17

Production of biofuels via hydrothermal conversion

P. Biller1,2,  and A.B. Ross1     1University of Leeds, Leeds, Yorkshire, United Kingdom     2Aarhus University, Aarhus, Denmark

Abstract

Hydrothermal processing has evolved as an alternative processing technology for wet biomass and waste materials in recent years. Using hot-compressed water as a reaction medium at temperatures of 200–500°C, materials with increased energy density can be obtained. The technology is particularly suited for wet and waste materials as drying of the feedstock is not required. Hydrothermal processing is divided into three separate areas depending on reaction severity: hydrothermal carbonization (HTC, 180–280°C), hydrothermal liquefaction (HTL, 280–375°C), and hydrothermal gasification (HTG, >350°C). Each of these hydrothermal routes results in energy densification by removal of oxygen to produce hydrochar (HTC), biocrude (HTL), or syngas (HTG). The process chemistry and reactions in hydrothermal media are described for each process. Suitable feedstocks and their considerations are reviewed as the quality of targeted biofuel is a function of feedstock and operating conditions. The quality of hydrochar influences its uses as a solid fuel while biocrude quality affects its use as a liquid fuel and feedstock for upgrading to drop-in replacement fuels, while HTG produces a syngas rich in either H2 or CH4. Hydrothermal processing results in a process water at all temperatures, typical decomposition products, treatments, and uses of the water byproduct are discussed. Advances in reactor design and scale-up efforts to demonstration and industrial scales are reviewed for each technology. An assessment is made of the current state of technology and further areas of research are discussed.

Keywords

Biocoal; Biocrude; Carbonization; Gasification; Hydrochar; Hydrogen; Hydrothermal; Liquefaction; Methane; Sub-critical; Supercritical; Waste

17.1. Introduction

Hydrothermal processing involves heating aqueous slurries of biomass or organic wastes at elevated pressures to produce an energy carrier with increased energy density and is essentially similar to how the earth has produced fossil fuels over millions of years. All fossil fuel reserves have been created by the transformation of organic matter under pressure and heat over long periods of time. The process of applying high pressures and temperatures to organic matter in modern hydrothermal processing is often described as accelerating nature's natural pathways to form a renewable fossil fuel.
Hydrothermal processing is divided into three separate processes, depending on the severity of the operating conditions. At temperatures between 180 and 250°C, it is known as hydrothermal carbonization (HTC). The main product is a hydrochar which has a similar property to that of a low-rank coal. At intermediate temperature ranges between 250 and 375°C, the process is known as hydrothermal liquefaction (HTL), resulting in the production of a liquid fuel known as biocrude. Biocrude can be upgraded to the whole distillate range of petroleum-derived products. At higher temperatures above 375°C, gasification reactions start to dominate and the process is known as hydrothermal gasification (HTG) or supercritical water gasification (SCWG), resulting in the production of a syngas. The produced gas is not the same as conventional syngas from gasification, which is comprised of hydrogen and carbon monoxide. Nevertheless, the gas is referred to as syngas but is typically high in either hydrogen or methane with carbon dioxide also present. The overall aim in each case is to generate a biofuel with a higher energy density from the original feedstock.
