 to 0.0029 m3 /kg due to the increase in
the total mass of water as steam is injected into the reactor, and we move from the right to the left on
the isobaric line at 1.6 MPa (Figure 6). For such semi-batch systems, it is important to make sure the
start and end points remain far enough away from the subcooled compression region. Temperature
increases above the initial target conditions due to use of superheated steam or exothermic reactions
could move the system diagonally upwards towards the subcooled compression region and high
pressure as steam is added. In the subcooled compression region, if a safety rupture disk valve is not
present to release at a preset pressure, the reactor pressure can exceed the tensile strength of reactor
material, and the reactor can explode.
4. Conclusions
There are many types of hydrothermal reactor systems being used with many process variations
in the literature. The analysis presented in this paper illustrates that a large range of conditions
need to be considered before labeling a reactor system VTC and HTC. The analysis of the process
conditions of seven different HTC/VTC cases reported in the literature through the use of the models
developed in this paper and a T-v phase diagram showed that the distinction between HTC and VTC
is often ambiguous. The models developed in this study for predicting pressure, the volume fraction
of liquid water and the distribution of water between phases as a function of reactor temperature
can be used to systematically analyze various HTC/VTC process conditions. Furthermore, this study
also demonstrates the importance of predicting the condition at which the reactor system enters the
subcooled compression liquid region to avoid the danger of explosion. Comparison of the reactor
pressures predicted by the models to the actual pressure for reactors filled with varying amounts
of water with and without initial pressurization showed reasonable agreement. However, higher
pressures can be expected with the addition of feedstock due to the production of CO2 and other
gases by the hydrothermal reactions and the decrease in headspace volume occupied by the feedstock.
In order to describe the amount of liquid water in physical contact with feedstock, we defined a new
solid content parameter %S(T) which changes with reaction temperature due to changes in the water
distribution between phases. This parameter is more useful in describing the solid content than the
nominal parameter %So typically reported in the literature. While the models developed here can help
determine whether steam or liquid water predominates at the specific process conditions, more research
and modeling on hydrothermal systems with feedstock present are required to understand the effect
of the water phase on the hydrothermal reactions. The tools presented here can help in designing
experiments to compare systems and in understanding results in future HTC research.
Author Contributions: K.S.R. conceived the original idea, conducted W-HTC experiments, analyzed the data,
and participated in manuscript writing. J.A.L. conducted experiments using an 18.75-L reactor system, and both
J.A.L. and A.A.-M. participated in all phases of the research, analyzed the data, simulated T-v diagrams,
and participated in manuscript writing. All authors have read and agreed to the published version of
the manuscript.
Funding: Financial help for A.A.M. came from the Junta de Extremadura and FEDER (Fondo Europeo de
Desarrollo Regional “Una manera de hacer Europa”) project IB16108, and also from the program “Ayudas a grupos
de la Junta de Extremadura” GR18150. The open access journal fee was supported by the Leibniz Association’s
Open Access Publishing Fund.
Acknowledgments: The authors would like to acknowledge the technical support provided by Melvin Johnson
of the USDA-Agricultural Research Service (ARS), Coastal Plains Soil, Water and Plant Research Center, Florence,
SC, and Marcus Fischer of the Leibniz Institute of Agricultural Engineering and Bioeconomy. This research was
supported by the USDA-ARS National Program 212 Soil and Air. Mention of trade names or commercial products
is solely for the purpose of providing specific information and does not imply recommendation or endorsement
by the U.S. Department of Agriculture.
Conflicts of Interest: The authors declare no conflict of interest.
108
Energies 2020, 13, 5733
Glossary
Symbol Description Unit
D overall reactor water density, MH2O /VR kg/m3
Mbiomass mass of feedstock in the reactor as dry matter (DM)
MH2O total mass of water as liquid and vapor water in the reactor kg
MC moisture content %
MWH2O molecular weight of water kg/kmol
P reactor pressure MPa
Po initial reactor pressure MPa
Psat saturated vapor pressure -
R universal gas constant (8.31451 × 10-3 m3 -MPa/kmol-K) -
% solid in reactor—ratio of mass of feedstock in DM to (total mass of
%S water + mass of feedstock in DM in reactor), -
Mbiomass /(MH2O + Mbiomass )
T reactor temperature ◦C
References
1. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.M.; Fuhner, C.; Bens, O.;
Kern, J.; et al. Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry,
processes and applications of wet and dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
2. Cao, X.; Ro, K.S.; Chappell, M.; Li, Y.; Mao, J. Chemical structures of swine-manure chars produced under
different carbonization conditions investigated by advanced solid-state 13C nuclear magnetic resonance
(NMR) spectroscopy. Energy Fuels 2011, 25, 388–397. [CrossRef]
3. Ro, K.S.; Flora, J.R.V.; Bae, S.; Libra, J.A.; Berge, N.D.; Álvarez -Murillo, A. Properties of animal-manure-based
hydrochars and predictions using published models. ACS Sustain. Chem. Eng. 2017, 5, 7317–7324. [CrossRef]
4. Roman, S.; Libra, J.A.; Berge, N.D.; Sabio, E.; Ro, K.S.; Li, L.; Ledesma, B.; Alvarez, A.; Bae, S. Hydrothermal
carbonization: Modeling, final properties design and applications: A review. Energies 2018, 11, 216.
[CrossRef]
5. Cao, X.; Ro, K.S.; Libra, J.A.; Kammann, C.; Lima, I.M.; Berge, N.D.; Li, L.; Li, Y.; Chen, N.; Yang, J.; et al.
Effects of biomass types and carbonization conditions on the chemical characteristics of hydrochars. J. Agric.
Food Chem. 2013, 61, 9401–9411. [CrossRef] [PubMed]
6. Ro, K.S.; Novak, J.M.; Johnson, M.G.; Szogi, A.A.; Libra, J.A.; Spokas, K.A.; Bae, S. Leachate water quality of
soils amended with different swine manure-based amendements. Chemosphere 2016, 142, 92–99. [CrossRef]
[PubMed]
7. Minaret, J.; Dutta, A. Comparison of liquid and vapor hydrothermal carbonization of corn husk for the use
as a solid fuel. Bioresour. Technol. 2016, 200, 804–811. [CrossRef] [PubMed]
8. Álvarez-Murillo, A.; Roman, S.; Ledesma, B.; Sabio, E. Study of variables in energy densification of oliver
stone by hydrothermal carbonization. J. Anal. Appl. Pyrolysis 2015, 113, 307–314. [CrossRef]
9. Safril, T.S.; Safril, B.I.; Yoshikawa, K. Commercial demonstration of solid fuel production from municipal
solid waste employing the hydrothermal treatment. Int. J. Environ. Sci. 2017, 2, 316–323. [CrossRef]
109
Energies 2020, 13, 5733
10. Shafie, S.A.; Al-attab, K.A.; Zainal, Z.A. Effect of hydrothermal and vapothermal carbonization of wet
biomass waste on bound moisture removal and combustion characteristics. Appl. Therm. Eng. 2018, 139,
187–195. [CrossRef]
11. Funke, A.; Reebs, F.; Kruse, A. Experimental comparison of hydrothermal and vapothermal carbonization.
Fuel Process. Technol. 2013, 115, 261–269. [CrossRef]
12. Berge, N.D.; Ro, K.S.; Mao, J.D.; Flore, J.R.V.; Chappell, M.A.; Bae, S. Hydrothermal carbonization of
municipal waste streams. Environ. Sci. Technol. 2011, 45, 5696–5703. [CrossRef] [PubMed]
13. Gomez, J.; Corsi, G.; Pino-Cortes, E.; Diaz-Robles, L.A.; Campos, V.; Cubillos, F.; Pelz, S.K.; Paczkowsk, S.;
Carrasco, S.; Silva, J.; et al. Modeling and simulation of a continuous biomass hydrothermal carbonization
process. Chem. Eng. Commun. 2020, 207, 751–768. [CrossRef]
14. Yeoh, K.-H.; Shafie, S.A.; Al-attab, K.A.; Zainal, Z.A. Upgrading agricultural wastes using three different
carbonization methods: Thermal, hydrothermal and vapothermal. Bioresour. Technol. 2018, 265, 365–371.
[CrossRef] [PubMed]
15. Akiya, N.; Savage, P.E. Roles of water for chemical reactions in high-temperature water. Chem. Rev. 2002,
102, 2725–2750. [CrossRef] [PubMed]
16. Lemmon, E.W.; Bell, I.H.; Huber, M.L.; McLinden, M.O. NIST Standard Seference Database 23: Reference
Fluid Thermodynamic and Transport Properties-REFPROP, Version 10.0; National Institute of Standards and
Technology (NIST): Gaithersburg, MD, USA, 2018.
17. Lemmon, E.W.; McLinden, M.O.; Friend, D.G. Thermophysical properties of fluid systems. In NIST Chemistry
Webbook; NIST Standard Reference Database No. 69; National Institute of Standards and Technology (NIST):
Gaithersburg, MD, USA, 2017.
18. IAPWS Thermodynamic Property Formulations. Available online: http://www.iapws.org/newform.html
(accessed on 1 October 2020).
19. Parr Stirred Reactors and Pressure Vessels. Chapter 2. Bulletin 4500, Volume 12. Available online: https://www.
parrinst.com/de/files/downloads/2013/08/4500MB-v12.0_Ch2_Parr_Stirred-Reactors-Literature.pdf (accessed
on 15 May 2020).
20. Parr Safety in the Operation of Laboratory Reactors and Pressure Vessels. No. 230M, 8; Moline, IL. 2009.
Available online: https://www.ccmr.cornell.edu/wp-content/uploads/sites/2/2015/11/ParrReactorSafetyInfo-
230m.pdf (accessed on 15 May 2020).
21. Corder, S.E. Properties and uses of bark as an energy source. In Proceedings of the XVI IUFRO World
Congress, Oslo, Norway, 20 June–2 July 1976; Volume 31.
22. Karpaky, H.; Maalouf, C.; Bliard, C.; Gacoin, A.; Lachi, M.; Polidori, G. Mechanical and thermal
characteriziation of a beet pulp-starch composite for building applications. E3S Web Conf. 2019, 85,
1–8. [CrossRef]
23. Oliveira, S.L.; Mendes, R.F.; Mendes, L.M.; Freire, T.P. Particleboard panels made from sugarcane bagasse:
Characterization for use in the furniture industry. Mater. Res. 2016, 19, 914–922. [CrossRef]
24. Pathak, P.D.; Mandavgane, S.A.; Kulkarni, B.D. Fruit peel waste: Characterization and its potential uses.
Curr. Sci. 2017, 113, 444–454. [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
110
energies
Article
Formation of Carbon Quantum Dots via Hydrothermal
Carbonization: Investigate the Effect of Precursors
Md Rifat Hasan 1 , Nepu Saha 2 , Thomas Quaid 2 and M. Toufiq Reza 2, *
1 Department of Chemical and Biomolecular Engineering, Ohio University, Athens, OH 45701, USA;
mh919116@ohio.edu
2 Department of Biomedical and Chemical Engineering and Sciences, Florida Institute of Technology,
Melbourne, FL 32901, USA; nsaha2019@my.fit.edu (N.S.); tquaid2018@my.fit.edu (T.Q.)
* Correspondence: treza@fit.edu; Tel.: +1-321-674-8578
Abstract: Carbon quantum dots (CQDs) are nanomaterials with a particle size range of 2 to 10 nm.
CQDs have a wide range of applications such as medical diagnostics, bio-imaging, biosensors,
coatings, solar cells, and photocatalysis. Although the effect of various experimental parameters,
such as the synthesis method, reaction time, etc., have been investigated, the effect of different
feedstocks on CQDs has not been studied yet. In this study, CQDs were synthesized from hydrox-
ymethylfurfural, furfural, and microcrystalline cellulose via hydrothermal carbonization at 220 ◦ C
for 30 min of residence time. The produced CQDs showed green luminescence behavior under
the short-wavelength UV light. Furthermore, the optical properties of CQDs were investigated
using ultraviolet-visible spectroscopy and emission spectrophotometer, while the morphology and
chemical bonds of CQDs were investigated using transmission electron microscopy and Fourier-
transform infrared spectroscopy, respectively. Results showed that all CQDs produced from various
precursors have absorption and emission properties but these optical properties are highly dependent
Citation: Hasan, M.R.; Saha, N.;
on the type of precursor. For instance, the mean particle sizes were 6.36 ± 0.54, 5.35 ± 0.56, and
Quaid, T.; Reza, M.T. Formation of 3.94 ± 0.60 nm for the synthesized CQDs from microcrystalline cellulose, hydroxymethylfurfural,
Carbon Quantum Dots via and furfural, respectively, which appeared to have similar trends in emission intensities. In addition,
Hydrothermal Carbonization: the synthesized CQDs experienced different functionality (e.g., C=O, O-H, C-O) resulting in different
Investigate the Effect of Precursors. absorption behavior.
Energies 2021, 14, 986.
https://doi.org/10.3390/en14040986 Keywords: carbon quantum dots; hydrothermal carbonization; hydroxymethylfurfural; furfural;
microcrystalline cellulose
Academic Editor: David Chiaramonti
water is more reactive and behaves as a non-polar solvent because of high ionic product
and low dielectric constant [28]. Therefore, carbohydrates undergo for hydrolysis and
the hydrolyzed products then undergo for simultaneous dehydration, decarboxylation,
condensation, and polymerization to make cross-linked polymeric materials [29]. The
particle sizes of the solid product, often referred as hydrohcar, are generally between 10 nm
to 100 μm [30]. As HTC is a bottom-up approach, the particles smaller than 10 nm remain
in the liquid phase (known as process liquid). HTC process liquid is referred to as the
waste product from a HTC process; thus far, it is considered as a liability for the HTC
development, as it requires expensive treatment. Various treatment technologies including
anaerobic digestion (AD), wet air oxidation (WAO), and membrane distillation (MD) have
been proposed to treat HTC process liquid [31–33]. However, the lack of value-added
product separation along with the additional cost of treatment often prohibits the adoption
of these aforementioned technologies for HTC process liquid.
In the recent past, various researchers have tried to separate the CQDs from the
liquid phase of the HTC process [3,4,34–36]. For instance, Mehta et al. produced highly
fluorescent CQDs from sugarcane via hydrothermal treatment at 120 ◦ C for 3 h [4]. The
obtained the CQDs were about 3.0 nm in size with highly blue fluorescence. On the other
hand, Sahu et al. reported highly photoluminescent CQDs with sizes 1.5 to 4.5 nm from
orange juice via HTC at 120 ◦ C for 2.5 h [34]. Papaioannou et al. studied the effect of
HTC residence time from 2 to 12 h at 200 ◦ C on the properties of CQDs produced from
D -(+)-glucose [35]. They reported that the sizes of CQDs decreased and their level of
crystallinity increased with an increase in reaction time.
From the above discussed literature, it is clear that various parameters such as reaction
time, residence time, etc. have been examined to understand their effect on the properties of
CQDs. Although researchers studied different feedstocks at various conditions, to the best
of the authors’ knowledge, no study has been reported about how the properties of CQDs
change with the variation of feedstock at the same experimental conditions (temperature
and residence time). The hypothesis was that simultaneous evolutions of amorphous
hydrochar and semi-crystalline CQDs occur during HTC reactions. The amorphous hy-
drochar agglomerates into micro-meter-sized supramolecules, whereas, nano-sized CQDs
remain in the liquid phase due to their high aqueous solubility. As various biopolymers
react differently under HTC conditions, the authors’ expectation was that the presence
and properties of CQDs will be different as well. Therefore, the objectives of the project
were to investigate the effect of feedstock on the properties of CQDs. To achieve this
goal, HTC of three organic precursors (i.e., furfural, 5-hydroxymethyl furfural (HMF),
and microcrystalline cellulose) have been performed at 220 ◦ C for 30 min of residence
time. In this study, the HTC experiments were conducted at a relatively low temperature
and residence time as they are reported to be favorable for CQD production. HMF and
furfural might react at temperatures lower than 220 ◦ C under an HTC environment, but the
literature indicates that microcrystalline cellulose starts to react at around 220 ◦ C [29,37].
As the purpose of this study was to investigate the effect of precursor, the authors wanted
to choose the lowest temperature where all the precursors could react under the HTC
environment. Multi-staged filtration was performed on the process liquids to remove
supramolecules. The presence of CQDs was confirmed by investigating the optical (lumi-
nescence, ultraviolet-visible absorption, and emission), morphology (transmission electron
microscopy), and chemical (Fourier Transform Infrared Spectroscopy) properties.
112
Energies 2021, 14, 986
purchased from Acros Organics (Fair Lawn, NJ, USA). Analytical-grade ethanol was
purchased from Fisher Scientific (Waltham, MA, USA).
2.2. Methods
2.2.1. Hydrothermal Carbonization and Separation of CQDs from HTC Process Liquid
HTC experiments of different precursors were performed in a glass-lined 100 mL Parr
batch reactor (reactor series 4590, Moline, IL, USA). A Parr proportional-integral-derivative
(PID) controller (model 4590) with an accuracy of ±2 ◦ C was used to control the reaction
temperature. The pressure was not controlled but was monitored during the experiment
with a pressure transducer and a gauge. HTC experiments were conducted at 220 ◦ C for
a residence time of 30 min. The reactor was loaded with 45 mL solution contained 10%
(w/v) of precursor and the rest was deionized (DI) water. The reactor was closed and
heated at a constant rate of 10 ◦ C/min until it reached the set temperature and then it
was maintained at isothermal conditions for 30 min. At the end of the residence time, the
heater was turned off, heating elements were removed from the reactor, and the reactor
was rapidly cooled to room temperature (~30 ◦ C) by placing it in an ice-water bath. Once
the reactor reached room temperature, the gaseous products produced during the reaction
were vented in a fume hood by opening the vent valve. Finally, the lid of the reactor was
opened and HTC process liquid was filtered by Whatman 41 filter paper. Dark brown HTC
process liquid samples were collected in centrifuge tubes and stored in the refrigerator
for further synthesis. As the major goal of this study was to evaluate the variation of
optical, morphological, and chemical properties of CQDs with precursors, the hydrochar
and gaseous products were not further characterized. However, the mass balance and
physico-chemical characterization of hydrochars can be found elsewhere [23]. All the
experiments were completed in duplicate to check reproducibility.
The dark brown HTC process liquid was centrifuged at 10,000 rpm for 15 min by
a Sorvall BIOS 16-series centrifuge from Thermo Fisher Scientific (Waltham, MA, USA)
to separate the larger particles from the liquid phase. The fluorescent CQD containing
liquid phase was then filtered with a standard syringe filter (0.22 μm). The filtered CQD
containing solution was then evaporated under reduced pressure by vacuum distillation.
The concentrated product from vacuum distillation was kept in the high vacuum freeze
dryer for 24 h to obtain powdered CQDs. The powdered CQDs were collected and
refrigerated in glass vials for further investigations. A schematic of the CQD synthesis and
characterization strategy is shown in Figure 1.
DI Water
Uv-Vis Absorbance
Centrifuged at 10,000 Vacuum distillation
Filtered out the rpm for 15 minutes Freeze dry
solids Emission
Emission
HTC process CQD containing Spectrometer
Powdered CQD
liquid liquid phase
10 % (w/v) TEM Morphology
precursor Separate the
larger particles
113
Energies 2021, 14, 986
114
Energies 2021, 14, 986
observed at 1670 cm−1 . On the other hand, both HMF and furfural showed various other
peaks, such as medium alcohol (O-H) peaks between 1330 and 1420 cm−1 , aromatic ester
(C-O) peaks between 1200 and 1300 cm−1 , and a sharp alkyl ether (C-O) peak at 1027 cm−1 .