One of the main advantages of hydrothermal processing technologies is that it is able to process wet feed. The feedstock does not require drying as the technology typically can handle slurries of biomass and waste with total solids of 10–30%. This is one reason why considerable attention has emerged in hydrothermal processing of microalgae for which traditional biofuel production pathways such as lipid extraction and transesterification to biodiesel require a drying step. This step can account to as much as 25% of the energy contained in the microalgae (Xu et al., 2011). While there have been considerable research efforts in the hydrothermal processing of algal biomass, other wet biomass feedstocks are also suitable for hydrothermal processing. Feedstock with high moisture and high ash content are particularly suitable for hydrothermal processing and include feedstocks such as AD digestate, manures, sewege sludge, dried distillers' grains with solubles (DDGS), food wastes, and municipal wastes.
The first reported experiments on hydrothermal processing were performed in 1913 by Bergius who applied hydrothermal carbonization to convert cellulose into a coal-like material (Specht, 1913). Hydrothermal gasification of biomass was first reported in the USA, at the Massachusetts Institute of Technology (Elliott, 2011). Similar technology was developed in Sweden for upgrading peat to coal-like fuels but more recently the focus has been on conversion of biomass to fuel products. Some of the earliest work on hydrothermal processing was performed at the Pittsburgh Energy Research Center in the 1970s and 1980s. Other early research in HTL was performed at the Royal Institute of Technology, Stockholm, the University of Arizona, and the University of Toronto (Elliott, 2011). HTL was extensively researched in the 1980s by Shell who developed a route for converting high-carbohydrate feedstock to liquid biofuels called the hydrothermal upgrading process (HTU®) (Frans and Jaap, 2015). Despite promising research, the project was shelved in 1988 mainly due to the very low crude oil price. The process was later picked up in 1997 and successfully tackled some of the challenges identified such as pumping of slurries and feed pretreatment. A pilot plant (100 kg/h) was constructed and the project ran until 2000 after demonstrating prolonged continuous operation in a 3-week investigation (Frans and Jaap, 2015). The earliest reported HTL of microalgae was carried out in the early 1990s at the National Institute for Resources and Environment in Tsubaka, Japan. The group led by Professor Minowa laid the foundations of microalgae HTL research (Inoue et al., 1994; Dote et al., 1994).
Academic research into hydrothermal processing increased around 2005, although the majority of this work has focused on algal biomass feedstocks. Hydrothermal carbonization was revived by Titirici et al. (2007), and has since received growing interest in academia and industry.
In this chapter, we aim to provide an overview of the status of hydrothermal processing technologies and recent research. The chapter includes a description of associated process chemistry including a description of how the properties of water change when heated under pressure and the reaction steps in each of the hydrothermal routes. The impact and influence of different feedstocks on process operation and product distribution and composition is reviewed. Recent advances in reactor design, product upgrading, commercialization and techno-economic and lifecycle analysis are presented before conclusions are drawn on the status of the technology.