Due to the presence of additional functionality, HMF and furfural could exhibit broader
absorbance compared to cellulose. In addition to the broad peaks, CQDs showed tails in
the visible region. These tails are typically related to nanoparticle functionalization and are
reported as lower energy surface centers [38]. These tails are also attributed the presence of
various π→π* (C=C) and n→π* (C=O and/or others) transitions [2]. As the aromatic rings
increase with the hydrothermal treatment, the energy gaps between π states gradually
decreased [15]. On the other hand, functional groups (i.e., carbonyl) with electron lone pairs
could be bonded with aromatic carbon that allows electron transition from n states [39].
2.5
Cellulose
HMF
2.0
Furfural
Absorbance
1.5
1.0
0.5
0.0
200 300 400 500 600 700
Wavenlength (nm)
Figure 2. UV-Vis spectra for carbon quantum dots derived from various precursors.
Aromatic ester/ether @1200−1300
Conjugated ketone @1666−1680
Transmission
Cellulose
Alcohol @1330−1420
Furfural
HMF
3000 2800 2600 2400 2200 2000 1800 1600 1400 1200 1000
Wavenumber (cm-1)
Figure 3. FTIR spectra of the CQDs synthesized from various precursors.
115
Energies 2021, 14, 986
be seen that the emission wavelength (360 nm) is longer than the excitation wavelength
(240 nm) shown in the absorbance spectra (Figure 2). This observation is in line with previous
literature, where the researchers reported the reason behind this is mainly due to bandgap
of the conjugated π domains and/or the presence of defects in the structures [39–41]. The
spectra also show that the emission capacity of cellulose-based CQDs is the highest among
the studied feedstocks, where furfural-derived CQDs show the least emission capacities. As
the absorbance intensity of the furfural-derived CQDs was the least, it was expected to have
the highest emission properties. However, the emission properties showed the similar trend
as absorption capacity. This could be due to the combined effect of functional groups, ligand
chain length, surface defects, and morphology of the CQDs [42].
ϴϬϬ͕ϬϬϬ
ĞůůƵůŽƐĞ
ϳϬϬ͕ϬϬϬ
,D&
ϲϬϬ͕ϬϬϬ &ƵƌĨƵƌĂů
;ĐŽƵŶƚƐƉĞƌƐĞĐŽŶĚͿ
ŵŝƐƐŝŽŶ/ŶƚĞŶƐŝƚLJ
ϱϬϬ͕ϬϬϬ ,ϮKďĂĐŬŐƌŽƵŶĚ
ϰϬϬ͕ϬϬϬ
ϯϬϬ͕ϬϬϬ
ϮϬϬ͕ϬϬϬ
ϭϬϬ͕ϬϬϬ
Ϭ
ϯϲϬ ϰϭϬ ϰϲϬ ϱϭϬ ϱϲϬ ϲϭϬ ϲϲϬ
tĂǀĞůĞŶŐƚŚ;ŶŵͿ
Figure 4. Emission spectra of carbon quantum dots derived from various precursors.
116
Energies 2021, 14, 986
(A)
(B)
(C)
Figure 5. TEM size images along with the particle size distribution of the different types of CQDs synthesized from:
(A) HMF; (B) Furfural; (C) Cellulose.
117
Energies 2021, 14, 986
4. Conclusions
In this study, CQDs were produced from HMF, furfural, and microcrystalline cellulose
via HTC at 220 ◦ C for 30 min. The optical and morphological properties of the produced
CQDs were investigated. It was observed that CQDs showed green luminescence under
short-wavelength UV light (254 nm). Additionally, they can absorb light in a broad range.
The variation of the absorption range was justified from the FTIR spectra where CQDs
from different precursors showed different chemical bonds (e.g., alcohol, ketone, etc.).
Although all the studied CQDs were within two to nine nm in diameter, the mean diameter
varied with the precursor. Depending on the particle size, the CQDs showed different
emission capacities. Finally, this study provides an insight into the effect of precursor on
the CQD properties. So, the precursors for the CQDs should be selected based on the
targeted applications.
Author Contributions: Conceptualization, M.R.H., N.S., M.T.R.; methodology, M.R.H., N.S.; formal
analysis, M.R.H., N.S., T.Q.; investigation, M.R.H., N.S., M.T.R.; writing—original draft preparation,
M.R.H., N.S., T.Q.; writing—review and editing, M.R.H., N.S., T.Q., M.T.R. All authors have read and
agreed to the published version of the manuscript.
Funding: The research is partially funded by the United States Department of Agriculture (USDA)
grant (2019-67019-2928) through M.T.R.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: The authors acknowledge Travis A. White from the Department of Chemistry
and Biochemistry at Ohio University for allowing to use emission spectrofluorometer. The authors
are also thankful to Kyle McGaughy from the biofuels lab at the Florida Institute of Technology for
his laboratory efforts in this project.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Liu, W.; Li, C.; Ren, Y.; Sun, X.; Pan, W.; Li, Y.; Wang, J.; Wang, W. Carbon dots: Surface engineering and applications. J. Mater.
Chem. B 2016, 4, 5772–5788. [CrossRef]
2. Wang, Y.; Hu, A. Carbon quantum dots: Synthesis, properties and applications. J. Mater. Chem. C 2014, 2, 6921–6939. [CrossRef]
3. Zhu, S.; Meng, Q.; Wang, L.; Zhang, J.; Song, Y.; Jin, H.; Zhang, K.; Sun, H.; Wang, H.; Yang, B. Highly photoluminescent carbon
dots for multicolor patterning, sensors, and bioimaging. Angew. Chem. 2013, 125, 4045–4049. [CrossRef]
4. Mehta, V.N.; Jha, S.; Kailasa, S.K. One-pot green synthesis of carbon dots by using Saccharum officinarum juice for fluorescent
imaging of bacteria (Escherichia coli) and yeast (Saccharomyces cerevisiae) cells. Mater. Sci. Eng. C 2014, 38, 20–27. [CrossRef]
[PubMed]
5. Qu, D.; Zheng, M.; Du, P.; Zhou, Y.; Zhang, L.; Li, D.; Tan, H.; Zhao, Z.; Xie, Z.; Sun, Z. Highly luminescent S, N co-doped
graphene quantum dots with broad visible absorption bands for visible light photocatalysts. Nanoscale 2013, 5, 12272–12277.
[CrossRef] [PubMed]
6. Choudhary, R.; Patra, S.; Madhuri, R.; Sharma, P.K. Equipment-free, single-step, rapid, “on-site” kit for visual detection of lead
ions in soil, water, bacteria, live cells, and solid fruits using fluorescent cube-shaped nitrogen-doped carbon dots. ACS Sustain.
Chem. Eng. 2016, 4, 5606–5617. [CrossRef]
7. Gao, X.; Lu, Y.; Zhang, R.; He, S.; Ju, J.; Liu, M.; Li, L.; Chen, W. One-pot synthesis of carbon nanodots for fluorescence turn-on
detection of Ag+ based on the Ag+-induced enhancement of fluorescence. J. Mater. Chem. C 2015, 3, 2302–2309. [CrossRef]
8. Kim, S.; Choi, Y.; Park, G.; Won, C.; Park, Y.-J.; Lee, Y.; Kim, B.-S.; Min, D.-H. Highly efficient gene silencing and bioimaging based
on fluorescent carbon dots in vitro and in vivo. Nano Res. 2017, 10, 503–519. [CrossRef]
9. Li, H.; Kang, Z.; Liu, Y.; Lee, S.-T. Carbon nanodots: Synthesis, properties and applications. J. Mater. Chem. 2012, 22, 24230–24253.
[CrossRef]
10. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst
for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [CrossRef] [PubMed]
118
Energies 2021, 14, 986
11. Loukanov, A.; Sekiya, R.; Yoshikawa, M.; Kobayashi, N.; Moriyasu, Y.; Nakabayashi, S. Photosensitizer-conjugated ultrasmall
carbon nanodots as multifunctional fluorescent probes for bioimaging. J. Phys. Chem. C 2016, 120, 15867–15874. [CrossRef]
12. Wang, Z.; Fu, B.; Zou, S.; Duan, B.; Chang, C.; Yang, B.; Zhou, X.; Zhang, L. Facile construction of carbon dots via acid catalytic
hydrothermal method and their application for target imaging of cancer cells. Nano Res. 2016, 9, 214–223. [CrossRef]
13. Wu, Y.-F.; Wu, H.-C.; Kuan, C.-H.; Lin, C.-J.; Wang, L.-W.; Chang, C.-W.; Wang, T.-W. Multi-functionalized carbon dots as
theranostic nanoagent for gene delivery in lung cancer therapy. Sci. Rep. 2016, 6, 21170. [CrossRef] [PubMed]
14. Zhou, J.; Booker, C.; Li, R.; Zhou, X.; Sham, T.-K.; Sun, X.; Ding, Z. An electrochemical avenue to blue luminescent nanocrystals
from multiwalled carbon nanotubes (MWCNTs). J. Am. Chem. Soc. 2007, 129, 744–745. [CrossRef] [PubMed]
15. Li, H.; He, X.; Kang, Z.; Huang, H.; Liu, Y.; Liu, J.; Lian, S.; Tsang, C.H.A.; Yang, X.; Lee, S.T. Water-soluble fluorescent carbon
quantum dots and photocatalyst design. Angew. Chem. 2010, 122, 4532–4536. [CrossRef]
16. Qiao, Z.-A.; Wang, Y.; Gao, Y.; Li, H.; Dai, T.; Liu, Y.; Huo, Q. Commercially activated carbon as the source for producing
multicolor photoluminescent carbon dots by chemical oxidation. Chem. Commun. 2010, 46, 8812–8814. [CrossRef]
17. Park, S.Y.; Lee, H.U.; Park, E.S.; Lee, S.C.; Lee, J.-W.; Jeong, S.W.; Kim, C.H.; Lee, Y.-C.; Huh, Y.S.; Lee, J. Photoluminescent green
carbon nanodots from food-waste-derived sources: Large-scale synthesis, properties, and biomedical applications. ACS Appl.
Mater. Interfaces 2014, 6, 3365–3370. [CrossRef]
18. Tang, L.; Ji, R.; Cao, X.; Lin, J.; Jiang, H.; Li, X.; Teng, K.S.; Luk, C.M.; Zeng, S.; Hao, J. Deep ultraviolet photoluminescence of
water-soluble self-passivated graphene quantum dots. ACS Nano 2012, 6, 5102–5110. [CrossRef] [PubMed]
19. Chen, B.; Li, F.; Li, S.; Weng, W.; Guo, H.; Guo, T.; Zhang, X.; Chen, Y.; Huang, T.; Hong, X. Large scale synthesis of photolumines-
cent carbon nanodots and their application for bioimaging. Nanoscale 2013, 5, 1967–1971. [CrossRef] [PubMed]
20. Wang, W.; Ni, Y.; Xu, Z. One-step uniformly hybrid carbon quantum dots with high-reactive TiO2 for photocatalytic application.
J. Alloy. Compd. 2015, 622, 303–308. [CrossRef]
21. Bian, J.; Huang, C.; Wang, L.; Hung, T.; Daoud, W.A.; Zhang, R. Carbon dot loading and TiO2 nanorod length dependence
of photoelectrochemical properties in carbon dot/TiO2 nanorod array nanocomposites. ACS Appl. Mater. Interfaces 2014,
6, 4883–4890. [CrossRef]
22. Wang, J.; Gao, M.; Ho, G.W. Bidentate-complex-derived TiO2 /carbon dot photocatalysts: In situ synthesis, versatile heterostruc-
tures, and enhanced H 2 evolution. J. Mater. Chem. A 2014, 2, 5703–5709. [CrossRef]
23. Saha, N.; Saba, A.; Reza, M.T. Effect of hydrothermal carbonization temperature on pH, dissociation constants, and acidic
functional groups on hydrochar from cellulose and wood. J. Anal. Appl. Pyrolysis 2019, 137, 138–145. [CrossRef]
24. Saha, N.; Volpe, M.; Fiori, L.; Volpe, R.; Messineo, A.; Reza, M.T. Cationic Dye Adsorption on Hydrochars of Winery and Citrus
Juice Industries Residues: Performance, Mechanism, and Thermodynamics. Energies 2020, 13, 4686. [CrossRef]
25. Saha, N.; Saba, A.; Saha, P.; McGaughy, K.; Franqui-Villanueva, D.; Orts, W.J.; Hart-Cooper, W.M.; Reza, M.T. Hydrothermal
carbonization of various paper mill sludges: An observation of solid fuel properties. Energies 2019, 12, 858. [CrossRef]
26. Reza, M.T. Upgrading biomass by hydrothermal and chemical conditioning. Ph.D. Dissertation, University of Nevada, Reno, NV,
USA, 2013.
27. Berge, N.D.; Ro, K.S.; Mao, J.; Flora, J.R.V.; Chappell, M.A.; Bae, S. Hydrothermal Carbonization of Municipal Waste Streams.
Environ. Sci. Technol. 2011, 45, 5696–5703. [CrossRef] [PubMed]
28. Bandura, A.V.; Lvov, S.N. The ionization constant of water over wide ranges of temperature and density. J. Phys. Chem. Ref. Data
2006, 35, 15–30. [CrossRef]
29. Sevilla, M.; Fuertes, A.B. The production of carbon materials by hydrothermal carbonization of cellulose. Carbon 2009,
47, 2281–2289. [CrossRef]
30. Sevilla, M.; Fuertes, A.B. Chemical and structural properties of carbonaceous products obtained by hydrothermal carbonization
of saccharides. Chem.–A Eur. J. 2009, 15, 4195–4203. [CrossRef] [PubMed]
31. Reza, M.T.; Freitas, A.; Yang, X.; Coronella, C.J. Wet air oxidation of hydrothermal carbonization (HTC) process liquid. ACS Sustain.
Chem. Eng. 2016, 4, 3250–3254. [CrossRef]
32. Silva, N.A. Evaluation of Membrane Distillation for Treating Hydrothermal Carbonization Aqueous Product (HAP) from Dairy
Manure. Master’s Thesis, University of Nevada, Reno, NV, USA, 2020.
33. Wirth, B.; Reza, T.; Mumme, J. Influence of digestion temperature and organic loading rate on the continuous anaerobic treatment
of process liquor from hydrothermal carbonization of sewage sludge. Bioresour. Technol. 2015, 198, 215–222. [CrossRef]
34. Sahu, S.; Behera, B.; Maiti, T.K.; Mohapatra, S. Simple one-step synthesis of highly luminescent carbon dots from orange juice:
Application as excellent bio-imaging agents. Chem. Commun. 2012, 48, 8835–8837. [CrossRef]
35. Papaioannou, N.; Titirici, M.-M.; Sapelkin, A. Investigating the Effect of Reaction Time on Carbon Dot Formation, Structure, and
Optical Properties. ACS Omega 2019, 4, 21658–21665. [CrossRef]
36. Yang, Z.-C.; Wang, M.; Yong, A.M.; Wong, S.Y.; Zhang, X.-H.; Tan, H.; Chang, A.Y.; Li, X.; Wang, J. Intrinsically fluorescent
carbon dots with tunable emission derived from hydrothermal treatment of glucose in the presence of monopotassium phosphate.
Chem. Commun. 2011, 47, 11615–11617. [CrossRef]
37. Saha, N.; McGaughy, K.; Reza, M.T. Elucidating hydrochar morphology and oxygen functionality change with hydrothermal
treatment temperature ranging from subcritical to supercritical conditions. J. Anal. Appl. Pyrolysis 2020, 152, 104965. [CrossRef]
119
Energies 2021, 14, 986
38. Carbonaro, C.M.; Corpino, R.; Salis, M.; Mocci, F.; Thakkar, S.V.; Olla, C.; Ricci, P.C. On the emission properties of carbon dots:
Reviewing data and discussing models. C—J. Carbon Res. 2019, 5, 60. [CrossRef]
39. Wang, R.; Lu, K.-Q.; Tang, Z.-R.; Xu, Y.-J. Recent progress in carbon quantum dots: Synthesis, properties and applications in
photocatalysis. J. Mater. Chem. A 2017, 5, 3717–3734. [CrossRef]
40. Baker, S.N.; Baker, G.A. Luminescent carbon nanodots: Emergent nanolights. Angew. Chem. Int. Ed. 2010, 49, 6726–6744.
[CrossRef] [PubMed]
41. Gokus, T.; Nair, R.; Bonetti, A.; Bohmler, M.; Lombardo, A.; Novoselov, K.; Geim, A.; Ferrari, A.; Hartschuh, A. Making graphene
luminescent by oxygen plasma treatment. Acs Nano 2009, 3, 3963–3968. [CrossRef] [PubMed]
42. Spirin, M.G.; Brichkin, S.B.; Gak, V.Y.; Razumov, V.F. Influence of photoactivation on luminescent properties of colloidal InP@ZnS
quantum dots. J. Lumin. 2020, 226, 117297. [CrossRef]
43. Zhang, Y.; Wang, Y.; Feng, X.; Zhang, F.; Yang, Y.; Liu, X. Effect of reaction temperature on structure and fluorescence properties
of nitrogen-doped carbon dots. Appl. Surf. Sci. 2016, 387, 1236–1246. [CrossRef]
44. Wongso, V.; Sambudi, N.S.; Sufian, S. The effect of hydrothermal conditions on photoluminescence properties of rice husk-derived
silica-carbon quantum dots for methylene blue degradation. Biomass Convers. Biorefin. 2020, 179. [CrossRef]
45. Zhao, S.; Song, X.; Chai, X.; Zhao, P.; He, H.; Liu, Z. Green production of fluorescent carbon quantum dots based on pine wood
and its application in the detection of Fe3+ . J. Clean. Prod. 2020, 263, 121561. [CrossRef]
46. Gao, X.; Zhou, X.; Ma, Y.; Qian, T.; Wang, C.; Chu, F. Facile and cost-effective preparation of carbon quantum dots for Fe3+ ion
and ascorbic acid detection in living cells based on the “on-off-on” fluorescence principle. Appl. Surf. Sci. 2019, 469, 911–916.
[CrossRef]
120
energies
Article
Integration of Air Classification and Hydrothermal
Carbonization to Enhance Energy Recovery of Corn Stover
Md Tahmid Islam 1 , Nepu Saha 1 , Sergio Hernandez 2 , Jordan Klinger 2 and M. Toufiq Reza 1, *
1 Department of Biomedical and Chemical Engineering and Sciences, Florida Institute of Technology,
150 W University Boulevard, Melbourne, FL 32901, USA; islamm2019@my.fit.edu (M.T.I.);
nsaha2019@my.fit.edu (N.S.)
2 Biomass Characterization Department, Idaho National Laboratory, 2525 Fremont Ave, Idaho Falls,
Idaho, ID 83402, USA; sergio.hernandez@inl.gov (S.H.); jordan.klinger@inl.gov (J.K.)
* Correspondence: treza@fit.edu; Tel.: +1-321-674-8578
Abstract: Air classification (AC) is a cost-effective technology that separates the energy-dense light
ash fraction (LAF) from the inorganic-rich high ash fraction (HAF) of corn stover. HAF could be
upgraded into energy-dense solid fuel by hydrothermal carbonization (HTC). However, HTC is a
high-temperature, high-pressure process, which requires additional energy to operate. In this study,
three different scenarios (i.e., AC only, HTC only, and integrated AC–HTC) were investigated for
the energy recovery of corn stover. AC was performed on corn stover at an 8 Hz fan speed, which
yielded 84.4 wt. % LAF, 12.8 wt. % HAF, and 2.8 wt. % below screen particles. About 27 wt. % ash
was reduced from LAF by the AC process. Furthermore, HTC was performed on raw corn stover
and the HAF of corn stover at 200, 230, and 260 ◦ C for 30 min. To evaluate energy recovery, solid
products were characterized in terms of mass yield, ash yield, ultimate analysis, proximate analyses,
and higher heating value (HHV). The results showed that the energy density was increased with
Citation: Islam, M.T.; Saha, N.;
the increase in HTC temperature, meanwhile the mass yield and ash yield were decreased with the
Hernandez, S.; Klinger, J.; Reza, M.T. increase in HTC temperature. Proximate analysis showed that fixed carbon increased 18 wt. % for
Integration of Air Classification and original char and 27 wt. % for HAF char at 260 ◦ C, compared to their respective feedstocks. Finally,
Hydrothermal Carbonization to the hydrochar resulting from HAF was mixed with LAF and pelletized at 180 bar and 90 ◦ C to densify
Enhance Energy Recovery of Corn the energy content. An energy balance of the integrated AC–HTC process was performed, and the
Stover. Energies 2021, 14, 1397. results shows that integrated AC with HTC performed at 230 ◦ C resulted in an additional 800 MJ/ton
https://doi.org/10.3390/en14051397 of energy recovery compared to the AC-only scenario.
Academic Editor: Byong-Hun Jeon Keywords: corn stover; air classification; hydrothermal carbonization; pelletization; energy recovery
and foreign materials are mostly dense materials compared to the lignocellulosic structure
of biomass. Therefore, AC separates a significant amount of these exogenous inorganics
and creates an enriched stream of biomass (often called the light ash fraction, LAF) and a
soil-laden reject stream (also known as the high ash fraction, HAF) [9–11]. Lacey et al. [11]
showed such separations for chipped forest residues after AC and reported remarkable
ash (wt. %) reductions (>32 wt. %) in the final throughputs. Emerson et al. [9] also found
quite reasonable reductions in ash (wt. %) for hybrid poplar (from 2.34 wt. % to 1.67 wt. %)
and shrub willow (from 2.60 wt. % to 2.14 wt. %). Recently, Thompson et al. [2] showed
that AC can effectively upgrade CS by removing 30 wt. % ash at a mild 7.5 Hz fan speed,
or approximately 2.1 m/s counter-stream air flow of AC. It has also been reported that a
large fraction of HAF byproduct is produced during AC, accounting for more than 12 wt.
% of the initial CS. This HAF of CS contains more than 80 wt. % organic content that can be
further utilized [2]. Instead, an estimated HAF disposal cost of USD 28.86 per ton has been
suggested by Reza et al. [12] and Humbird et al. [13], accounting for 2.5 cents of the USD
2.15 per gallon minimum ethanol selling price. In order to make AC more economically
viable, the further utilization of HAF needs to be implemented in conjunction with AC
technologies and system-wide economics and sustainability for full material utilization.
However, to the best of the authors’ knowledge, no work has been done to date to utilize
the waste energy in value-added fuel. Therefore, to make use of the vital unrecovered
energy from this solid waste and integrate it with the AC technology, the HAF could be
converted into essential products.
Hydrothermal carbonization (HTC) is a promising thermochemical process that trans-
forms biomass, such as CS, into a carbon-rich solid product called hydrochar [5,14,15]. The
reaction temperature varies between 180 and 260 ◦ C, and the pressures are maintained
above the saturation pressure. The ionic products of water increase three orders of magni-
tude (KH2O (573K) /KH2O (298K) = 103 ) [5,15,16] and the dielectric constant of water reduces
from 78.5 (298 K, 0.1 MPa) to 27.1 (523 K, 5 MPa) [15,17]. As a result, biomass undergoes a
series of reactions, i.e., hydrolysis, decarboxylation, dehydration, condensation polymeriza-
tion, and aromatization, in subcritical water due to the increased reactivity and solvent-like
properties [18,19]. For agricultural residues, the HTC reaction products can be divided into
three streams: 41–90 wt. % solids with 80–95 wt. % of the original calorific value; 2–10 wt.
% gas consisting mainly of CO2 ; the rest is process liquid [20–24]. The energy-dense solid
hydrochar can be pelletized to improve the mass and energy density of the feedstock and
reduce the cost of transportation, handling and storage [25–27].
During HTC, the low density and low viscosity of subcritical water, in combination
with the acidic process liquid and the modification of the biomass structure, aid in the
leaching of inorganics from biomass structures [18,28–31]. The reduction of ash is partic-
ularly important for hydrochars’ application as fuel, as this can significantly reduce the
ash-fusion temperatures, leading to slagging in fuel boilers [1]. Therefore, HTC could be
used to recover the energy in the HAF stream while reducing ash from the hydrochar.
However, HTC is an energy-intensive process as it requires high temperatures and high
pressure. Therefore, an energy balance in the integrated AC–HTC is warranted to evalu-
ate the net energy recovery from the integrated process compared to the individual AC
and HTC processes. Therefore, the main goal of this work was to study the feasibility of
three different scenarios, to determine the energy recovery from CS: 1. AC-only process;
2. HTC-only process; 3. Integrated AC–HTC process. First, the AC of CS was performed
to separate LAF and HAF. Then HTC on HAF was performed at various temperatures
and the hydrochars were pelletized with LAF. The energy recovery from the integrated
AC–HTC process was compared with AC-only and HTC-only, wherein the original (ORG)
CS was used for both technologies.
122
Energies 2021, 14, 1397
2.2. Methods
2.2.1. Air Classification of Corn Stover
The ORG CS intended for HTC were sampled using mechanical sampling procedures
and equipment according to Solid Biofuels Sampling ISO 18135:2017. A super sack of the
deconstructed CS bale was divided using a custom rotary splitter that consists of a conveyor
and eight bins mounted on a rotating table. In this unit, a feed hopper accommodates
approximately 120 L of sample and a live-bottom style belt feeder slowly dispenses the
samples to 8 sample bins below. The 45 L bins rotate over 100 times during each splitting
operation and make at least one full rotation for each belt flight to ensure the samples
are representative of the original bulk solid. Both the belt feeder and the rotary table are
equipped with speed control devices (variable frequency drive and/or DC potentiometer)
to adjust the processing parameters according to the sample feed behavior to ensure
analytical splitting (an image of AC can be found in the supplementary information
(Figure S1)). AC separates the samples into two fractions as the material is passed over a
screen-covered fan: LAF and HAF. The HAF is blown upward and removed while the LAF
(air classified material) remains. Air classification was performed on CS feedstocks using
a 2× Air Cleaner equipped with an Iso-flo dewatering infeed shaker (Key Technologies,
Walla Walla, WA, USA). CS was air classified using a fan speed of 1.8 m/s. Both light and
heavy fractions were collected after classification for downstream processing and analyses.
123
Energies 2021, 14, 1397
labeled according to the HTC temperature. For instance, CS ORG HTC200 represents the
hydrochar produced from ORG CS at 200 ◦ C.
2.3.2. Ash
The ASTM D1102 method was followed to determine the ash content of the dry solid
samples by using a muffle furnace (Thermo Scientific, Model # FB1415M, Waltham, MA) at
575 ◦ C for 5 h 30 min. The ash yield (AY) was calculated by using Equation (2). Triplicates
were performed for each sample to report reproducibility.
124
Energies 2021, 14, 1397
and vanadium oxide (V2 O5 ) as a conditioner for the samples, which were combusted
around 950 ◦ C in ultra-high purity oxygen with helium carrier gas and passed over copper
oxide pellets and then electrolytic copper. The produced gases were then analyzed by a
thermal conductivity detector (TCD), with the peak areas of detection being compared to
those of BBOT standards. The oxygen (O) content was found by subtraction method. The
following equation was used to find the oxygen wt. %.
O (wt. %) = 100 wt. % − C (wt. %) − H (wt. %) − N (wt. %) − S (wt. %) − ash (wt. %) (5)
Figure 1. Photo of the original (ORG) prepared corn stover (CS) and the air classified separated CS with high ash fraction
(HAF) and light ash fraction (LAF) divisions.
125
Energies 2021, 14, 1397
Table 1. Proximate and ultimate analysis of ORG and separated CS. Measurements made using
ASTM D7582, ASTM D5373, and ASTM D4239.
126
Table 2. Proximate and ultimate analysis of air classified CS hydrochars. DB is dry basis.
HAF HTC 200 62.2 ± 1.9 68.9 ± 2.8 18.6 ± 0.2 64.1 ± 3.8 2.3 ± 0.1 10.6 ± 0.6 69.5 ± 0.4 17.6 ± 0.3 43.9 ± 0.2 5.1 ± 0.1 0.9 ± 0.2 BD** 39.4 ± 0.2
64.9 ±
HTC 230 76.7 ± 2.6 19.9 ± 0.1 67.1 ± 1.3 1.9 ± 0.4 10.6 ± 0.2 66.9 ± 1.5 20.5 ± 1.9 47.5 ± 1.9 4.7 ± 0.3 1.1 ± 0.2 BD** 35.9 ± 1.9
1.9 *
50.1 ±
HTC 260 66.4 ± 0.4 22.2 ± 0.2 69.7 ± 1.7 1.1 ± 0.2 14.3 ± 0.4 40.3 ± 0.2 44.4 ± 0.0 51.1 ± 0.5 3.5 ± 0.1 1.7 ± 0.2 BD** 29.4 ± 0.5
0.4 *
* values refer to duplicate experiments. All other experiments were triplicated; ** below detection limit.
127
Energies 2021, 14, 1397
The AY showed slight changes in their values (Table 2) for HAF hydrochars, but
these changes were significant for ORG hydrochars. AY showed remarkable reductions
of ~50 wt. % and ~36 wt. % from raw ORG to ORG HTC200 and raw HAF to HAF
HTC200, respectively, but it increased for successive HTC temperatures. Qadi et al. [40]
and Chen et al. [41] found that HTC aids in the removal of loose minerals from the biomass.
One possible explanation is that CS can be accompanied by large amounts of loose dirt
during harvesting, depending on method, and this portion was removed significantly
during the HTC200 process. It is likely that this agronomic practice contributed to the
trapped soil in this sample, as discussed on the previous section. On the other hand,
the ORG HTC260 and HAF HTC260 showed the highest AY values of ~86 wt. % and
~70 wt. %, respectively. This could be due to the adsorption of the inorganics from the
liquid phase to the solid phase (hydrochar) at elevated temperatures. The oversaturation
with minerals in the liquid phase at high temperatures could be responsible for this
precipitation phenomenon. Several researchers found that HTC enhances the degradation
process with higher temperatures, and produces sugar monomers, furfurals, and organic
acids, which leave porous structures of hydrochar. As such, some entrapped/loosely
bonded inorganics in the crosslinked matrix might have been adsorbed into the pores of
hydrochar during HTC [18,24,40]. Since cellulose and lignin start degrading at around
230 ◦ C and 260 ◦ C, respectively, and create porous structures, they might permit the
insoluble inorganics to be absorbed from the process liquid into the hydrochar surface,
resulting in higher AY.
The ultimate analysis shown in Table 2 indicates that with the increase in HTC temper-
ature, the carbon content was increased by ~10 wt. % for ORG HTC260 and ~11 wt. % for
HAF HTC260 from raw ORG and HAF, respectively, wherein the oxygen content dropped
about 13 wt. % and 14 wt. %. Regarding hydrogen, although ORG chars showed essentially
no change, HAF chars showed a slight drop (~1.5 wt. %) for HTC 260 from raw HAF. The
nitrogen showed insignificant changes in the hydrochars with respect to the feedstock.
The higher carbon content and lower oxygen content rationalized the rising HHV with
the increase in HTC process temperature. The fuel quality was also analyzed with the
van-Krevelen diagram (Figure 2). The van-Krevelen diagram depicts that the closer to
the origin of the data the points are, the better the fuel is [33]. The HAF fuel quality was
128
Energies 2021, 14, 1397
very low compared to the LAF quality, as expected due to the higher concentration of
non-combustible species. Even considering this on an inorganics-free basis, there is a small
discrepancy with a lower calorific value in the HAF. As mentioned above, this is attributed
to the partitioning of tissues during AC. In addition to having higher inorganics content,
these HAF tissues (larger portions of leaf, sheaths) have higher amounts of extractives
(cutin resin, protein, etc.). With the increase in HTC temperature, the hydrochar fuel quality
(calorific value, and carbon, oxygen, and hydrogen content) increases (up to 32 wt. % in the
HAF). Hydrochars produced at 260 ◦ C showed the best fuel quality among all the chars,
and this agrees with prior studies undertaken in these condition ranges [5]. As the rise in
HTC temperature favored the dehydration and decarboxylation reactions (see Figure 2),
this could justify the increase in elemental C and decrease in elemental O content during
the carbonization process.
The TGA of the ORG chars showed a decreasing trend (Table 2) of VM and a rising
trend of FC. The ORG HTC260 showed ~21 wt. % VM decrease and ~19 wt. % FC increase
from the raw ORG, whereas HAF HTC260 showed a significant decrease (~25 wt. %)
in VM and increase (~27 wt. %) in FC compared to raw HAF. However, a minimum
change in both FC and VM was observed for both feedstocks at 200 ◦ C. Previous studies
suggested that the degradation of cellulose at around 220 ◦ C might be responsible for
this phenomenon [36,42,43]. Sharma et al. [39] and Titirici et al. [44] further explained
cellulose degradation via two routes: (1) cellulose > glucose > 5-hydroxymethyl furfural >
carbonized structure; (2) cellulose > aromatic structure. In this study, a possible explanation
for receiving higher FC for both ORG and HAF HTC260 is the cellulose carbohydrate’s
degradation into more carbonaceous particles at successively higher temperatures.
The highest energy content was observed at the 260 ◦ C hydrochars for both ORG and
HAF (about 22 MJ/kg), as shown in Table 2, which was almost 4 MJ/kg and 5 MJ/kg higher
than with the ORG CS (~18 MJ/kg) and HAF (~17 MJ/kg), respectively. Earlier studies
demonstrated that with the increasing HTC temperature, the HHV and corresponding
EY showed upward and downward trends, respectively [20,22,24]. The EY reported in
Table 2 showed a descending trend for ORG chars, but a more complex pattern for HAF
hydrochars. A possible explanation for the discrepancies between the hydrochar EY values
is that the HAF HTC230 had higher MY, which could be due to the retention of cellulose in
the hydrochar due to partial degradation. In addition, the enriched energy content and
mass yield is convolved with the reduction in inorganics to create a potentially complex
optimization for the energy yield of the HAF.
129
Energies 2021, 14, 1397
The overall energy balances for the three scenarios were calculated using Equation (6)–(11).
For process units, the specific energy consumptions (SECs) were found from [45,46] and
shown in Table 3. Specific energy consumption for each process unit. The SEC of AC
was calculated as 0.0036 MJ/kg by INL and the basis was taken as per kg dry LAF. The
pellet press SEC was found to be 0.4 MJ/kg dry feed for the raw LAF and LAF–HTC
mixtures and 0.2 MJ/kg dry hydrochar for the ORG chars [45]. Earlier, it was demonstrated
that the pellet press SEC increases with a higher feed rate, feed moisture (wt. %), and
length-to-diameter (L/D) ratio of pellet die, etc. [47]. For example, the moisture (wt. %) in
the biomass feed is higher than in the hydrochar feed due to the higher hydrophilicity [48],
which can increase the overall SEC of biomass pelletization. Here, the SEC for biomass and
biomass–hydrochar mixtures under pelletization (Table 3. Specific energy consumption
for each process unit) was considered higher than the hydrochar SEC due to the higher
feed, moistures (wt. %), and/or L/D ratio. For the case of SEC of HTC, the required
SEC was found to be 5.9, 6.3, and 6.7 MJ/kg dry feed for HTC200, HTC230, and HTC260,
respectively (for both ORG and HAF chars) [46]. As higher temperatures require higher
energy consumption [46], HTC260 has the highest SEC (6.7 MJ/kg dry feed) among all the
chars. Table 4 shows the overall energy balance of the processes.
As per the AC process (Table 4), the overall net energy in was 1833.6 MJ (Equations (6),
(7), and (11) and Table 4) and the energy out was 1588.3 MJ (Equation (8), Table 4). The
recoverable energy of the HAF was found by the difference between the ORG and LAF
energies, which was ~245 MJ (Equation (9), Table 4). Since the HAF waste was not processed
and eliminated after AC, the recoverable energy was 0 MJ for the AC-only
In the case of the HTC-only process (Table 4), the overall energy in was 2439.6 MJ,
2474.1 MJ, and 2507.9 MJ for ORG HTC200, 230 and 260, respectively (Equations (6), (7) and
Table 4). The overall energy out was 1217.6 MJ, 1116.4 MJ and 953.8 MJ for HTC200, 230 and
260, respectively (Equation (8) and Table 4). The energy consumptions for the bulk quantity
of HTC of ORG significantly elevated the process’ overall energy demand. Since the HTC
process was highly energy-intensive, the energy recovery was not feasible. process.
130
Energies 2021, 14, 1397
As per the integrated AC–HTC process (Table 4), the HAF waste was hydrothermally
carbonized to utilize the recoverable energy present (~245 MJ) in the product, and further
mixed with the LAF stream for taking advantage of AC preprocessing (Equation (10),
and Table 4). The results in Table 4 show that the energy recovery was the maximum
(~80 MJ) for LAF:HAF under HTC230, which was about ~11 MJ and ~29 MJ higher than
LAF:HAF HTC200 and LAF:HAF HTC260, respectively (using Equations (6)–(8), (10), (11),
and Table 4). The higher FC and elemental carbon for HAF HTC230 than HAF HTC200
could contribute as a better solid fuel compared to LAF: HAF HTC200. On the other hand,
although HAF HTC260 has higher FC and elemental carbon than HAF HTC230, due to
the higher volatile matters as well as the elemental oxygen in HAF HTC230, this could
enhance the energy content of the LAF:HAF HTC230. Since volatiles are mainly made of
short chain hydrocarbons, long chain hydrocarbons and aromatic hydrocarbons that are
easy to distill off, it might be possible that the heavier hydrocarbons break into lighter gases
during combustion and react with limited oxygen via partial oxidation, releasing more
131
Energies 2021, 14, 1397
energy as heat [49]. In this context, integrating the LAF and HAF HTC230 process streams
might give a better energy recovery out of the AC reject stream. Moreover, in this mode,
the HTC process energy consumptions were significantly less than the HTC-only mode,
which might impact the overall energy recovery in a positive dimension. On the other
hand, the final product energy value was much higher than the AC-only mode, which bears
the importance of combining the AC and HTC processes together. Overall, the integrated
AC–HTC could become a better option to take the advantage of the potential waste energy
as valuable utilizable energy, and minimize the overall energy lost.
Equations used:
Net energy (MJ) of LAF pellet = LAF pellet-Pellet press for LAF (11)
4. Conclusions
This study investigated an integrated AC–HTC process to recover the energy from
corn stover. An air classifier was used to beneficiate CS into a purified biomass stream with
27 wt. % reduction in ash content from LAF with 84.4 wt. % mass recovery. The waste from
this fractionation technique was used as a feedstock for HTC to produce a high-energy fuel
pellet (19–22 MJ/kg). When CS was hydrothermally carbonized, both MY and AY were
decreased with the HTC temperature; meanwhile, the energy content of the hydrochar was
increased. This study showed that the AC and HTC processes cannot recover any energy
from the HAF stream individually. On the other hand, along with the high-energy densified
hydrochar, the integrated AC–HTC process further showed significant energy recovery
(~800 MJ/tonne) from the HAF. Therefore, this study provides evidence of an HTC that can
be integrated with AC to reduce the inorganic content and recover energy. Adopting HTC
with AC could potentially transform CS into an advanced biorefinery feedstock, while
still utilizing the otherwise lost organics from the AC process. Further sustainability and
technoeconomic assessment of AC and HTC are needed to justify the economic viability
of the integrated process. However, the energy consumption of this study was calculated
based on the laboratory-scale data, which could vary at a large scale. Process economics
with large-scale data could reveal additional data that might assist technology maturation.