17.2. Process chemistry

17.2.1. Hot compressed water

Hydrothermal processing has several advantages over other thermochemical processing routes of which the ability to utilize wet feedstock is the most obvious. The chemical and physical transformation of biomass takes place in water at elevated temperature and pressure between 200 and 600°C and 5–40 MPa respectively. The water is kept in the liquid state by operating at or above the saturation point and therefore largely minimizes the enthalpy change associated with the latent heat of vaporization of water. The large enthalpy change associated with the latent heat (2260 J/g) means heating water to 100°C under pressure requires six times less energy than generating steam at 100°C. The energy penalty of the latent heat of vaporization is avoided in all of the hydrothermal processing pathways. Conversion technologies that require predrying of feed, such as pyrolysis, need to overcome the latent heat of vaporization to remove water prior to processing.
An advantage of water as a reaction medium is that it is ecologically safe, cheap, and readily available but it also has some unique properties when heated under pressure. The hydrogen bonds are weakened, resulting in a change in dielectric constant, acidity, and polarity, each of which can increase opportunities for water to take part in chemical reactions. This change in properties allows water to act as a catalyst, lowering activation energies and allowing reaction pathways which would not occur at ambient conditions. Water is consumed by hydrolysis reactions and formed by dehydration reactions. It is also consumed in the water gas shift reaction which is active at higher temperatures. The critical point of water is at 374°C and 22.1 MPa, below this point the vapor pressure curve separates the liquid from the gaseous phase and the conditions are described as subcritical. Approaching the critical point, the densities of the two phases become more and more alike, and finally become identical at the critical point (Peterson et al., 2008). Above this point, the density of supercritical water is interchangeable without any phase transitions over a wide range of conditions. Depending where in the phase diagram the process conditions take place determines whether HTC, HTL, or HTG reaction conditions are favored, as can be seen in Fig. 17.1.
image
Figure 17.1 Hydrothermal processing conditions in the water phase diagram. Data from Perry, R.H., Green, D.W., 1997. Perry's Chemical Engineers' Handbook, seventh ed., McGraw-Hill.
At ambient conditions, the miscibility of water for hydrocarbons and gases is poor but it is a good solvent for salts due to its high dielectric constant of 78.5 (Kruse and Dinjus, 2007). Just below the critical point, the miscibility of hydrocarbons is improved as the dielectric constant reduces to levels equivalent to solvents such as dichloromethane, decreasing further in the supercritical region. The change in dielectric constant with temperature can be seen in Fig. 17.2 which shows the effect of temperature at constant pressure (30 MPa on the properties of water). Even under HTC conditions (250°C), the dielectric constant has more than halved, increasing the solubility of organics and opening new reaction pathways. The reaction rates in hydrothermal media can be finely tuned by controlling the temperature and pressure influencing the activation energy of reactions. Above the supercritical point of water, the miscibility of hydrocarbons and gases is very high, increasing the rate of reactions. Below the supercritical point, miscibility of these compounds is not complete but is still increased compared to ambient conditions. When the process products cool back down to ambient conditions, the water and the organic fraction will separate once more; this makes distillation or other costly separation techniques unnecessary.
At high temperature and pressure below the supercritical point, the ionic product (pH) is up to three orders of magnitude higher than under ambient conditions and is plotted in Fig. 17.2. The high ionic product supports acid or base catalyzed reactions and can act as an acid/base catalyst precursor because of the relatively high concentrations of H3O+ and OH ions from the self-dissociation of water (Kruse and Dinjus, 2007). The advantage of this is that the addition of acid or base catalysts can be avoided. The concentration of ions is at its maximum at 275°C, representing an optimum temperature for acid/base catalyzed reactions. Above 350°C, pH increases rapidly by five orders of magnitude or more (Kruse and Dinjus, 2007). The change in density of water at 30 MPa with temperature is also plotted in Fig. 17.2 and indicates that the most dramatic change takes place in the region of the critical point (375°C). Between 300 and 450°C, the density at 30 MPa changes from a liquid-like 750 kg/m3 to a gas-like 150 kg/m3, despite this dramatic change, there is no phase change taking place. This change in density directly correlates with properties such as solvation power, degree of hydrogen bonding, polarity, dielectric strength, diffusivity, and viscosity (Peterson et al., 2008).
image
Figure 17.2 Density (Wagner and Pruss, 2002), static dielectric constant (Uematsu and Frank, 1980) at 30 MPa and pH (Bandura and Lvov, 2006) of water at 25 MPa.