132
Energies 2021, 14, 1397
References
1. U.S. Billion-Ton Update: Biomass Supply for a Bioenergy and Bioproducts Industry; U.S. Department of Energy, Energy Efficiency and
Renewable Energy, Office of the Biomass Program: Washington, DC, USA, 2011; p. 235.
2. Thompson, V.S.; Lacey, J.A.; Hartley, D.; Jindra, M.A.; Aston, J.E.; Thompson, D.N. Application of air classification and formulation
to manage feedstock cost, quality and availability for bioenergy. Fuel 2016, 180, 497–505. [CrossRef]
3. Perlack, R.D.; Wright, L.L.; Turhollow, A.F.; Graham, R.L.; Stokes, B.J.; Erbach, D.C. Biomass as Feedstock for a Bioenergy and
Bioproducts Industry: The Technical Feasibility of a Billion-Ton Annual Supply; Oak Ridge National Laboratory: Oak Ridge, TN, USA,
2005; p. 1216415. [CrossRef]
4. Zhichao, W.; Dunn, J.B.; Wang, M.Q. Updates to the Corn Ethanol Pathway and Development of an Integrated Corn and Corn Stover
Ethanol Pathway in the GREET™ Model; No. ANL/ESD-14/11; Argonne National Lab. (ANL): Argonne, IL, USA, 2014.
5. Machado, N.; Castro, D.; Queiroz, L.; Santos, M.; Costa, C. Production and Characterization of Energy Materials with Adsorbent
Properties by Hydrothermal Processing of Corn Stover with Subcritical H2 O. J. Appl. Solut. Chem. Model. 2016, 5, 117–130. [CrossRef]
6. Kim, S.; Dale, B.E. Global potential bioethanol production from wasted crops and crop residues. Biomass Bioenergy 2004,
26, 361–375. [CrossRef]
7. Zhang, Y.; Jiang, Q.; Xie, W.; Wang, Y.; Kang, J. Effects of temperature, time and acidity of hydrothermal carbonization on the
hydrochar properties and nitrogen recovery from corn stover. Biomass Bioenergy 2019, 122, 175–182. [CrossRef]
8. Biswas, B.; Pandey, N.; Bisht, Y.; Singh, R.; Kumar, J.; Bhaskar, T. Pyrolysis of agricultural biomass residues: Comparative study of
corn cob, wheat straw, rice straw and rice husk. Bioresour. Technol. 2017, 237, 57–63. [CrossRef]
9. Emerson, R.M.; Hernandez, S.; Williams, C.L.; Lacey, J.A.; Hartley, D.S. Improving bioenergy feedstock quality of high moisture
short rotation woody crops using air classification. Biomass Bioenergy 2018, 117, 56–62. [CrossRef]
10. Williams, C.L.; Emerson, R.M.; Hernandez, S.; Klinger, J.L.; Fillerup, E.P.; Thomas, B.J. Preprocessing and hybrid biochemi-
cal/thermochemical conversion of short rotation woody coppice for biofuels. Front. Energy Res. 2018, 6, 74. [CrossRef]
11. Lacey, J.A.; Aston, J.E.; Westover, T.L.; Cherry, R.S.; Thompson, D.N. Removal of introduced inorganic content from chipped
forest residues via air classification. Fuel 2015, 160, 265–273. [CrossRef]
12. Reza, M.T.; Emerson, R.; Uddin, M.H.; Gresham, G.; Coronella, C.J. Ash reduction of corn stover by mild hydrothermal
preprocessing. Biomass Convers. Biorefinery 2014. [CrossRef]
13. Humbird, D.; Davis, R.; Tao, L.; Kinchin, C.; Hsu, D.; Aden, A. Process Design and Economics for Biochemical Conversion of
Lignocellulosic Biomass to Ethanol: Dilute-Acid Pretreatment and Enzymatic Hydrolysis of Corn Stover; Technical Report NREL/TP-
5100-47764; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2011; p. 1013269.
14. Möller, M.; Nilges, P.; Harnisch, F.; Schröder, U. Subcritical Water as Reaction Environment: Fundamentals of Hydrothermal
Biomass Transformation. ChemSusChem 2011, 4, 566–579. [CrossRef]
15. Brunner, G. Heat Transfer. In situ Spectroscopic Techniques at High Pressure; Elsevier: Amsterdam, The Netherlands, 2014; Volume 5,
pp. 227–263. [CrossRef]
16. Öztürk, I.; Irmak, S.; Hesenov, A.; Erbatur, O. Hydrolysis of kenaf (Hibiscus cannabinus L.) stems by catalytical thermal treatment
in subcritical water. Biomass Bioenergy 2010, 34, 1578–1585. [CrossRef]
17. Machado, N.; De Castro, D.; Santos, M.; Araújo, M.; Lüder, U.; Herklotz, L.; Werner, M.; Mumme, J.; Hoffmann, T. Process
analysis of hydrothermal carbonization of corn Stover with subcritical H2 O. J. Supercrit. Fluids 2018, 136, 110–122. [CrossRef]
18. Funke, A.; Ziegler, F. Hydrothermal carbonization of biomass: A summary and discussion of chemical mechanisms for process
engineering. Biofuels Bioprod. Biorefin. 2010, 4, 160–177. [CrossRef]
19. Minaret, J.T. Hydrothermal Carbonization of Corn Residuals to Produce a Solid Fuel Replacement for Coal; The University of Guelph:
Keiu Lake, ON, Canada, 2015; p. 133.
20. Kobayashi, N.; Okada, N.; Hirakawa, A.; Sato, T.; Kobayashi, J.; Hatano, S.; Itaya, Y.; Mori, S. Characteristics of Solid Residues
Obtained from Hot-Compressed-Water Treatment of Woody Biomass. Ind. Eng. Chem. Res. 2009, 48, 373–379. [CrossRef]
21. Lynam, J.G.; Coronella, C.J.; Yan, W.; Reza, M.T.; Vasquez, V.R. Acetic acid and lithium chloride effects on hydrothermal
carbonization of lignocellulosic biomass. Bioresour. Technol. 2011, 102, 6192–6199. [CrossRef] [PubMed]
22. Yan, W.; Acharjee, T.C.; Coronella, C.J.; Vásquez, V.R. Thermal pretreatment of lignocellulosic biomass. Environ. Prog. Sustain.
Energy 2009, 28, 435–440. [CrossRef]
23. Hoekman, S.K.; Broch, A.; Robbins, C. Hydrothermal Carbonization (HTC) of Lignocellulosic Biomass. Energy Fuels 2011,
25, 1802–1810. [CrossRef]
133
Energies 2021, 14, 1397
24. Reza, M.T.; Lynam, J.G.; Uddin, M.H.; Coronella, C.J. Hydrothermal carbonization: Fate of inorganics. Biomass Bioenergy 2013,
49, 86–94. [CrossRef]
25. Tu, R.; Sun, Y.; Wu, Y.; Fan, X.; Wang, J.; Cheng, S.; Jia, Z.; Jiang, E.; Xu, X. Improvement of corn stover fuel properties via
hydrothermal carbonization combined with surfactant. Biotechnol. Biofuels 2019, 12, 1–19. [CrossRef]
26. Wang, T.; Zhai, Y.; Zhu, Y.; Li, C.; Zeng, G. A review of the hydrothermal carbonization of biomass waste for hydrochar formation:
Process conditions, fundamentals, and physicochemical properties. Renew. Sustain. Energy Rev. 2018, 90, 223–247. [CrossRef]
27. Wu, Q.; Yunqiao, P.; Hao, N.; Wells, T.; Meng, X.; Li, M.; Pu, Y.; Liu, S.; Ragauskas, A.J. Characterization of products from
hydrothermal carbonization of pine. Bioresour. Technol. 2017, 244, 78–83. [CrossRef] [PubMed]
28. Smith, A.M.; Singh, S.; Ross, A.B. Fate of inorganic material during hydrothermal carbonisation of biomass: Influence of feedstock
on combustion behaviour of hydrochar. Fuel 2016, 169, 135–145. [CrossRef]
29. Wagner, W.; Pruß, A. The IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water Substance for General
and Scientific Use. J. Phys. Chem. Ref. Data 2002, 31, 387–535. [CrossRef]
30. Archers, D.G.; Wang, P. The Dielectric Constant of Water and Debye-HOckel Limiting Law Slopes. J. Phys. Chem. Ref. Data 1990,
19, 41.
31. Bandura, A.V.; Lvov, S.N. The Ionization Constant of Water over Wide Ranges of Temperature and Density. J. Phys. Chem.
Ref. Data 2006, 35, 15–30. [CrossRef]
32. Kaliyan, N.; Morey, R.V. Densification Characteristics of Corn Stover and Switchgrass. Trans. ASABE 2009, 52, 907–920. [CrossRef]
33. Saba, A.; Saha, P.; Reza, M.T. Co-Hydrothermal Carbonization of coal-biomass blend: Influence of temperature on solid fuel
properties. Fuel Process. Technol. 2017, 167, 711–720. [CrossRef]
34. Saha, N.; Xin, D.; Chiu, P.C.; Reza, M.T. Effect of Pyrolysis Temperature on Acidic Oxygen-Containing Functional Groups and
Electron Storage Capacities of Pyrolyzed Hydrochars. ACS Sustain. Chem. Eng. 2019, 7, 8387–8396. [CrossRef]
35. Thompson, V.S.; Aston, J.E.; Lacey, J.A.; Thompson, D.N. Optimizing Biomass Feedstock Blends with Respect to Cost, Supply,
and Quality for Catalyzed and Uncatalyzed Fast Pyrolysis Applications. BioEnergy Res. 2017, 10, 811–823. [CrossRef]
36. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.-M.; Fühner, C.; Bens, O.; Kern, J.; et al.
Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry, processes and applications of wet and
dry pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
37. Pastor-Villegas, J.; Pastor-Valle, J.; Rodríguez, J.M.; García, M.G. Study of commercial wood charcoals for the preparation of
carbon adsorbents. J. Anal. Appl. Pyrolysis 2006, 76, 103–108. [CrossRef]
38. Kumar, S.; Gupta, R.; Lee, Y.; Gupta, R.B. Cellulose pretreatment in subcritical water: Effect of temperature on molecular structure
and enzymatic reactivity. Bioresour. Technol. 2010, 101, 1337–1347. [CrossRef] [PubMed]
39. Sharma, R.; Jasrotia, K.; Singh, N.; Ghosh, P.; Srivastava, S.; Sharma, N.R.; Singh, J.; Kanwar, R.; Kumar, A. A Comprehensive
Review on Hydrothermal Carbonization of Biomass and Its Applications. Chem. Afr. 2019, 3, 1–19. [CrossRef]
40. Qadi, N.; Takeno, K.; Mosqueda, A.; Kobayashi, M.; Motoyama, Y.; Yoshikawa, K. Effect of Hydrothermal Carbonization Conditions
on the Physicochemical Properties and Gasification Reactivity of Energy Grass. Energy Fuels 2019, 33, 6436–6443. [CrossRef]
41. Chen, S.-F.; Mowery, R.A.; Scarlata, C.J.; Chambliss, C.K.; Chambliss, K. Compositional Analysis of Water-Soluble Materials in
Corn Stover. J. Agric. Food Chem. 2007, 55, 5912–5918. [CrossRef]
42. Saha, N.; Saba, A.; Reza, M.T. Effect of hydrothermal carbonization temperature on pH, dissociation constants, and acidic
functional groups on hydrochar from cellulose and wood. J. Anal. Appl. Pyrolysis 2019, 137, 138–145. [CrossRef]
43. Peterson, A.A.; Vogel, F.; Lachance, R.P.; Fröling, M.; Antal, J.M.J.; Tester, J.W. Thermochemical biofuel production in hydrothermal
media: A review of sub- and supercritical water technologies. Energy Environ. Sci. 2008, 1, 32–65. [CrossRef]
44. Titirici, M.M.; Thomas, A.; Yu, S.-H.; Müller, A.J.-O.; Antonietti, M. A Direct Synthesis of Mesoporous Carbons with Bicontinuous
Pore Morphology from Crude Plant Material by Hydrothermal Carbonization. Chem. Mater. 2007, 19, 4205–4212. [CrossRef]
45. Tumuluru, J.S. High moisture corn stover pelleting in a flat die pellet mill fitted with a 6 mm die: Physical properties and specific
energy consumption. Energy Sci. Eng. 2015, 3, 327–341. [CrossRef]
46. Lucian, M.; Fiori, L. Hydrothermal Carbonization of Waste Biomass: Process Design, Modeling, Energy Efficiency and Cost
Analysis. Energies 2017, 10, 211. [CrossRef]
47. Tumuluru, J.S. Specific energy consumption and quality of wood pellets produced using high-moisture lodgepole pine grind in a
flat die pellet mill. Chem. Eng. Res. Des. 2016, 110, 82–97. [CrossRef]
48. Reza, M.T.; Lynam, J.G.; Vasquez, V.R.; Coronella, C.J. Pelletization of biochar from hydrothermally carbonized wood.
Environ. Prog. Sustain. Energy 2012, 31, 225–234. [CrossRef]
49. Wang, T. An overview of IGCC systems. In Integrated Gasification Combined Cycle (IGCC) Technologies; Wang, T., Stiegel, G., Eds.;
Woodhead Publishing: Cambridge, UK, 2017; pp. 1–80. [CrossRef]
134
energies
Article
Sewage Sludge Treatment by Hydrothermal Carbonization:
Feasibility Study for Sustainable Nutrient Recovery and
Fuel Production
Gabriel Gerner 1, *, Luca Meyer 1 , Rahel Wanner 1 , Thomas Keller 2 and Rolf Krebs 1
1 Institute of Natural Resource Sciences, Campus Grüental, Zurich University of Applied Sciences (ZHAW),
CH-8820 Wädenswil, Switzerland; luca.meyer@zhaw.ch (L.M.); rahel.wanner@zhaw.ch (R.W.);
rolf.krebs@zhaw.ch (R.K.)
2 Institute of Chemistry and Biotechnology, Campus Reidbach, Zurich University of Applied Sciences (ZHAW),
CH-8820 Wädenswil, Switzerland; thomas.keller2@zhaw.ch
* Correspondence: gabriel.gerner@zhaw.ch; Tel.: +41-58-934-5588
Abstract: Phosphorus recovery from waste biomass is becoming increasingly important, given that
phosphorus is an exhaustible non-renewable resource. For the recovery of plant nutrients and
production of climate-neutral fuel from wet waste streams, hydrothermal carbonization (HTC) has
been suggested as a promising technology. In this study, digested sewage sludge (DSS) was used as
waste material for phosphorus and nitrogen recovery. HTC was conducted at 200 ◦ C for 4 h, followed
by phosphorus stripping (PS) or leaching (PL) at room temperature. The results showed that for PS
and PL around 84% and 71% of phosphorus, as well as 53% and 54% of nitrogen, respectively, could be
recovered in the liquid phase (process water and/or extract). Heavy metals were mainly transferred
to the hydrochar and only <1 ppm of Cd and 21–43 ppm of Zn were found to be in the liquid phase
Citation: Gerner, G.; Meyer, L.; of the acid treatments. According to the economic feasibility calculation, the HTC-treatment per dry
Wanner, R.; Keller, T.; Krebs, R. ton DSS with an industrial-scale plant would cost around 608 USD. Between 349–406 kg of sulfuric
Sewage Sludge Treatment by acid are required per dry ton DSS to achieve a high yield in phosphorus recovery, which causes
Hydrothermal Carbonization: additional costs of 96–118 USD. Compared to current sewage sludge treatment costs in Switzerland,
Feasibility Study for Sustainable
which range between 669 USD and 1173 USD, HTC can be an economically feasible process for DSS
Nutrient Recovery and Fuel
treatment and nutrient recovery.
Production. Energies 2021, 14, 2697.
https://doi.org/10.3390/en14092697
Keywords: hydrothermal carbonization; phosphorus recovery; digested sewage sludge; hydrochar;
Academic Editor: M. Toufiq Reza
nutrient recovery; climate-neutral fuel; energy efficiency; economic feasibility
contrast to phosphorus, which relies on a definite source, nitrogen fertilizer can be synthe-
sized as ammonia from air by the Haber–Bosch process [13]. The negative aspect of this
process is that to produce ammonia fertilizer, it consumes between 1–2% of global energy
and produces around 1.4% of global CO2 emissions [14]. Countries like Switzerland and
Germany approved new regulations to reduce the dependency on phosphorus imports.
It will be mandatory to recycle phosphorus from sewage sludge by 2026 in Switzerland
and by 2029 in Germany [15]. This attempt to reduce mismanagement of this exhaustible
resource and the need for high-quality fertilizer has increased the necessity of a phosphorus
recycling process with high nutrient recovery and low environmental burden. Hydrother-
mal carbonization (HTC) is a thermochemical process allowing the direct usage of wet
feedstock without drying them and converting it under high pressure and temperature
to a coal slurry, which can be separated into an energy-rich solid phase (hydrochar) and
nutrient-rich liquid phase (process water) [16]. The conversion leads to reduced NO emis-
sion in the hydrochar combustion [17] and simultaneously improves dewaterability of
the carbonized sludge [18,19], which is crucial for an energy-efficient separation. In this
process, phosphorus is mainly incorporated in the hydrochar and has to be removed by
acid leaching [20–22]. Other processes use sewage sludge ash for P-leaching [23], where the
sludge’s fuel property is mainly used as process energy for the incineration. Therefore, it is
lost to industrial processes like the cement industry or coal power plants as climate-neutral
fuel.
Modern wastewater treatment plants (WWTP) employ biological and/or chemical
phosphorus removal technologies to transfer around 90% of the input phosphorus load
into the sewage sludge [24]. Larger WWTPs utilize the sludge as a feedstock for anaerobic
digestion (AD) producing methane, which can be converted to electricity and heat to cover
energy demands. Implementing an HTC plant on-site allows the methane yield to be
increased by feeding the AD with HTC process water (PW), as a supplemental feedstock
with high organic carbon content [20,25–30]. Recovered energy can at the same time be
utilized for the HTC process and increase the energy efficiency.
Different studies show the potential of P-recovery by applying acids before or after
the HTC treatment [20,22,31,32] and the need for using digested sewage sludge. For the
acid application, digested sludge has a higher P availability compared to raw sludge, while
with raw sludge only 50% of P can be recovered at pH 2 [23]. For a profitable application
of the process a cost-effective usage of acid must be achieved.
The aim of this study was to investigate the recovery of plant nutrients (phosphorus
and nitrogen) from digested sewage sludge for a sustainable fertilizer and fuel production
with the main focus on minimizing the acid usage and increasing the PW utilization.
Two paths of P-recovery are investigated in lab-scale experiments, with immediate acid
application before and after liquid–solid separation and without prior drying to simulate
industrial processing. Process liquids are analyzed for their nutrient and heavy metal
content and hydrochars are examined for their fuel properties as possible substitutes for
fossil fuels. Furthermore, the economic feasibility of the process in an industrial-scale
HTC-plant is evaluated.
136
Energies 2021, 14, 2697
Germany). The slowly dried stock was homogenized by grinding to a fine powder (GM200,
Retsch GmbH, Haan, Germany) and stored in airtight 1 L glass bottles (DURAN® GLS
80® laboratory wide mouth bottle, DURAN Group GmbH, Mainz, Germany) for HTC
trials. These sludge pre-treatments are conducted to achieve a better reproducibility of
the lab-scale experiments. In large-scale HTC-treatments sewage sludge will be treated
directly after dewatering, without previous drying step.
137
Energies 2021, 14, 2697
Figure 1. Flow charts of experimental set-ups with and without acid addition. (HTC = Hydrothermal carbonization, C =
Control, PS = P-Stripping, PL = P-Leaching, DSS = Digested sewage sludge, PW = Process water, Ext = Leaching extract, VF
= Vacuum filtration).
O (wt.%) = 100 − C (wt.%) − H (wt.%) − N (wt.%) − S (wt.%) − Ash content (wt.%) (2)
Total content of HM (Cd, Cu, Ni, Pb and Zn) and phosphorus (P) was measured for
liquid and solid samples spectroscopically by ICP-OES (Agilent 5100, Agilent Technologies,
CA, USA). 2 mL of liquid sample or 0.35 g of solid sample (dried at 40 ◦ C) was added
to 8 mL of aqua regia and microwave digested (Speedwave Four, Berghof Products +
Instruments GmbH, Eningen, Germany) for 35 min at 175 ◦ C (in dependence on SN EN
13346, protocol C). After the acid digestion samples were transferred to a 25 mL volumetric
flask and diluted with ultrapure water. Final results for solid samples were corrected
by the remaining water content (% DM105C ). Acid digestion using aqua regia did not
lead to a complete digestion for solid char samples. Therefore, measurement results were
utilized for material balance and recovery efficiency was determined. The results showed
good overall recovery with >83% for HM (excl. Pb, Cd and Hg) and nutrients. Dilution
effects in liquid samples faced detection limitations, especially for elements with low
138
Energies 2021, 14, 2697
concentrations as Pb, Cd and Hg. Phosphorus contents for hydrochars were calculated by
difference. All samples were digested in triplicates and average results are reported. For
the determination of mercury (Hg) samples dried at 40 ◦ C were measured according to DIN
EN 1483: 08.97. Mercury was analyzed using a cold vapor atomic absorption spectrometer
(CV-AAS) (novAA® 350, Analytik Jena GmbH, Jena, Germany) equipped with a hydride
generator (HS 60A, Analytik Jena GmbH, Jena, Germany).
Liquid samples (process water and leaching extract) were analyzed for total organic
carbon (TOC), measured as non-purgeable organic carbon (NPOC), according to ASTM
D7573 and total bound nitrogen (TNb) according to DIN EN 12260 with a TOC-LCSH
analyzer equipped with a TNM-L unit (Shimadzu, Kyoto, Japan). For TOC measurements
the samples were automatically acidified with HCl to pH < 3 and sparged with purified air
to remove inorganic carbon. Purgeable organic carbon (POC) may also be lost during this
sample treatment. The pH values of fresh process water and leachate were measured with
a portable multi-parameter meter (HQ40d, Hach Lange, Düsseldorf, Germany).
Hydrochars were further characterized for its fuel properties. Regarding energy con-
tent, the higher heating value (HHV) of the raw material and the hydrochar was measured
using a calorimeter (IKA C 200, Breisgau, Germany). The volumetric emissions of CO2 and
SO2 were calculated according to Equations (3) and (4) adapted from Kaltschmitt [35], on
the assumption of a complete combustion. Weight percentage of carbon (c) and sulfur (s)
are given by the elemental analysis:
m3 CO2 c
CO2 emission = 22.41 (3)
kg DM 12
m3 SO2 s
SO2 emission = 22.41 (4)
kg DM 32
The hydrochar yield and energy efficiency (EE) were calculated by Equations (5) and
(6), respectively [36–38].
DMhydrochar (kg)
Hydrochar yield (%) = × 100 (5)
DMraw material (kg)
⎛ ⎞
MJ
HHVhydrochar kg
Energy efficiency (%) = (Hydrochar yield) × ⎝ ⎠ (6)
MJ
HHVraw material kg
139
Energies 2021, 14, 2697
grade H2 SO4 98% was applied. Specific overall costs per dry ton DSS were determined by
Equation (7).
Annual fixed cos ts (CHF per year)
Specific overall cos ts (CHF per ton DSS) = (7)
Annual troughput (tons per year)
The industrial scale HTC-plant from GRegio Energie AG is operating continuously
with a reactor volume of 5.6 m3 . The maximum flow rate of the current system is 2 m3 /h
and is limited by the heating system and heat recovery unit. The optimal DM content
of the sludge input material is around 15%. For the calculation of the disposal fees the
throughput was set at 11,200 tons dewatered sewage sludge (15%DM) per year. With a
carbonization time of 4 h and a yearly operation time of 8000 h, the flow rate was set to
1.4 m3 /h.
Table 1. Ultimate analysis and higher heating values (HHV) of digested sewage sludge (DSS) and its
derived hydrochars (HTC-xx). (C = Control; PS = P stripping; PL = P leaching; HC = Hydrochar) (n =
2).
Table 1 summarizes the results from ultimate analyses, as well as gross calorific values
on raw material and hydrochar samples. The ash content of the hydrochars decreased by
10% after the acid treatment, with a minimum ash content of 45.1% after P stripping. Volu-
metric carbon dioxide and sulfur dioxide emissions c are shown in Figure 2. Hydrochars
after acid addition show an increase in sulfur content by 9–10 times, which has a direct
influence on the SO2 emissions. Hydrochars from HTC-PS and HTC-PL contained more
than 6 wt.% sulfur which, according to the Swiss Clean Air Act (LRV, Annex 5 Number 2),
is 3% above the recommended value for coal, coal briquettes and cokes [39]. The carbon
content of the chars as well as the CO2 emissions are very stable and did not show any
major changes (Table 1, Figure 2).
140
Energies 2021, 14, 2697
Figure 2. Volumetric carbon dioxide and sulfur dioxide emission per kg dry matter (DM) (Error bars
show standard deviation of n = 2).
Figure 3. Recovery of main nutrients in the process water (PW) and leaching extract (Ext) (Error bars
show standard deviation of n = 2).
Tables 2 and 3 show the nutrient and heavy metal contents in the raw material,
hydrochars and liquid phase. Regarding heavy metal contaminations in the liquid phase,
only cadmium and zinc could be detected and quantified. The appearance of Cd could
141
Energies 2021, 14, 2697
not be explained by the authors, as it was not found in the raw material and must have
accumulated in the PW. A release of Zn was also observed by Becker et al. [22]. Elemental
analysis confirmed that most HM stayed in the hydrochar. An explanation for the high
selectivity is the sorption property of the char [22] and the reduction of HM solubility after
the HTC-treatment [40].
Table 2. Content analysis of raw material and hydrochars (HC). (C = Control; PS = P stripping; PL = P leaching; BLD =
bellow limit of detection) (n = 2).
Table 3. Content analysis of process water (PW) and leaching extract (Ext). (C = Control; PS = P stripping; PL = P leaching; BLD =
bellow limit of detection) (n = 2).
Because of high sulfur and heavy metal content in the hydrochar, it can only be used
for energy production in industrial-scale plants (e.g., cement industry, coal-fired power
plant, mono-incineration) with appropriate fume cleaning systems. To decrease the sulfur
content in the char, sulfuric acid could be substituted with nitric or organic acids (e.g., citric
acid) for the P-recovery treatment. However, other mineral and organic acids can increase
the leaching costs [22].
While the N concentrations were for all four liquid samples (PW and Ext) in a similar
range (6195–6960 mg N/L), differences in P concentration could be observed. The PW after
acid stripping contained 8113 mg P/L, whereas the highest P concentration was achieved
in the leachate extract (HTC-PL_Ext) with 16,050 mg P/L. The acid consumption was for
both methods (HTC-PS and HTC-PL) at around 14.2 kg (±0.3 kg) H2 SO4 per kg P.
142
Energies 2021, 14, 2697
Figure 4. Dewaterability (DW), hydrochar (HC) yield and energy efficiency of the raw material (DSS
as received), the control and acid treated hydrochars (Error bars show standard deviation of n = 2).
3.4. Profitability
Calculations for annual fixed costs are presented in Table 4. The specific costs per ton
DSS for the HTC treatment depend on the DM content of the delivered sludge and can
range between 91 USD and 183 USD with 15% and 30% DM content, respectively. On a dry
basis, the specific overall costs are 608 USD per ton DM. In Switzerland average disposal
costs (incl. transportation) per dry ton sewage sludge amounted in 2001 to: (i) 827 USD
for sewage sludge incineration plant; (ii) 882 USD for municipal solid waste incineration
plant; and (iii) 669 USD for cement plant (including drying) [45]. The WWTP Werdhölzli in
Zurich is one of the biggest WWTPs in Switzerland and operates its own sludge incinerator.
The sludge capacity of the plant is designed so that external WWTPs can also deliver their
sewage sludge for disposal. The disposal of sludge at the WWTP Werdhölzli costs currently
1173 USD per dry ton [46]. Besides the disposal costs, by 2026 the acid treatment will be an
important cost factor for WWTPs. In the current study the acid usage per dry ton DSS was
for PS 406 kg and for PL 349 kg, which adds additional costs of 111 USD and 96 USD per
dry ton DSS, respectively. Based on the acid usage per kg P recovered, both methods are
equal with acid costs of 3.96 USD per kg P.
Table 4. Calculation of net costs for the HTC process from GRegio Energie AG (amortization time 10
years, interest rate 2%; db = dry basis; DSS = digested sewage sludge).
143
Energies 2021, 14, 2697
4. Conclusions
Two methods for the nutrient recovery and fuel production were evaluated. For
the treatments PS and PL the acid consumption per kg P recovered were, at around 14.2
kg (±0.3 kg) H2 SO4 , equal, whereas for PL an advantage was the reduction of acidified
PW and doubling the P-concentration in the extract. The additional treatment step with
sedimentation and decantation and the reduced P-recovery efficiency compared to PS are
the only drawbacks of this treatment. In comparison, PS produced a char with a slightly
higher HHV, which could be favorable as a fuel substitute for industrial-scale plants. Both
treatments exceed the regulatory requirements for P recovery and additionally transfer
around 50% of N into the liquid phase, while leaving the main part of heavy metals in
the hydrochar. A reduction in char-N content improves as well the combustion emissions
regarding NOx . Regarding the high char-S content, only industrial plants with appropriate
exhaust gas treatments are applicable. Therefore, sewage sludge will be still available as a
climate-neutral fuel for the cement industry and coal power plants.
HTC can compete with current costs for sewage sludge disposal and offers an advan-
tageous alternative to mono-incineration as a sludge treatment, regarding fuel and nutrient
recovery. For an implementation in current WWTP infrastructures plants with AD are
favored because of their sludge property and the possibility of PW fermentation for higher
methane recovery. To promote the P recovery from waste streams like sewage sludge,
further investigations in a sustainable and economically feasible fertilizer production are
needed.
Author Contributions: Conceptualization, G.G. and L.M.; methodology, G.G.; validation, G.G.
and T.K.; formal analysis, G.G., L.M. and T.K.; investigation, G.G., L.M., R.W. and T.K.; resources,
G.G., L.M., R.W. and T.K.; data curation, G.G. and L.M.; writing—original draft preparation, G.G.;
writing—review and editing, G.G. and R.K.; visualization, G.G.; supervision, G.G. and R.K.; project
administration, G.G.; funding acquisition, G.G. All authors have read and agreed to the published
version of the manuscript.
Funding: This research was funded by cemsuisse, the Association of the Swiss Cement Industry. The
article processing charges for the open access publication was funded by ZHAW Zurich University
of Applied Sciences.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: The authors would like to thank for the financial support received from cemsu-
isse, the Association of the Swiss Cement Industry and ZHAW Zurich University of Applied Sciences
for funding the open access publication. The authors would also like to thank Samuel Solin and
Simon Kaiser, Institute of Bioenergy and Resource Efficiency at the FHNW University of Applied
Sciences and Arts Northwestern Switzerland, for the help with the elemental analysis.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript.
The project funders agreed to publish the results.
References
1. Mathieux, F.; Ardente, F.; Bobba, S.; Nuss, P.; Blengini, G.; Alves Dias, P.; Blagoeva, D.; Torres De Matos, C.; Wittmer, D.; Pavel, C.;
et al. Critical Raw Materials and the Circular Economy—Background Report; Publications Office of the European Union: Luxembourg,
2017.
2. European Commission. Communication from the Commission to the European Parliament, the Council, the European Economic and Social
Committee and the Committee of the Regions; On the Review of the List of Critical Raw Materials for the EU and the Implementation
of the Raw Materials Initiative; European Commission: Brussels, Belgium, 2014.
3. Binder, C.R.; de Baan, L.; Wittmer, D. Phosphorflüsse in Der Schweiz. Stand, Risiken Und Handlungsoptionen; Umwelt-Wissen;
Bundesamt für Umwelt: Bern, Switzerland, 2009; p. 161.
4. Daneshgar, S.; Callegari, A.; Capodaglio, A.; Vaccari, D. The Potential Phosphorus Crisis: Resource Conservation and Possible
Escape Technologies: A Review. Resources 2018, 7, 37. [CrossRef]
144
Energies 2021, 14, 2697
5. Cordell, D.; White, S. Life’s Bottleneck: Sustaining the World’s Phosphorus for a Food Secure Future. Annu. Rev. Environ. Resour.
2014, 39, 161–188. [CrossRef]
6. Hermann, L. Rückgewinnung von Phosphor aus der Abwassereinigung; Umwelt-Wissen; Bundesamt für Umwelt: Bern, Switzerland,
2009; p. 196.
7. Bigalke, M.; Ulrich, A.; Rehmus, A.; Keller, A. Accumulation of Cadmium and Uranium in Arable Soils in Switzerland. Environ.
Pollut. 2017, 221, 85–93. [CrossRef] [PubMed]
8. Bigalke, M.; Schwab, L.; Rehmus, A.; Tondo, P.; Flisch, M. Uranium in Agricultural Soils and Drinking Water Wells on the Swiss
Plateau. Environ. Pollut. 2018, 233, 943–951. [CrossRef]
9. Bigalke, M.; Imseng, M.; Schneider, S.; Schwab, L.; Wiggenhauser, M.; Keller, A.; Müller, M.; Frossard, E.; Wilcke, W. Uranium
Budget and Leaching in Swiss Agricultural Systems. Front. Environ. Sci. 2020, 8, 54. [CrossRef]
10. Schnug, E.; Haneklaus, N. Uranium in phosphate fertilizers—Review and outlook. In Uranium—Past and Future Challenges;
Merkel, B.J., Arab, A., Eds.; Springer International Publishing: Cham, Germany, 2015; pp. 123–130. ISBN 978-3-319-11058-5.
11. Tulsidas, H.; Gabriel, S.; Kiegiel, K.; Haneklaus, N. Uranium Resources in EU Phosphate Rock Imports. Resour. Policy 2019, 61,
151–156. [CrossRef]
12. Roth, N.; FitzGerald, R. Human and Environmental Impact of Uranium Derived from Mineral Phosphate Fertilizers; Swiss Centre for
Applied Human Toxicology: Basel, Switzerland, 2015; p. 50.
13. Nutrient Source Specifics—Ammonia. Available online: www.ipni.net/specifics (accessed on 19 April 2021).
14. Kyriakou, V.; Garagounis, I.; Vourros, A.; Vasileiou, E.; Stoukides, M. An Electrochemical Haber-Bosch Process. Joule 2020, 4,
142–158. [CrossRef]
15. Krämer, J. Phosphorrecycling: Wer, Wie, Was?—Umsetzung Einer Iterativen, Zielgruppenorientierten Kommunikationsstrategie; Deutsche
Phosphor Plattform: Frankfurt, Germany, 2019; p. 108.
16. Libra, J.A.; Ro, K.S.; Kammann, C.; Funke, A.; Berge, N.D.; Neubauer, Y.; Titirici, M.-M.; Fühner, C.; Bens, O.; Kern, J.; et al.
Hydrothermal Carbonization of Biomass Residuals: A Comparative Review of the Chemistry, Processes and Applications of Wet
and Dry Pyrolysis. Biofuels 2011, 2, 71–106. [CrossRef]
17. Zhao, P.; Chen, H.; Ge, S.; Yoshikawa, K. Effect of the Hydrothermal Pretreatment for the Reduction of NO Emission from Sewage
Sludge Combustion. Appl. Energy 2013, 111, 199–205. [CrossRef]
18. Escala, M.; Zumbühl, T.; Koller, C.; Junge, R.; Krebs, R. Hydrothermal Carbonization as an Energy-Efficient Alternative to
Established Drying Technologies for Sewage Sludge: A Feasibility Study on a Laboratory Scale. Energy Fuels 2013, 27, 454–460.
[CrossRef]
19. Ahmed, M.; Andreottola, G.; Elagroudy, S.; Negm, M.S.; Fiori, L. Coupling Hydrothermal Carbonization and Anaerobic Digestion
for Sewage Digestate Management: Influence of Hydrothermal Treatment Time on Dewaterability and Bio-Methane Production.
J. Environ. Manag. 2021, 281, 111910. [CrossRef] [PubMed]
20. Marin-Batista, J.D.; Mohedano, A.F.; Rodríguez, J.J.; de la Rubia, M.A. Energy and Phosphorous Recovery through Hydrothermal
Carbonization of Digested Sewage Sludge. Waste Manag. 2020, 105, 566–574. [CrossRef]
21. Maurizio, V.; Luca, F.; Fabio, M.; Antonio, M. Andreottola Gianni Hydrothermal Carbonization as an Efficient Tool for Sewage
Sludge Valorization and Phosphorous Recovery. Chem. Eng. Trans. 2020, 80, 199–204. [CrossRef]
22. Becker, G.C.; Wüst, D.; Köhler, H.; Lautenbach, A.; Kruse, A. Novel Approach of Phosphate-Reclamation as Struvite from Sewage
Sludge by Utilising Hydrothermal Carbonization. J. Environ. Manag. 2019, 238, 119–125. [CrossRef]
23. Egle, L.; Rechberger, H.; Zessner, M. Phosphorrückgewinnung aus dem Abwasser; Bundesministerium für Land- und Forstwirtschaft,
Umwelt und Wasserwirtschaft, Sektion VIIWasser: Vienna, Austria, 2014; p. 323.
24. Fux, C.; Theiler, M.; Irzan, T. Studie Phosphorrückgewinnung aus Abwasser und Klärschlamm; TBF+Partner AG: Zürich, Switzerland,
2015; p. 92.
25. Blöhse, D. Anaerobe Verwertung von HTC-Prozesswässern. In Proceedings of the Biokohle im Blick—Herstellung, Einsatz und
Bewertung, 73. Symposium des ANS e.V., Berlin, Germany, 20 September 2012.