17.2.2. Hydrothermal reactions

Hydrothermal carbonization is the mildest of the three hydrothermal routes, operating at a process temperature ranging from 180 to 250°C and long residence times of typically several hours. The main product from HTC is a coal like hydrochar or biocoal, which is more energy-dense, more easily friable, and more hydrophobic than the starting material. This is achieved by reducing the oxygen and hydrogen content of the feed (described by the molecular O/C and H/C ratio), destroying the colloidal structures, and reducing the hydrophilic functional groups. The other products in HTC include a process water containing polar organic compounds and mineral matter and a gaseous fraction.
The water provides a medium for a complex series of reactions which involve removal of hydroxyl groups through dehydration, removal of carboxyl and carbonyl groups though decarboxylation, and cleavage of many ester and ether bonds through hydrolysis (Funke and Ziegler, 2010). The different steps in HTC are summarized in Fig. 17.3.
The initial step is thought to be the hydrolysis of carbohydrates and proteins which consumes water. Hemicellulose appears to degrade first, at temperatures generally below 200°C, cellulose degrades at 200–230°C and lignin degrades between 220 and 260°C (Pastor-Villegas et al., 2006; Libra et al., 2011; Reza et al., 2013). The lignin is thought to remain relatively intact at lower temperatures but begins to degrade further above 250°C. The soluble hydrolysis products undergo further dehydration reactions yielding water and decarboxylation reactions yielding CO2. These compounds undergo condensation and polymerization to larger molecules which undergo further aromatization (Funke and Ziegler, 2010; Kruse et al., 2013).
For lignocellulosic biomass, the net result is an increase in aromaticity with phenolic structures derived from the dehydration of lignin combined with aromatization resulting from carbonization of carbohydrates. Condensation and polymerization of the fragments can also form “humic-acid-like” material and “bitumen-like” material which tend to readhere to the hydrochar (Funke and Ziegler, 2010). The reaction pathways involved and composition and yield of products are strongly dependent on the type and composition of the feed (Funke and Ziegler, 2010) and are discussed in more detail in Section 17.3. The proportions of the biochemical components also influence the degradation pathway and operating temperature. In order to produce a coal-like product using hydrothermal carbonization, a reaction temperature of between 200 and 250°C is typically used (Funke and Ziegler, 2010).
When the temperature is increased to 250–375°C, hydrothermal liquefaction reactions start to dominate. During liquefaction, the biomass is decomposed to smaller molecules which are reactive and can repolymerize into hydrophobic, oily compounds (Zhang et al., 2010). The main reaction steps during liquefaction have been summarized by Garcia Alba et al. (2011) as follows:
image
Figure 17.3 Summary of the reaction steps in the different hydrothermal processes.
1. Hydrolysis of macromolecules into smaller fragments;
2. Conversion by dehydration and decarboxylation, into smaller compounds;
3. Rearrangement via condensation, cyclization, and polymerization producing new larger, hydrophobic macromolecules.
The products from hydrothermal liquefaction consist of a biocrude fraction, a water fraction, a gaseous fraction, and a solid residue fraction. The majority of the inorganic material is concentrated in the solid residue and process water. The product distribution is largely affected by the biochemical composition of the feedstock. Lipids, for example, are almost entirely fractionated to the biocrude as fatty acids and alkanes. Carbohydrates on the other hand tend to form char. This can be avoided by keeping the water alkali to increase the biocrude yields. The processing of lignocellulosics generally requires an alkali catalyst to achieve satisfactory conversion to biocrude and to avoid excessive char formation. In HTC, the water is acidic and so favors the formation of char. Proteins are converted readily to biocrude via a stepwise reaction into amino acids and finally ammonium, increasing the pH further and promoting biocrude formation. This is why catalysts are not required in HTL when processing feedstock high in protein. A series of interactions between different biochemical components in the biomass can occur. For example, feedstocks containing protein and lipids can result in formation of fatty acid amides. Carbohydrates and protein fragments can react via the Maillard reaction resulting in a range of N compounds. The different steps in HTL are summarized in Fig. 17.3.
Hydrothermal gasification occurs in the higher-temperature region above the supercritical point where water is in the supercritical state. The primary product now becomes a syngas high in combustible gases such as H2, CH4, CO, and light hydrocarbons (C2–C3); however CO2 is also produced. The energy requirements to maintain the water in its supercritical state are higher than the subcritical due to the higher operating temperatures required.
An overall simplified net reaction of biomass under supercritical gasification can be summarized as:

CHx+(2y)H2OCO2+(2y+x2)H2

image [17.i]

where x and y represent the molar rations of H/C and O/C in biomass, respectively (Guo et al., 2007). The intermediate and interacting reactions during HTG are mainly steam-reforming, water-gas shift, and methanation reactions:

CHxOy+(1y)H2OCO+(1y+x2)H2

image [17.ii]

CO+H2OCO2+H2

image [17.iii]

CO+3H2CH4+H2O

image [17.iv]

image
Figure 17.4 Equilibrium gas yields from 5 wt% wood sawdust. Adapted from Guo, L.J., et al., 2007. Hydrogen production by biomass gasification in supercritical water: a systematic experimental and analytical study. Catalysis Today 129(3–4), 275–286.
HTG has many promising attributes for the formation of H2 from biomass as it can use a variety of feedstocks and presents a sustainable alternative to steam reforming of natural gas. Fig. 17.4 shows the equilibrium gas yields from a slurry of 5 wt% (weight percentage) wood sawdust adapted from Guo et al. (2007). When a high H2 yield is required, reaction [17.iv] is undesired and should be suppressed to avoid the formation of methane from H2. The yield of methane drops steadily from 400 to 800°C while the fraction of H2 increases. Above 800°C, gasification is complete and reaction [17.i] can be used to predict the molar ratio of CO2 and H2. Therefore from a thermodynamic point of view, higher temperatures are not required during HTG. At lower temperatures, CH4 production is favored. Although the amount of water to biomass is also significant in this context, a higher water content will move the equilibrium toward hydrogen and away from methane production by reaction [17.ii].

17.2.3. Catalytic hydrothermal processing

Catalysis has been investigated in all hydrothermal processing routes to improve product selectivity and quality. In HTC, homogeneous acid catalysts are most commonly employed to improve the hydrochar yield. However catalysts are not necessarily required during HTC due to the acidic conditions associated with the process water during processing.
In HTL, heterogeneous and homogeneous catalysts have been applied “in situ” during hydrothermal processing in an attempt to improve biocrude quality. Reductions in oxygen, increased yields, and higher energy recoveries have been obtained. However for feedstock high in protein, which results in high levels of nitrogen in the biocrude, the use of “in situ” catalysts has been less successful. In situ catalysis is often used during liquefaction of lignocellulosic feedstocks where an alkali is added such as NaCO3, KOH, or K2CO3. For algal and other high N feedstocks, homogeneous catalysts such as formic and acetic acids have applied as well as heterogeneous catalysts; Pt/Al, NiMo, Ni-Ra, CoMo, Pt-C and many more (Ross et al., 2010; Biller and Ross, 2011; Duan and Savage, 2010; Luo et al., 2015). The effects of alkali catalysts on lignocellulosic feedstocks are reported to be quite beneficial in terms of yield improvements and biocrude quality. The alkali catalyst reduces acid catalyzed decomposition of carbohydrates, phenol aldehyde cross linking reactions and favors more desirable aldol condensation reactions. During HTL of lignocellulosics without alkali, the high concentration of produced acids results in a low pH (<4) which favors dehydration, polymer and char formation (Zhu et al., 2014). For microalgae feedstock, where the majority of HTL work has taken place, the use of catalysts only increases the yields slightly and reduces the oxygen content of biocrudes. Due to the high nitrogen content of algae, the biocrudes generally contain high levels of N (5–7%), which is an issue during combustion and hydrotreating. None of the “in situ” catalytic work to date has been able to address the issue of N in biocrude satisfactorily. Attention has therefore shifted toward the upgrading of the biocrude using either catalytic hydrotreating or catalytic hydrothermal treatment (Bai et al., 2014; Duan and Savage, 2011a,b; Elliott et al., 2013b). Hydrotreating involves processing the biocrude with hydrogen over a catalyst. Hydrogenation reactions convert oxygen, nitrogen, and sulfur into H2O, NH3, and H2S, respectively. A more detailed description of the catalytic upgrading of the biocrude generated from HTL is described by Elliott in chapter “Production of biofuels via bio-oil upgrading and refining.”
Catalysis appears to have been most successful in HTG where good conversion rates of biomass to syngas can be achieved at temperatures below the supercritical point, hence performing subcritical catalytic hydrothermal gasification (CHG) (Elliott, 2015). Early work in HTG suggested that supercritical conditions were necessary to achieve reasonable gas formation rates from biomass but more recent work has shown that with the right catalysts, HTG can operate successfully at lower temperatures. The use of catalysis can also affect the selectivity toward H2 or CH4; NaOH for example shifts the reaction equilibrium toward the production of H2, while the presence of Ni increases the production of CH4. Catalysts also allow operation at much lower temperatures resulting in so-called catalytic HTG (or CHG). Temperatures can be lowered as far as 350–400°C if the primary aim is to produce methane rather than hydrogen. Ru and Ni catalysts have been shown to be most effective during HTG (Elliott, 2008). One of the main issues in catalytic HTG is the poisoning of active sites by sulfur. Researchers at the Paul Scherrer Institute have developed catalyst and reactor set-ups to avoid this problem, by precipitating salts such as sulfates prior to the catalyst bed (Luterbacher et al., 2009).