26. Ferrentino, R.; Merzari, F.; Fiori, L.; Andreottola, G. Coupling Hydrothermal Carbonization with Anaerobic Digestion for Sewage
Sludge Treatment: Influence of HTC Liquor and Hydrochar on Biomethane Production. Energies 2020, 13, 6262. [CrossRef]
27. Zhao, K.; Li, Y.; Zhou, Y.; Guo, W.; Jiang, H.; Xu, Q. Characterization of Hydrothermal Carbonization Products (Hydrochars and
Spent Liquor) and Their Biomethane Production Performance. Bioresour. Technol. 2018, 267, 9–16. [CrossRef] [PubMed]
28. Parmar, K.R.; Ross, A.B. Integration of Hydrothermal Carbonisation with Anaerobic Digestion; Opportunities for Valorisation of
Digestate. Energies 2019, 12, 1586. [CrossRef]
29. Aragón-Briceño, C.; Ross, A.B.; Camargo-Valero, M.A. Evaluation and Comparison of Product Yields and Bio-Methane Potential
in Sewage Digestate Following Hydrothermal Treatment. Appl. Energy 2017, 208, 1357–1369. [CrossRef]
30. Merzari, F.; Langone, M.; Andreottola, G.; Fiori, L. Methane Production from Process Water of Sewage Sludge Hydrothermal
Carbonization. A Review. Valorising Sludge through Hydrothermal Carbonization. Crit. Rev. Environ. Sci. Technol. 2019, 49,
947–988. [CrossRef]
31. Shi, Y.; Luo, G.; Rao, Y.; Chen, H.; Zhang, S. Hydrothermal Conversion of Dewatered Sewage Sludge: Focusing on the
Transformation Mechanism and Recovery of Phosphorus. Chemosphere 2019, 228, 619–628. [CrossRef]
32. Aragón-Briceño, C.I.; Pozarlik, A.K.; Bramer, E.A.; Niedzwiecki, L.; Pawlak-Kruczek, H.; Brem, G. Hydrothermal Carbonization
of Wet Biomass from Nitrogen and Phosphorus Approach: A Review. Renew. Energy 2021, 171, 401–415. [CrossRef]
145
Energies 2021, 14, 2697
33. Biswas, B.K.; Inoue, K.; Harada, H.; Ohto, K.; Kawakita, H. Leaching of Phosphorus from Incinerated Sewage Sludge Ash by
Means of Acid Extraction Followed by Adsorption on Orange Waste Gel. J. Environ. Sci. 2009, 21, 1753–1760. [CrossRef]
34. BBodSch, V. Federal Soil Protection and Contaminated Sites Ordinance; German Federal Council: Berlin, Germany, 1999; p. 1554.
35. Kaltschmitt, M.; Hartmann, H.; Hofbauer, H. (Eds.) Energie aus Biomasse: Grundlagen, Techniken und Verfahren; Springer:
Berlin/Heidelberg, Germany, 2016; ISBN 978-3-662-47438-9.
36. Gao, Y.; Yu, B.; Wu, K.; Yuan, Q.; Wang, X.; Chen, H. Physicochemical, Pyrolytic, and Combustion Characteristics of Hydrochar
Obtained by Hydrothermal Carbonization of Biomass. BioResources 2016, 11, 4113–4133. [CrossRef]
37. Li, Z.; Yi, W.; Li, Z.; Tian, C.; Fu, P.; Zhang, Y.; Zhou, L.; Teng, J. Preparation of Solid Fuel Hydrochar over Hydrothermal
Carbonization of Red Jujube Branch. Energies 2020, 13, 480. [CrossRef]
38. Lucian, M.; Fiori, L. Hydrothermal Carbonization of Waste Biomass: Process Design, Modeling, Energy Efficiency and Cost
Analysis. Energies 2017, 10, 211. [CrossRef]
39. Luftreinhalte-Verordnung; Swiss Federal Council: Bern, Switzerland, 1985; (Status as of 1 April 2020); p. 82.
40. Liu, M.; Duan, Y.; Bikane, K.; Zhao, L. The Migration and Transformation of Heavy Metals in Sewage Sludge during Hydrothermal
Carbonization Combined with Combustion. BioMed Res. Int. 2018, 2018, 1913848. [CrossRef]
41. Wei, H.; Gao, B.; Ren, J.; Li, A.; Yang, H. Coagulation/Flocculation in Dewatering of Sludge: A Review. Water Res. 2018, 143,
608–631. [CrossRef]
42. Böhler, M.; Siegrist, H.; Pinnow, D.; Müller, D.; Krauss, W.; Brauchli, H. Neue Presstechnologie zur verbesserten Klärschlammen-
twässerung; EAWAG: Dübendorf, Switzerlnd, 2003; p. 42.
43. Funke, A.; Ziegler, F. Hydrothermal Carbonization of Biomass: A Summary and Discussion of Chemical Mechanisms for Process
Engineering. Biofuels Bioprod. Biorefin. 2010, 4, 160–177. [CrossRef]
44. Chen, Y.; Yang, H.; Gu, G. Effect of Acid and Surfactant Treatment on Activated Sludge Dewatering and Settling. Water Res. 2001,
35, 2615–2620. [CrossRef]
45. Laube, A.; Vonplon, A. Klärschlammentsorgung in Der Schweiz—Mengen- Und Kapazitätserhebung; Umwelt-Materialien; BUWAL:
Bern, Switzerland, 2004; p. 47.
46. Schlamm Aus Klärwerken—Preise Für Die Anlieferung von Schlamm Aus Klärwerken. Available online: www.stadt-zuerich.ch/
ted/de/index/entsorgung_recycling/sauberes_wasser/entsorgen/schlamm_aus_klaerwerken.html (accessed on 21 January
2021).
146
energies
Article
Thermal Analysis and Kinetic Modeling of Pyrolysis and
Oxidation of Hydrochars
Gabriella Gonnella 1 , Giulia Ischia 1 , Luca Fambri 2 and Luca Fiori 1, *
1 Department of Civil, Environmental and Mechanical Engineering, University of Trento, 38123 Trento, Italy;
ga.gonnella@gmail.com (G.G.); giulia.ischia-1@unitn.it (G.I.)
2 Department of Industrial Engineering, University of Trento, Via Sommarive 9, 38123 Trento, Italy;
luca.fambri@unitn.it
* Correspondence: luca.fiori@unitn.it
Abstract: This study examines the kinetics of pyrolysis and oxidation of hydrochars through thermal
analysis. Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) techniques
were used to investigate the decomposition profiles and develop two distributed activation energy
models (DAEM) of hydrochars derived from the hydrothermal carbonization of grape seeds produced
at different temperatures (180, 220, and 250 ◦ C). Data were collected at 1, 3, and 10 ◦ C/min between
30 and 700 ◦ C. TGA data highlighted a decomposition profile similar to that of the raw biomass for
hydrochars obtained at 180 and 220 ◦ C (with a clear distinction between oil, cellulosic, hemicellulosic,
and lignin-like compounds), while presenting a more stable profile for the 250 ◦ C hydrochar. DSC
showed a certain exothermic behavior during pyrolysis of hydrochars, an aspect also investigated
through thermodynamic simulations in Aspen Plus. Regarding the DAEM, according to a Gaussian
model, the severity of the treatment slightly affects kinetic parameters, with average activation
energies between 193 and 220 kJ/mol. Meanwhile, the Miura–Maki model highlights the distributions
of the activation energy and the pre-exponential factor during the decomposition.
Citation: Gonnella, G.; Ischia, G.; Keywords: hydrothermal carbonization; modeling; kinetics; thermal analysis; Aspen Plus; biomass;
Fambri, L.; Fiori, L. Thermal Analysis hydrochar
and Kinetic Modeling of Pyrolysis
and Oxidation of Hydrochars.
Energies 2022, 15, 950. https://
doi.org/10.3390/en15030950
1. Introduction
Academic Editors: M. Toufiq Reza The global increase in energy demand, driven mainly by the growing global popula-
and Javier Fermoso tion, urbanization, and economic growth, is depleting the world’s reserves of fossil fuels,
Received: 30 December 2021
already strained by their high consumption rate, heightening environmental concerns, re-
Accepted: 24 January 2022 source depletion, and rising costs. To face the problem, different renewable energy sources
Published: 27 January 2022 have been investigated, like wind, solar, hydropower, geothermal, and biomass. Among
the different biomasses, those deriving from waste or agricultural residues, agro-industrial,
Publisher’s Note: MDPI stays neutral
horticulture, and wood processing, are cheap and convenient [1,2]. Besides, the use and
with regard to jurisdictional claims in
valorizing of waste biomass are part of a circular economy context, where residues become
published maps and institutional affil-
the input for a new process [3]. At present, most of the waste produced in Europe is dis-
iations.
posed of without any further treatment resulting in wastefulness of energy and resources
and in serious environmental impact. A high-efficient use of waste not only increases its
economic value but also results in environmental benefits [4].
Copyright: © 2022 by the authors. In the last decades, biomass waste treatment and management have seen an increasing
Licensee MDPI, Basel, Switzerland. interest from the research community. Currently, biomass waste has characteristics that limit
This article is an open access article its use, including a low heating value, low fixed carbon content, high moisture content, low
distributed under the terms and grindability, and non-uniformity. Studies on the conversion of biomass into high-quality
conditions of the Creative Commons biofuels have been performed to avoid the problems caused by direct combustion and
Attribution (CC BY) license (https:// find more effective methods of utilization, such as pyrolysis, gasification, torrefaction, and
creativecommons.org/licenses/by/ hydrothermal carbonization [5,6].
4.0/).
148
Energies 2022, 15, 950
collected and used to determine kinetic parameters through two DAEMs (Gaussian and
Miura–Maki DAEM). To do this, TGA data at different heating rates were collected, and
activation energies and pre-exponential factors were computed. Analyses were also per-
formed on the raw biomass and the oil contained inside the seeds. To investigate more
in-depth the decomposition, derivative conversion curves were deduced, and the heat
involved during the pyrolysis or oxidation was determined through differential scanning
calorimetry (DSC). A simple thermodynamic model was also developed in Aspen Plus to
compare experimental outputs with thermodynamic predictions. Through the software,
pyrolysis and oxidation of the hydrochars were simulated and the outputs were compared
with experimental measures from DSC to understand if this modeling approach can be
used as a preliminary study of the heat of reactions.
Table 1. Grape seeds and hydrochar (HC) properties, from the literature [32] (C: carbon, H: hydrogen,
O: oxygen, N: nitrogen, S: sulfur, HHV: higher heating value).
m0 − m(t) v(t)
α(t) = = (1)
m0 − mf v∞
where m0 , m(t), and mf are the mass at time zero (dry), at a given time t, and final time,
respectively. This corresponds to the ratio between v(t), which represents the amount of
volatiles evolved up to a certain time t, and v∞ , which represents the amount of volatiles
that evolved up to the end of the process.
DSC was performed in a Mettler DSC20 calorimeter between 40 and 600 ◦ C, at a
heating rate of 10 ◦ C/min, under nitrogen or air at a flow rate of 100 mL/min. About 15 mg
of material was placed into a 40 μL aluminum crucible. The heat released/absorbed was
computed by integrating the measured heat flow over time in the 40–600 ◦ C range, while
the sample mass variation was recorded after every run.
149
Energies 2022, 15, 950
FTIR spectra of samples were acquired using PerkinElmer FT-IR spectrometer Spec-
trum One. Spectra were obtained over a wavenumber range of 4000–650 cm−1 . Dried
samples were ground to be compacted into a thin film and a flat top-plate Zinc Selenide
crystal was used for the detection.
where k0 represents the pre-exponential factor, E is the activation energy, and R is the
universal gas constant.
To deduce the kinetic parameters, f(α) is usually assumed in advance (in the so called
model-fitting methods) and then forced to fit experimental data [23]. In particular, non-
isothermal models usually adopt a constant heating rate β equal to dT/dt, allowing the
re-arrangement of Equation (3) into (4), which can be solved using thermogravimetric
data [23].
dα ∼ dα dt dα 1 k(T) k0 E
= = = f(α) = exp − f(α) (4)
dT dt dT dt β β β RT
Among model-fitting methods, DAEM assumes that pyrolysis can be described
through an infinite number of irreversible first-order parallel reactions, with different
rate parameters, which occur simultaneously [23,34]. In particular, the DAEM for a reaction
order equal to 1 can be obtained by re-arranging Equations (2)–(4) as:
∞
T
k0 E
1−α = exp − exp − dT f(E)dE (5)
0 T0 β RT
Equation (5) does not have an analytical solution and there are two families of methods
to solve it. One is referred to as distribution-fitting method and the other as isoconver-
sional methods. The distribution-fitting method assumes f(E) (like a Gaussian, Weibull, or
Boltzmann distribution) and forces to fit the TGA data by applying a certain numerical
method [23]. Meanwhile, isoconversional methods do not assume f(α), but adopt a series of
thermogravimetric data at different heating rates to directly compute the activation energy.
The reaction rate at a certain α is a function of temperature. Among these distribution-free
methods, the Miura–Maki integral method is a common one [35].
1 (E − E0 )2
f(E) = √ exp (6)
σ 2π 2σ2
150
Energies 2022, 15, 950
Applying the Coats–Redfern [36] and Fisher et al. [37] approximations, Equation (7)
is obtained.
∞ k0 RT2 E 1 ( E − E0 ) 2
1−α = exp − exp − √ exp dE (7)
0 βE RT σ 2π 2σ2
To solve Equation (7), the authors applied the simplification reported by Anthony
and Howard [38], which consists in keeping k0 constant and equal to 1.67 × 1013 s−1 , and
the minimization procedure proposed by Güneş and Güneş [39] that solves the integral
using the Simpson’s 1/3 rule. To avoid oscillations in the results, dE was kept constant
at 50 kJ/mol, while the extremes of integration were iteratively adjusted according to E0 .
Then, the best solution was calculated by iterating the solution of the integral for several
values of E0 and σ and minimizing h, i.e., the root mean quadratic error between the model
output and the experimental data:
2
∑nj=1 αTGA,j − αDAEM,j
h= (8)
n
where αTGA,j and αDAEM,j are the experimental and calculated extent of conversion, respec-
tively, and n is the number of available data points.
The code was developed in MATLAB and a direct search technique was adopted [39].
The code is reported in Supplementary Materials. Firstly, a larger grid procedure was used
for both E0 and σ, with a step size equal to 5. The first three values minimizing h were
obtained. Then, for these last three values, an iteration process using E0 ± 5 and σ ± 5 with
a step equal to 1 was set. Finally, final values were used to solve the integral of Equation (7)
and to calculate αDAEM .
the extent of conversion for the j-th reaction is expressed by Equation (12):
Δv ∼ k0 RT2 E
1− = exp − exp − (12)
Δv∞ βE RT
151
Energies 2022, 15, 950
where Δv and Δv∞ represent the amount of volatiles evolved and the effective volatile
content for the j-th reaction.
After some mathematical steps [35], and after imposing 1 − Δv/(Δv∞ ) = Φ(E, T) ∼ =
0.58, the representative equation of the Miura–Maki model is given by Equation (13).
β k0 R E
ln 2 = ln + 0.6075 − (13)
T E RT
Compared to the Gaussian DAEM, the approach of Miura [35] and Miura and Maki [33]
relies on the dependence of k0 on E. In particular, they proposed a simple procedure to
estimate f(E) and k0 , consisting of the following steps:
(1) measurement of α vs. T relationship for at least three heating rates;
(2) calculation of β/T2 at selected α values from the α vs. T relationship obtained in
(1) for each heating rate;
(3) plotting of ln (β/T2 ) vs. −1/(RT) at the selected α values, and determination of E
and k0 from the Arrhenius plot at different α values using the relationship represented by
Equation (13): E can be estimated by the slope, and k0 by the intercept of each line with the
y axis; and
(4) plotting the α value against the activation energy E, as calculated in (3), and
differentiating the α vs. E relationship to obtain f(E).
152
Energies 2022, 15, 950
GĮGW V
& &
&
&
&
7HPSHUDWXUH & 7HPSHUDWXUH &
(a) (b)
+&& &PLQ +&& &PLQ
&
&PLQ &PLQ
& &PLQ &PLQ
GĮGW V
& &
7HPSHUDWXUH & 7HPSHUDWXUH &
(c) (d)
Figure 2. Cont.
153
Energies 2022, 15, 950
+&& &PLQ
&PLQ
&PLQ
&
(e)
Figure 2. Derivative thermogravimetric (DTGA) curves: (a) during pyrolysis at 10 ◦ C/min; (b) during
oxidation at 10 ◦ C/min; (c) effect of the heating rate during pyrolysis of hydrochar 180 ◦ C; (d) effect
of the heating rate during pyrolysis of hydrochar 220 ◦ C; (e) effect of the heating rate during pyrolysis
of hydrochar 250 ◦ C. HC: hydrochars, dα/dt: derivative conversion curve of the extent of conversion
α with respect to time t.
Table 3. DTGA peak temperature and decomposition rate of pyrolysis and oxidation (10 ◦ C/min).
Pyrolysis Oxidation
DTGA Peak Grape HC HC HC Grape HC HC HC
Oil
Seeds 180 ◦ C 220 ◦ C 250 ◦ C Seeds 180 ◦ C 220 ◦ C 250 ◦ C
T (◦ C) - 280 - - - 280 - - -
1 dα/dt
- 5.9 × 10−4 - - - 4.4 × 10−4 - - -
(s−1 )
T (◦ C) - 343 353 358 - 346 334 336 -
2 dα/dt
- 8.2 × 10−4 1.0 × 10−3 1.1 × 10−3 - 7.4 × 10−4 6.6 × 10−4 5.7 × 10−4 -
(s−1 )
T (◦ C) 423 421 424 428 400 415 415 406 409
3 dα/dt
2.0 × 10−3 6.6 × 10−4 7.8 × 10−4 7.7 × 10−4 8.2 × 10−4 4.6 × 10−4 6.2 × 10−4 5.8 × 10−4 5.3 × 10−4
(s−1 )
T (◦ C) - - - - - 502 502 482 480–523
4 dα/dt
- - - - - 1.1 × 10−3 1.3 × 10−3 1.4 × 10−3 1.6–1.1 × 10−3
(s−1 )
Overall, both grape seeds and hydrochars mostly decompose between 200 and 500 ◦ C,
indicating that most of the volatile matter is removed in this range. Grape seeds present
three main decomposition peaks, occurring at 280, 343, and 421 ◦ C under a pyrolytic atmo-
sphere. These peaks can be attributed to the overlapped decomposition of hemicellulose,
lignin plus cellulose, and oil, which are the main constituents of grape seeds (around 7%
cellulose, 31% hemicellulose, 44% lignin, and 10–15% oil [41,42]). Indeed, it is well known
that under pyrolytic conditions, due to its amorphous and little polymerized structure,
hemicellulose is very reactive at low temperatures (in the range of 250–300 ◦ C [21,43]). In
the DTGA curves, this reactivity translates into the first decomposition peak. Meanwhile,
the second peak can be attributed to the overlapped decomposition of lignin and cellulose,
which both generally decompose in the 300–350 ◦ C range [43]. Considering the low content
of cellulose (around 7% [41,42]), lignin clearly dominates at this temperature. Beyond this
decomposition, lignin also contributes by forming a decomposition “baseline” to the profile,
overlapping the other components. Indeed, due to its high polymerized structure and
higher heterogeneity, it decomposes also (less intensively than at 300–350 ◦ C) in a broad
range of temperatures, from around 230◦ up to 600 ◦ C [43]. Therefore, during the pyrolysis
of grape seeds, its decomposition could overlap to that of the other constituents, forming a
“baseline” to the overall profile. Meanwhile, comparing the measured curve with that of
154
Energies 2022, 15, 950
grape seed oil from previous work by our group [28], it is clear that the peak occurring at
423 ◦ C can be attributed to the decomposition of the oil. Regarding the hydrochars, through
HTC, grape seeds undergo carbonization, decreasing their atomic O/C and H/C ratios
with a natural reduction of the volatile matter as the harshness of the process increases [32].
In DTGA curves, this translates in flattening the first decomposition peak at around 280 ◦ C,
which is indeed absent for all the hydrochars. Therefore, the authors can affirm that
hemicellulosic compounds are mostly degraded during HTC. Interestingly, the 250 ◦ C
hydrochar exhibits a broad and not intense decomposition between 200 and 300 ◦ C. Since
hemicellulose degrades during HTC, this reactivity can be explained by the presence of
re-polymerized compounds from the aqueous phase. Indeed, during HTC, sugar-derived
compounds dissolved in the aqueous phase (like 5-HMF) can undergo condensation and
re-polymerization, forming a solid phase called “secondary char” [44]. Beyond this, fatty
acids formed during the hydrolysis of the oil could also be embedded in the solid phase,
conferring to the sample an extra-reactivity at low temperatures. Both grape seeds and 180
and 220 ◦ C hydrochars show their highest reactivity at around 350 ◦ C, attributable to the
degradation of lignin and cellulosic derived compounds. Differences can be due to both the
heterogeneity of the feedstock and the synergistic behavior among the constituents, which
are present in different ratios in the various samples. Conversely, the 250 ◦ C hydrochar,
after the initial slow decomposition at 200–300 ◦ C, shows only one main decomposition
peak at 400 ◦ C. Therefore, while the 180 and 220 ◦ C hydrochars tend to follow the feedstock
profile (2nd and 3rd peaks), the 250 ◦ C hydrochar highly deviates from that, highlighting
the impact of the HTC operating temperature on the devolatilization profile. Overall, it is
important to highlight to positive effect of the HTC treatment: after HTC, curves tend to
shift towards the right region of the graph, demonstrating a higher thermal stability during
the degradation process.
Regarding oxidative conditions (Figure 2b), as expected by typical combustion profiles
of biomass fuels [20], all the profiles can be divided into two regions: a first phase during
which the feedstock devolatilizes to char, and a second phase where char oxidation occurs.
The first phase, between 250 and 450 ◦ C, resembles the pyrolytic behavior: hemicellulose,
lignin plus cellulose, and oil volatilize in the same temperature ranges as during pyrolysis.
Then, the second phase, 480–520 ◦ C, corresponds to the oxidation of the char formed during
the first phase. This phase is predominant with respect to the previous phase. As for
pyrolysis, the 180 and 220 ◦ C hydrochars resemble the behavior of grape seeds, exhibiting
three main DTGA peaks. Meanwhile, the 250 ◦ C hydrochar presents only two DTGA peaks,
demonstrating the harshness of severe HTC conditions on the feedstock structure and
composition. Oxidation and pyrolysis TGA curves where (1-α) is plotted vs. T are reported
in Supplementary Materials Figure S1.
Figure 2c–e show the effect of the heating rate on DTGA curves. The heating rate
does not affect the shape of the decomposition profile. Meanwhile, the DTGA peak moves
towards higher temperatures at increasing heating rates, which can be due to heat and
mass transfer phenomena and intrinsic devolatilization kinetics. Indeed, at higher rates, the
mismatch between the temperature in the furnace and the particle is higher, and therefore
there is a delay between the temperature measured and the decomposition stage [45]. In
addition, the dα/dt is higher at higher rates because, at a fixed α, dt is smaller. In addition,
higher rates cause a bigger gradient between the surface and core temperature of the
sample [45].
155
Energies 2022, 15, 950
3\URO\VLV 2[LGDWLRQ
5DZ
+HDW)ORZ :J
+HDW)ORZ :J
+&&
+&&
+&&
H[R
5DZ
+&&
+&&
+&& H[R
í
7HPSHUDWXUH & 7HPSHUDWXUH &
(a) (b)
Figure 3. Differential scanning calorimetry (DSC) curves of grape seeds and hydrochars (HC) at
10 ◦ C/min: (a) under inert and (b) oxidative atmosphere (> 0 exothermic, < 0 endothermic); per
grams of dry feedstock.
Table 4. Details of DSC curves: integrals, mass loss, and transition point from endothermic to
exothermic behavior; per kilograms of dry feedstock. The positive values of “Net energy” correspond
to exothermal behavior during thermal treatments.
Pyrolysis Oxidation
Raw HC 180 ◦ C HC 220 ◦ C HC 250 ◦ C Raw HC 180 ◦ C HC 220 ◦ C HC 250 ◦ C
Endothermic 1
118 90 67 23 98 54 31 19
(J/g)
Exothermic 1 (J/g) 705 587 585 598 4406 4928 6678 7533
Net energy 1 (J/g) 587 497 518 575 4309 4874 6647 7514
Mass loss 1 (%) 61.2 56.2 55.4 43.7 73.7 69.0 73.3 69.0
1 Integral and mass losses computed/measured in the range 40–600 ◦ C.
As for DTGA curves, samples show heat profiles with local peaks. Indeed, grape seeds
show three main peaks, at 280, 352, and 410–425 ◦ C, in correspondence with the degradation
of hemicellulose, lignin plus cellulose, and oil. The 180 and 220 ◦ C hydrochars show two
peaks, while the 250 ◦ C shows only one at 420 ◦ C. A comparison between DTGA and DSC
curves under pyrolytic atmosphere is reported in Supplementary Materials Figure S2.
As expected from the nature of the process, under oxidative conditions (Figure 3b),
the amount of heat released increases as the carbonization degree of the sample increases.
All the samples show a slight endothermic behavior at low temperature, justified by the
absence or low presence of oxidative reactions at this condition. Overall, the heat release
profile is broad and rises to the final observed temperature. Indeed, at 600 ◦ C, a certain
quantity of matter that undergoes oxidation is still present. Even if less sharp than under
an inert atmosphere, all the samples show a peak at 350–450 ◦ C that can be associated with
156
Energies 2022, 15, 950
the oxidation of lignin-like compounds. Grape seeds also show a small inflection at 280 ◦ C,
probably due to the oxidation of hemicellulosic compounds.
3.3. FTIR
Figure 4 reports the FTIR spectra of the samples. The deflection at 3300 cm−1 indicates
the stretching of the hydroxyl group, occurring generally in the range 3700–3200 cm−1 [47].
Increasing HTC temperature results in a decreasing intensity in O-H stretching, because
dehydration phenomena occur during HTC treatment. The peak at 3010 cm−1 represents
=C-H group stretch, which tends to gradually disappear in the hydrochars. This fact can
be related to double bonds present in the oil, which degrades, and it is even not present
in HC 250 ◦ C. Purnomo et al. [48] identified hydrogen bonded to the unsaturated carbon
chain (C-H), whereas peaks at lower wavelength represent attachment to the saturated
carbon chain. The peak at 1744 cm−1 indicates a typical bending of carbonyl group >C=O,
possibly in the form of aldehyde [5], and/or the ester group in glyceride compounds. It
tends to attenuate by increasing HTC treatment severity. Whereas grape seeds and HC
180 ◦ C still present a peak, the HC 220 ◦ C spectrum moves toward a lower wavelength. In
the HC 250 ◦ C spectrum the peak disappeared, confirming degradation and formation of
new compounds during the hydrothermal treatment. Furthermore, Diaz et al. [41] state
that peaks around 1030 cm−1 indicate C-O and C-O-C stretching, confirming that both
decarboxylation and dehydration occurred. The grape seed spectrum exhibits a shoulder
at a wavelength equal to 1655 cm−1 , attributed to stretching of double bond >C=C< [47],
that tends to flatten out in hydrochars. Unsaturated double-bound >C=C< could be related
to the presence of oil [6]—grape seed oil is mainly composed of linoleic acid [48,49], having
two double bonds C=C in position 9 and 12. In general, functional groups act a stretch
toward the region at a shorter wavelength.
2í+
7UDVPLWWDQFH
*UDSHVHHGV
+&& &í+
+&& & 2 &í2
+&&
Figure 4. Fourier transform infrared spectroscopy (FTIR) curves of grape seeds and hydrochars (HC).
157
Energies 2022, 15, 950
Table 5. Values of E0 and σ for the Gaussian model for the different substrates and heating rates; per
moles of dry feedstock. k0 is let constant and equal to 1.67 × 1013 s−1 .
σ 1 ◦ C/min - - 15 16 17
Pyrolysis
(kJ/mol) 3 ◦ C/min - - 14 15 16
10 ◦ C/min 16 11 15 15 16
1 ◦ C/min - - 0.083 0.090 0.089
h (-) 3 ◦ C/min - - 0.061 0.072 0.073
10 ◦ C/min 0.084 0.022 0.064 0.062 0.076
1 ◦ C/min - - 211 214 219
E0 (kJ/mol) 3 ◦ C/min - - 215 215 220
10 ◦ C/min 213 - 214 213 221
σ 1 ◦ C/min - - 14 15 15
Oxidation 3 ◦ C/min - - 15 14 15
(kJ/mol)
10 ◦ C/min 17 - 16 15 15
1 ◦ C/min - - 0.057 0.065 0.065
h (-) 3 ◦ C/min - - 0.071 0.066 0.066
10 ◦ C/min 0.102 - 0.084 0.077 0.077
Data demonstrate that for both pyrolysis and oxidation, the effect of the heating rate
on E0 and σ is negligible. Therefore, every substrate can be associated with one single pair
of E0 and σ, regardless of the heating rate, which agrees with previous studies performed
on other biomass substrates [28]. Under both inert and oxidative atmospheres, the Gaussian
DAEM adopted does not highlight the effects of both HTC and its severity on the starting
feedstock (their maximum relative difference is 3.7%). This is not surprising: the model
basis on the assumption of approximating the entire decomposition profile, which includes
more stages, through a one-step profile characterized by a Gaussian distribution. Therefore,
the final values of E0 and σ can be considered as a sort of weighted average among the
values of the single components (biomass constituents), i.e., hemicellulosic, lignin-like
plus cellulosic compounds, and oil. Probably, the difference in the relative composition
of the hydrochars is not enough to affect the final values. In particular, under an inert
atmosphere, E0 passes from 193 kJ/mol of grape seeds to 200–209 kJ/mol of hydrochars,
while σ ranges between 14 and 17 kJ/mol. Values agree with those reported in the literature
on the single components. For example, during pyrolysis, the activation energies usually
range in 80–116 kJ/mol for hemicellulose, 195–286 kJ/mol for cellulose, and 176–300 kJ/mol
for lignin [35,42]. The pyrolysis of oil alone has an E0 of 218 kJ/mol. With values between
211 and 215 kJ/mol, the activation energy of oxidation is the same for grape seeds and
hydrochars obtained at 180 and 220 ◦ C. Meanwhile, it slightly increases at 220 kJ/mol for
the 250 ◦ C hydrochar.
158
Energies 2022, 15, 950
Figure 5. 1-α vs. temperature: comparison between experimental data (dots) and predicted data
(lines) computed through the Gaussian model, during pyrolysis of different samples. (a) hydrochar
180 ◦ C; (b) hydrochar 220 ◦ C; (c) hydrochar 250 ◦ C.
159
Energies 2022, 15, 950
Figure 7. Miura–Maki diagrams of pyrolysis of: (a) hydrochar 180 ◦ C; (b) hydrochar 220 ◦ C;
(c) hydrochar 250 ◦ C.
Under pyrolysis, the fitting is satisfactory, with a R2 always higher than 0.95, except
for one outlier at α = 0.7 for the 220 ◦ C hydrochar. Results show that α hugely affects
the kinetic parameters. All the hydrochars present an E that increases with the degree
of conversion: it rapidly passes from 100–147 kJ/mol (at α = 0.1) to 371–386 kJ/mol in a
160
Energies 2022, 15, 950
range of α 0.6–0.7. This phenomenon can be explained by the progressive increase of the
degree of carbonization with temperature. As pyrolysis proceeds, the biomass undergoes
volatilization and is converted into a high-carbon content matrix. This char is more difficult
to degrade than the volatile matter and therefore causes a higher E. Generally, values are
considerably higher than those commonly found in literature for biomass substrates (for
example, for lignin it is 237–267 kJ/mol, and for pine wood it is 186–271 kJ/mol [27]),
which can be explained by the nature of the substrate. Meanwhile, k0 reaches its maximum
at an α of 0.5–0.6, with maximum values of 1024–27 for the 180 and 250 ◦ C hydrochars and
1038 for the 220 ◦ C one.
Figure 6b shows how E of different hydrochars vary during oxidation. For the 180 and
220 ◦ C hydrochars, E increases up to α = 0.3–0.5 to values of 357–379 kJ/mol. Meanwhile,
the 250 ◦ C hydrochar shows a very similar trend to the 220 ◦ C one up to an α of 0.35, and
then stabilizes at much lower values of 221–225 kJ/mol. The 180 ◦ C hydrochar shows higher
values of E at the beginning of the conversion. Therefore, the HTC severity decreases the
kinetic parameters of both volatile matter and char produced during the process. Similar
trends were observed by Bach et al. [20] on hydrochars produced from wood residues.
Meanwhile, pre-exponential factors k0 show a very similar trend to E (they reach their
maxima in correspondence of the maximum of α, arriving up to 1024–27 for the 180 and
220 ◦ C hydrochars, and 1014 for the 250 ◦ C one).
Table 6. Average values of E, σ, and k0 computed through the Gaussian and Miura–Maki models (k0
constant and equal to 1.67 × 1013 s−1 for the Gaussian model).
Hydrochar
Model Atmosphere Parameter Grape Seeds Oil
180 ◦ C 220 ◦ C 250 ◦ C
E (kJ/mol) 193 218 203 203 206
Pyrolysis
σ (kJ/mol) 16 11 15 15 16
Gaussian
E (kJ/mol) 213 - 213 214 220
Oxidation
σ (kJ/mol) 17 - 15 15 15
E (kJ/mol) - - 265 339 239
Pyrolysis
k0 (s−1 ) - - 7.6 × 1027 6.0 × 1038 9.8 × 1024
Miura-Maki
E (kJ/mol) - - 222 215 181
Oxidation
k0 (s−1 ) - - 3.4 × 1026 1.0 × 1026 2.9 × 1013
Due to its starting hypothesis, i.e., approximating the profile with a single Gaussian
curve, the Gaussian model does not highlight the effect of the severity of the HTC pro-
cess on E, which ranges between 193 and 206 kJ/mol for pyrolysis. This does not occur
for the Miura–Maki model, which assumes a distribution of E that varies on the entire
range leading to more diverse average values. Indeed, the Gaussian approach models a
decomposition involving several stages through a single decomposition stage character-
ized by a Gaussian profile of the activation energy. Therefore, this assumption “averages”
the profiles, canceling the differences among the samples obtained at different severities.
Meanwhile, the Miura–Maki model avoids this averaging since it does not impose a similar
strong assumption on the distribution.
Since the literature lacks the modeling of the kinetics of pyrolysis of hydrochar derived
from grape seeds, the comparison of the results obtained was possible considering typical
values of activation energy for decomposition of the main constituents of lignocellulosic
materials. In particular, cellulose decomposition activation energy ranges between 175
and 279 kJ/mol [21,27,50] hemicellulose is in the range 132–186 kJ/mol [27,51,52], whereas
161
Energies 2022, 15, 950
lignin ranges from 62 to 271 kJ/mol [27,50–52], with the broadest range. These values seem
to prove the consistency of the results obtained in the present work.
In general, the Gaussian model has the advantage of being valid over the entire range
of conversion, while the Miura–Maki model requires always that the temperature sequence
must be satisfied, an aspect that is not obvious (in this study this occurs only for α in the
range of 0.1–0.8). Overall, both the models suffer from a strong starting hypothesis: k0
fixed for the Gaussian model and a strong integral simplification in the Miura–Maki model
that can lead to an overestimation of the average values of the kinetics parameters [53].
To improve the prediction, the authors suggest extending the single Gaussian model
to a Multi Gaussian, in which the entire decomposition profile is divided into single
decomposition peaks, each one approximated by a Gaussian curve [21]. Indeed, the
single approach suits well homogeneous substrates (like the oil) characterized by a single
decomposition peak, but seems too simple for complex substrates like biomasses. As a
result, future works should include a comparison with other kinetics models.
Then, applying other DAEMs (like Kissinger–Akahira–Sunose method and the Coats–
Redfern method) on the same feedstock could help to validate the results. Apart from
contributing to understanding the mechanisms behind the decomposition, kinetic parame-
ters can help in technological design and optimization. Moving to a larger scale, integration
with other modeling techniques is necessary. Among these, it is worth mentioning models
that consider the effects of particle distribution and geometry on heat transfer phenomena
(like fractal models [54,55]), statistical models [56], and comprehensive computational
models [57].
162
Energies 2022, 15, 950
3\URO\VLV
3\URO\VLV
(QHUJ\ -J
(QHUJ\ -J
H[R
HQGR
í 5DZ
H[R 5DZ
+&&
+&&
+&&
+&& +&&
í +&&
&RDO
í
7HPSHUDWXUH & 7HPSHUDWXUH &
(a) (b)
2[LGDWLRQ 5DZ 2[LGDWLRQ
+&&
+&&
+&&
(QHUJ\ -J
(QHUJ\ -J
5DZ
+&&
+&&
H[R +&&
H[R
&RDO
7HPSHUDWXUH & 7HPSHUDWXUH &
(c) (d)
Figure 8. Energy released (> 0) or absorbed (< 0) by the samples computed from: (a) Aspen Plus
during pyrolysis; (b) DSC data during pyrolysis at 10 ◦ C/min; (c) Aspen Plus during oxidation;
(d) DSC data during oxidation at 10 ◦ C/min; per grams of dry feedstock.
Figure 9. Aspen Plus results: main products from pyrolysis of hydrochar 180 ◦ C.
Besides, Figure 8b shows the progressive integration of pyrolytic DSC data; comput-
ing the energy integral is necessary because DSC measures the instantaneous heat flow
absorbed or released (which depends on the “history” of the sample due to the thermal
program), while thermodynamics considers the energy absorbed or released when convert-
ing all the reactants into products at that determined temperature (and therefore the values
obtained from thermodynamics do not depend on the thermal history of the sample). Over-
163
Energies 2022, 15, 950
all, DSC data show a trend towards a cumulative exothermic behavior, whose final value
corresponds to the integral reported in Table 4. The different behavior concerning Aspen
data can be due to the nature of the thermodynamic approach. Indeed, real systems do not
behave as predicted by thermodynamics at low temperatures because of reaction kinetics
constraints. Moreover, as previously mentioned, thermodynamics does not consider the
“history” of what happened before a certain condition. Indeed, inside the DSC, the starting
material progressively undergoes pyrolysis, and the material at every time is the result of
what happened before (i.e., the thermal program). In addition, thermodynamics does not
consider typical “out of equilibrium” products, like tarry compounds, always present in
real systems. Therefore, the thermodynamic approach requires particular attention and
criticism before being used.
Regarding oxidation, both thermodynamic and DSC curves are clearly exothermic,
with values that progressively increase with temperature. Under both atmospheres, ab-
solute values differ between the Aspen and DSC approaches. This is due to the mass
involved: DSC curves measure the energy absorbed/released by an amount of material
that progressively decreases with time (the process is occurring), while Aspen predicts
the energy absorbed/released by the same mass at every temperature. Since the energy
is normalized by the amount of starting feedstock (and not the mass measured at every
temperature), the amount of heat computed through the DSC approach is smaller than
the Aspen one. Finally, it is interesting to note that in an oxidative environment, and
for the whole range of temperature investigated, the thermodynamic approach foresees
the complete oxidation of the biomass with the production essentially of CO2 , H2 O, and
residual ash (equal to about 2% of the biomass fed to the reactor, Table 1). Conversely, in an
oxidative environment in DSC, at the maximum temperature reached of 600 ◦ C, there is
still a substantial mass of non-oxidized carbon: the residual mass at 600 ◦ C is of the order
of 30% (Table 4). Through this critical comparison, the two approaches, thermodynamic
and DSC, turn out to actually be very distant from each other, which is an aspect worthy
of interest.