17.3. Process layout

17.3.1. Hydrothermal carbonization

The reactor design in hydrothermal processing can be either batch or continuous. Continuous reactors require feeding systems that operate under pressure and include either slurry-based pumps or lock and hopper systems for larger particles. The use of larger particles usually requires more severe reaction conditions to achieve similar carbonization to smaller particles. The feedstock may require pretreatment to reduce the particle size and/or to remove contaminants such as stone, metal, or plastic. The uniform feed is fed into the hydrothermal reactor where it is processed typically for several hours depending on the reactor design. The slurry is then typically mechanically dewatered and dried. The hydrochar can be pelletized depending on final use. The process water is either treated to remove nutrients, recycled or both. A general process layout for HTC is summarized in Fig. 17.5.
image
Figure 17.5 Schematic layout of an HTC process.
The process water is considered a waste product but advancements have recently been made in treating this stream by anaerobic digestion resulting in enhanced biogas yields (Fettig, 2010; Danso-Boateng et al., 2015; Reza et al., 2014a,b). The process water from HTC contains high levels of TOC and therefore linking to AD can produce significant levels of biogas. Levels of biological methane potential range from 0.5 to 1 L/g TOC (Wirth and Mumme, 2014; Oliveira et al., 2013). Some challenges remain, such as reducing inhibition to methanogens but integration has clear advantages and the integration of HTC with AD shows particular promise.
It is generally considered that for continuous operation, the process waters will need to be recycled to minimize water usage (Uddin et al., 2014). During HTC, the soluble organics in the water phase are both being produced and adsorbed although there is evidence that the production dominates (Uddin et al., 2014), which means it will continue to increase on multiple cycling. The level of inorganics will also increase leading to a build-up of inorganics in the reactor (Uddin et al., 2014). This may cause problems with salt corrosion and may require additional measures to precipitate salts with possible recovery of nutrients. Recycling of process water can be considered for each of the hydrothermal process routes and in HTC it has been shown to increase the HTC yields of the hydrochar (Uddin et al., 2014).

17.3.2. Hydrothermal liquefaction

There is growing interest in the commercial development of hydrothermal liquefaction for the treatment of biosolids, emerging feedstock like algae and lignocellulosic feedstocks. Despite this, the vast majority of research reported on HTL is performed in small batch reactors. This work provides valuable insight into parametric operating conditions, allows reaction pathways to be investigated and product yields and qualities to be predicted. For industrial applications of HTL, continuous reactor systems are required to improve the economics of the process. For HTC, reactor design can be continuous or batch but the latter is unlikely for HTL and HTG as lower residence times are favored. Continuous systems also have the potential to implement advanced heat recovery.
The advantage of continuous flow systems include higher throughput and improved heat recovery. Continuous operation generally has a lower residence time than batch operation although there are benefits of higher residence times for product distribution as described further in Section 17.6. As residence times decrease due to increasing flow rate, the reactor output increases and energy requirements decrease. Careful consideration should be given to flow rates, heating rates, and residence times in terms of biocrude yield, chemical energy recovery, capital reactor cost, and energy requirements. The optimum conditions for continuous HTL however need to be carefully adjusted as, for instance, operation at lower residence times also reduces the chemical energy recovery. The effect of this trade-off should become more apparent with the transition from lab-scale to pilot- and full-scale HTL systems.
The general steps anticipated in a modern hydrothermal liquefaction facility are summarized in Fig. 17.6(a) for lignocellulosic biomass. The processing and pumping of feed slurries of low particle size may require pretreatment for certain feedstock. For woody biomass this typically includes size reduction and alkaline treatment to obtain a pumpable and stable slurry.
The feedstock is then processed via HTL at temperatures of around 350°C and pressures of 180 bar for approximately 15 min. Phase separation occurs spontaneously after the reaction resulting in a gaseous phase rich in CO2, solid residue, the biocrude, and an aqueous phase. The process water can be recycled reducing water requirements which can enhance biocrude yields. Waste process water can be treated anaerobically or treated via catalytic HTG or CHG to produce either a methane-rich or hydrogen-rich syngas (Elliott et al., 2013b; Cherad et al., 2015). Anaerobic digestion of HTL process waters has not been demonstrated experimentally and could prove difficult due to compounds in the water phase such as furfurals and phenols which are inhibitory to AD. There is potential for the recovery of N and P following treatment of the water by either AD or hydrothermal gasification. The produced biocrude requires upgrading via hydrotreating to produce final fuels or a refinery feedstock.
image
Figure 17.6 Schematic layout of the HTL process for lignocellulosics (a) and microalgae (b).
A conceptual scheme for the production of biofuel from microalgae by hydrothermal liquefaction is described in Fig. 17.6(b). A benefit of processing algae by HTL is the potential for recycling of nutrients back to cultivation.
Microalgae require no pretreatment due to their small particle size and are pumped more easily as a slurry. The processing of other biomass may require grinding or maceration, adding additional cost. Algae still require dewatering during harvesting to produce a slurry containing approximately 20% solids. Feedstock such as sewage sludge would operate similarly to algae but still require dewatering and thickening before processing. Some research has shown it is possible to recycle the process water after dilution to algal cultivation and the algae can utilize the organic carbon by heterotrophic growth however this has not been demonstrated at scale (Biller et al., 2012).