4. Conclusions
This work investigated the kinetic behavior during pyrolysis and oxidation of hy-
drochars derived from an agro-industrial residue (grape seeds). The topic was approached
through experimental techniques (TGA and DSC) that were used to develop two different
DAEMs for biomass decomposition. Indeed, there is a certain lack in the literature regard-
ing hydrochar decomposition, especially for the computation of the kinetic parameters.
Therefore, this work covers both the experimental aspects and the comparison of two
DAEMs using a critical approach. Thermogravimetric analysis and differential scanning
calorimetry highlighted the importance of the HTC severity on the decomposition profiles.
Indeed, hydrochars obtained at 180 and 220 ◦ C still decompose with two/three stages
(pyrolysis/oxidation) due to the presence of lignocellulosic constituents and oil that only
partially degraded during HTC. Meanwhile, the 250 ◦ C hydrochar shows a more stable
profile, with one/two decomposition peaks, highlighting that HTC temperature hugely af-
fected the structure of the biomass. Interestingly, all hydrochars show a certain exothermic
behavior during pyrolysis.
DAEMs predicted the kinetic parameters involved during the pyrolysis and oxida-
tion processes, i.e., the activation energy and the pre-exponential factor. The single-stage
Gaussian model does not highlight differences among the various hydrochars obtained
at different HTC temperatures, with average values of activation energy in the range
203–206 kJ/mol for pyrolysis and 213–220 kJ/mol for oxidation. Actually, the Gaussian
DAEM turned out to be unsatisfactory to model complex feedstock such as hydrochars
characterized by multi-decomposition peaks. Meanwhile, the Miura–Maki model, even if
not applicable over the entire decomposition region, enabled the determination of the dis-
tribution of activation energies: the more severe the HTC process, the lower the activation
energy values.
164
Energies 2022, 15, 950
Supplementary Materials: The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/en15030950/s1, Figure S1: (1-α) at different temperatures of
feedstock during: (a) pyrolysis; (b) oxidation. Heating rate: 10 ◦ C/min.; Figure S2: Comparison
between DTGA and DSC curves during pyrolysis at 10◦ C/min of: (a) grape seeds and hydrochar
180 ◦ C; (b) hydrochars 220 and 250 ◦ C; Figure S3: Comparison between experimental and predicted
data (1-α vs temperature) computed through the Gaussian model, during oxidation of different
samples: (a) hydrochar 180 ◦ C; (b) hydrochar 220 ◦ C; (c) hydrochar 250 ◦ C; Table S1: Details of
Miura-Maki model, pyrolysis (on a dry basis); Table S2: Details of Miura-Maki model, oxidation (on a
dry basis); Gaussian DAEM – MATLAB code.
Author Contributions: G.G.: methodology, software, data curation, writing—original draft prepara-
tion; G.I.: data curation, visualization, writing—original draft preparation; L.F. (Luca Fambri): exper-
imental planning, writing—review and editing, supervision, project administration; L.F. (Luca Fiori):
writing—review and editing, supervision, project administration. All authors have read and agreed
to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Informed consent was obtained from all subjects involved in the study.
Data Availability Statement: Not applicable.
Acknowledgments: All the authors acknowledge Claudia Gavazza for thermal analysis. G.I. ac-
knowledges the financial support provided by the ERICSOL project of the University of Trento, which
partially covered her Ph.D. scholarship. All individuals included in this section have consented to
the acknowledgement.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Li, J.; Pan, L.; Suvarna, M.; Tong, Y.W.; Wang, X. Fuel Properties of Hydrochar and Pyrochar: Prediction and Exploration with
Machine Learning. Appl. Energy 2020, 269, 115166. [CrossRef]
2. Ahmed, M.; Andreottola, G.; Elagroudy, S.; Negm, M.S.; Fiori, L. Coupling Hydrothermal Carbonization and Anaerobic Digestion
for Sewage Digestate Management: Influence of Hydrothermal Treatment Time on Dewaterability and Bio-Methane Production.
J. Environ. Manage. 2021, 281, 111910. [CrossRef] [PubMed]
3. De Medeiros, A.D.M.; Da Silva Junior, C.J.G.; De Amorim, J.D.P.; Do Nascimento, H.A.; Converti, A.; De Santana Costa, A.F.;
Sarubbo, L.A. Biocellulose for Treatment of Wastewaters Generated by Energy Consuming Industries: A Review. Energies 2021,
14, 1–18. [CrossRef]
4. Yu, C.; Ren, S.; Wang, G.; Xu, J.; Teng, H.; Li, T.; Huang, C.; Wang, C.; Yu, C.; Ren, S.; et al. Kinetic Analysis and Modeling of Maize
Straw Hydrochar Combustion Using a Multi- Gaussian-Distributed Activation Energy Model Kinetic Analysis and Modeling of
Maize Straw Hydrochar Combustion Using a Multi-Gaussian-Distributed Activation. Energy Model. 2021, 28, 611–620.
5. Zhang, N.; Wang, G.; Zhang, J.; Ning, X.; Li, Y.; Liang, W.; Wang, C. Study on Co-Combustion Characteristics of Hydrochar and
Anthracite Coal. J. Energy Inst. 2020, 93, 1125–1137. [CrossRef]
6. Chen, Z.; Hu, M.; Zhu, X.; Guo, D.; Liu, S.; Hu, Z.; Xiao, B.; Wang, J.; Laghari, M. Characteristics and Kinetic Study on Pyrolysis of
Five Lignocellulosic Biomass via Thermogravimetric Analysis. Bioresour. Technol. 2015, 192, 441–450. [CrossRef]
7. Chen, J.; Wang, Y.; Lang, X.; Ren, X.; Fan, S. Evaluation of Agricultural Residues Pyrolysis under Non-Isothermal Conditions:
Thermal Behaviors, Kinetics, and Thermodynamics. Bioresour. Technol. 2017, 241, 340–348. [CrossRef]
8. Ferrentino, R.; Merzari, F.; Fiori, L.; Andreottola, G. Coupling Hydrothermal Carbonization with Anaerobic Digestion for Sewage
Sludge Treatment: Influence of HTC Liquor and Hydrochar on Biomethane Production. Energies 2020, 13, 6262. [CrossRef]
9. Funke, A.; Ziegler, F. Hydrothermal Carbonization of Biomass: A Summary and Discussion of Chemical Mechanisms for Process
Engineering. Biofuels, Bioprod. Biorefining 2010, 4, 160–177. [CrossRef]
10. Olszewski, M.P.; Arauzo, P.J.; Wadrzyk,
˛ M.; Kruse, A. Py-GC-MS of Hydrochars Produced from Brewer’s Spent Grains. J. Anal.
Appl. Pyrolysis 2019, 140, 255–263. [CrossRef]
11. Ischia, G.; Cazzanelli, M.; Fiori, L.; Orlandi, M.; Miotello, A. Exothermicity of Hydrothermal Carbonization: Determination of
Heat Profile and Enthalpy of Reaction via High-Pressure Differential Scanning Calorimetry. Fuel 2022, 310, 122312. [CrossRef]
12. Ischia, G.; Fiori, L. Hydrothermal Carbonization of Organic Waste and Biomass: A Review on Process, Reactor, and Plant
Modeling. Waste and Biomass Valorization 2021, 12, 2797–2824. [CrossRef]
13. Scrinzi, D.; Andreottola, G.; Fiori, L. Composting Hydrochar-OFMSW Digestate Mixtures: Design of Bioreactors and Preliminary
Experimental Results. Appl. Sci. 2021, 11, 1–15. [CrossRef]
165
Energies 2022, 15, 950
14. Lin, J.C.; Mariuzza, D.; Volpe, M.; Fiori, L.; Ceylan, S.; Goldfarb, J.L. Integrated Thermochemical Conversion Process for Valorizing
Mixed Agricultural and Dairy Waste to Nutrient-Enriched Biochars and Biofuels. Bioresour. Technol. 2021, 328, 124765. [CrossRef]
15. Hu, Q.; Yang, H.; Xu, H.; Wu, Z.; Lim, C.J.; Bi, X.T.; Chen, H. Thermal Behavior and Reaction Kinetics Analysis of Pyrolysis and
Subsequent In-Situ Gasification of Torrefied Biomass Pellets. Energy Convers. Manag. 2018, 161, 205–214. [CrossRef]
16. Román, S.; Libra, J.; Berge, N.; Sabio, E.; Ro, K.; Li, L.; Ledesma, B.; Alvarez, A.; Bae, S. Hydrothermal Carbonization: Modeling,
Final Properties Design and Applications: A Review. Energies 2018, 11, 216. [CrossRef]
17. Sharma, H.B.; Panigrahi, S.; Dubey, B.K. Hydrothermal Carbonization of Yard Waste for Solid Bio-Fuel Production: Study on
Combustion Kinetic, Energy Properties, Grindability and Flowability of Hydrochar. Waste Manag. 2019, 91, 108–119. [CrossRef]
18. Islam, M.A.; Kabir, G.; Asif, M.; Hameed, B.H. Combustion Kinetics of Hydrochar Produced from Hydrothermal Carbonisation
of Karanj (Pongamia Pinnata) Fruit Hulls via Thermogravimetric Analysis. Bioresour. Technol. 2015, 194, 14–20. [CrossRef]
19. He, C.; Giannis, A.; Wang, J.-Y. Conversion of Sewage Sludge to Clean Solid Fuel Using Hydrothermal Carbonization: Hydrochar
Fuel Characteristics and Combustion Behavior. Appl. Energy 2013, 111, 257–266. [CrossRef]
20. Bach, Q.V.; Tran, K.Q.; Skreiberg, Ø. Combustion Kinetics of Wet-Torrefied Forest Residues Using the Distributed Activation
Energy Model (DAEM). Appl. Energy 2017, 185, 1059–1066. [CrossRef]
21. Zhang, J.; Chen, T.; Wu, J.; Wu, J. Multi-Gaussian-DAEM-Reaction Model for Thermal Decompositions of Cellulose, Hemicellulose
and Lignin: Comparison of N2 and CO2 Atmosphere. Bioresour. Technol. 2014, 166, 87–95. [CrossRef]
22. Vand, V. A Theory of the Irreversible Electrical Resistance Changes of Metallic Films Evaporated in Vacuum. Proc. Phys. Soc. 1943,
55, 222–246. [CrossRef]
23. Wang, S.; Dai, G.; Yang, H.; Luo, Z. Lignocellulosic Biomass Pyrolysis Mechanism: A State-of-the-Art Review. Prog. Energy
Combust. Sci. 2017, 62, 33–86. [CrossRef]
24. Radojević, M.; Janković, B.; Jovanović, V.; Stojiljković, D.; Manić, N. Comparative Pyrolysis Kinetics of Various Biomasses Based
on Model-Free and DAEM Approaches Improved with Numerical Optimization Procedure. PLoS ONE 2018, 13, 1–25. [CrossRef]
25. Hameed, S.; Sharma, A.; Pareek, V.; Wu, H.; Yu, Y. A Review on Biomass Pyrolysis Models: Kinetic, Network and Mechanistic
Models. Biomass Bioenergy 2019, 123, 104–122. [CrossRef]
26. White, J.E.; Catallo, W.J.; Legendre, B.L. Biomass Pyrolysis Kinetics: A Comparative Critical Review with Relevant Agricultural
Residue Case Studies. J. Anal. Appl. Pyrolysis 2011, 91, 1–33. [CrossRef]
27. Cai, J.; Wu, W.; Liu, R.; Huber, G.W. A Distributed Activation Energy Model for the Pyrolysis of Lignocellulosic Biomass. Green
Chem. 2013, 15, 1331–1340. [CrossRef]
28. Fiori, L.; Valbusa, M.; Lorenzi, D.; Fambri, L. Modeling of the Devolatilization Kinetics during Pyrolysis of Grape Residues.
Bioresour. Technol. 2012, 103, 389–397. [CrossRef]
29. Basso, D.; Weiss-Hortala, E.; Patuzzi, F.; Baratieri, M.; Fiori, L. In Deep Analysis on the Behavior of Grape Marc Constituents
during Hydrothermal Carbonization. Energies 2018, 11, 1379. [CrossRef]
30. Cai, J.M.; Bi, L.S. Kinetic Analysis of Wheat Straw Pyrolysis Using Isoconversional Methods. J. Therm. Anal. Calorim. 2009, 98, 325.
[CrossRef]
31. Branca, C.; Albano, A.; Di Blasi, C. Critical Evaluation of Global Mechanisms of Wood Devolatilization. Thermochim. Acta 2005,
429, 133–141. [CrossRef]
32. Ischia, G.; Orlandi, M.; Fendrich, M.A.; Bettonte, M.; Merzari, F.; Miotello, A.; Fiori, L. Realization of a Solar Hydrothermal
Carbonization Reactor: A Zero-Energy Technology for Waste Biomass Valorization. J. Environ. Manage. 2020. [CrossRef] [PubMed]
33. Miura, K.; Maki, T. A Simple Method for Estimating f(E) and K0(E) in the Distributed Activation Energy Model. Energy and Fuels
1998, 12, 864–869. [CrossRef]
34. Gao, L.; Volpe, M.; Lucian, M.; Fiori, L.; Goldfarb, J.L. Does Hydrothermal Carbonization as a Biomass Pretreatment Reduce Fuel
Segregation of Coal-Biomass Blends during Oxidation? Energy Convers. Manag. 2019, 181, 93–104. [CrossRef]
35. Miura, K. A New and Simple Method to Estimate f(E) and K0(E) in the Distributed Activation Energy Model from Three Sets of
Experimental Data. Energy and Fuels 1995, 9, 302–307. [CrossRef]
36. Coats, A.W.; Redfern, J.P. Kinetic Parameters from Thermogravimetric Data. Nature 1964, 201, 68–69. [CrossRef]
37. Fischer, P.E.; Jou, C.S.; Gokalgandhi, S.S. Obtaining the Kinetic Parameters from Thermogravimetry Using a Modified Coats and
Redfern Technique. Ind. Eng. Chem. Res. 1987, 26, 1037–1040. [CrossRef]
38. Anthony, D.B.; Howard, J.B. Coal Devolatilization and Hydrogastification. AIChE J. 1976, 22, 625–656. [CrossRef]
39. Güneş, M.; Güneş, S. A Direct Search Method for Determination of DAEM Kinetic Parameters from Nonisothermal TGA Data
(Note). Appl. Math. Comput. 2002, 130, 619–628. [CrossRef]
40. Ischia, G.; Fiori, L.; Gao, L.; Goldfarb, J.L. Valorizing Municipal Solid Waste via Integrating Hydrothermal Carbonization and
Downstream Extraction for Biofuel Production. J. Clean. Prod. 2021, 289, 125781. [CrossRef]
41. Spanghero, M.; Salem, A.Z.M.; Robinson, P.H. Chemical Composition, Including Secondary Metabolites, and Rumen Fer-
mentability of Seeds and Pulp of Californian (USA) and Italian Grape Pomaces. Anim. Feed Sci. Technol. 2009, 152, 243–255.
[CrossRef]
42. Moldes, D.; Gallego, P.P.; Rodríguez Couto, S.; Sanromán, A. Grape Seeds: The Best Lignocellulosic Waste to Produce Laccase by
Solid State Cultures of Trametes Hirsuta. Biotechnol. Lett. 2003, 25, 491–495. [CrossRef]
43. Hubble, A.H.; Goldfarb, J.L. Synergistic Effects of Biomass Building Blocks on Pyrolysis Gas and Bio-Oil Formation. J. Anal. Appl.
Pyrolysis 2021, 156, 105100. [CrossRef]
166
Energies 2022, 15, 950
44. Volpe, M.; Fiori, L. From Olive Waste to Solid Biofuel through Hydrothermal Carbonisation: The Role of Temperature and Solid
Load on Secondary Char Formation and Hydrochar Energy Properties. J. Anal. Appl. Pyrolysis 2017, 124, 63–72. [CrossRef]
45. Mackul’ak, T.; Prousek, J.; Olejníková, P.; Bodík, I. Slovak Society of Chemical Engineering Institute of Chemical and Environmental
Engineering Slovak University of Technology in Bratislava. Direct 2010, 1407–1412.
46. Volpe, M.; Goldfarb, J.L.; Fiori, L. Hydrothermal Carbonization of Opuntia Ficus-Indica Cladodes: Role of Process Parameters on
Hydrochar Properties. Bioresour. Technol. 2018, 247, 310–318. [CrossRef]
47. Guo, S.; Dong, X.; Wu, T.; Shi, F.; Zhu, C. Characteristic Evolution of Hydrochar from Hydrothermal Carbonization of Corn Stalk.
J. Anal. Appl. Pyrolysis 2015, 116, 1–9. [CrossRef]
48. Purnomo, C.W.; Castello, D.; Fiori, L. Granular Activated Carbon from Grape Seeds Hydrothermal Char. Appl. Sci. 2018, 8, 1–16.
[CrossRef]
49. Diaz, E.; Manzano, F.; Villamil, J.A.; Rodriguez, J.; Mohedano, A. Low-Cost Activated Grape Seed-Derived Hydrochar through
Hydrothermal Carbonization and Chemical Activation for Sulfamethoxazole Adsorption. Appl. Sci. 2019, 9, 5127. [CrossRef]
50. Jiang, G.; Nowakowski, D.J.; Bridgwater, A.V. A Systematic Study of the Kinetics of Lignin Pyrolysis. Thermochim. Acta 2010, 498,
61–66. [CrossRef]
51. Wang, S.; Ru, B.; Lin, H.; Sun, W.; Luo, Z. Pyrolysis Behaviors of Four Lignin Polymers Isolated from the Same Pine Wood.
Bioresour. Technol. 2015, 182, 120–127. [CrossRef] [PubMed]
52. Gomez, I.J.; Arnaiz, B.; Cacioppo, M.; Arcudi, F.; Prato, M. Nitrogen-Doped Carbon Nanodots for Bioimaging and Delivery of
Paclitaxel. J. Mater. Chem. B 2018, 6, 5540–5548. [CrossRef] [PubMed]
53. Cai, J.; Li, T.; Liu, R. A Critical Study of the Miura-Maki Integral Method for the Estimation of the Kinetic Parameters of the
Distributed Activation Energy Model. Bioresour. Technol. 2011, 102, 3894–3899. [CrossRef] [PubMed]
54. Xu, P.; Yu, B. The Scaling Laws of Transport Properties for Fractal-like Tree Networks. J. Appl. Phys. 2006, 100, 104906. [CrossRef]
55. Liang, M.; Fu, C.; Xiao, B.; Luo, L.; Wang, Z. A Fractal Study for the Effective Electrolyte Diffusion through Charged Porous
Media. Int. J. Heat Mass Transf. 2019, 137, 365–371. [CrossRef]
56. Mäkelä, M.; Yoshikawa, K. Simulating Hydrothermal Treatment of Sludge within a Pulp and Paper Mill. Appl. Energy 2016, 173,
177–183. [CrossRef]
57. Álvarez-Murillo, A.; Sabio, E.; Ledesma, B.; Román, S.; González-García, C.M. Generation of Biofuel from Hydrothermal
Carbonization of Cellulose. Kinetics Modelling. Energy 2016, 94, 600–608. [CrossRef]
167
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
www.mdpi.com
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are
solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s).
MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from
any ideas, methods, instructions or products referred to in the content.
Academic Open
Access Publishing