17.3.3. Hydrothermal gasification

Hydrothermal gasification is the least developed commercially of the three hydrothermal processes although there is extensive understanding of the process chemistry. Smaller continuous facilities have been in operation in a number of laboratories for some time. Most notable are the facilities at the Paul Sherrer Institute led by F. Vogel and at PNNL led by D.C. Elliott (Luterbacher et al., 2009; Elliott et al., 2004). Feeding of biomass is usually via slurry pumps or lock and hopper devices. Continuous gasification of a range of feedstock have been reported including lignocellulose, sewage sludge, macroalgae, and glycerol. Salt management has been identified to be critical in operation in CHG and HTG. The Paul Sherrer Institute has developed technology for the continual withdrawal of inorganics during gasification, resulting in a concentrated brine solution with potential for nutrient recovery (Schubert et al., 2010a,b, 2012). This approach has the added advantage of reducing catalyst poisoning. Using a suitable catalyst can improve gasification and increase the formation of methane at lower temperatures. The process water from HTG and CHG contains low TOC although higher temperature operation can produce some tar (Müller and Vogel, 2012). Syngas produced at lower temperature in the presence of catalysts is high in methane although higher-temperature HTG can produce high hydrogen yields. Syngas high in methane can be used for process heat requirements. Integration of HTL and CHG/HTG has been proposed as a means of increasing energy recovery from the HTL process water and providing a source of hydrogen for upgrading of the biocrude.
The general steps anticipated in a modern hydrothermal gasification facility are summarized in Fig. 17.7.
image
Figure 17.7 Schematic layout of an HTG process. Adapted from (Haiduc, A., et al., 2009. SunCHem: an integrated process for the hydrothermal production of methane from microalgae and CO2 mitigation. Journal of Applied Phycology 21(5), 529–541; Stucki, S., et al., 2009. Catalytic gasification of algae in supercritical water for biofuel production and carbon capture. Energy & Environmental Science 2(5), 535–541).

17.4. Feedstock considerations

Modern research in hydrothermal processing has concentrated on the utilization of biomass and wet wastes. The process has clear advantages for wet feedstocks but biomass with low moisture content such as woody biomass has also been investigated. Most biomass can be processed by hydrothermal conversion but one of the main considerations is often the ease of pumping and material transfer. The feed is typically fed as a water slurry due to the hydrophilic nature of biomass and the reasonable ease in forming water slurries with biomass particles at pumpable concentrations. This is typically possible up to 30% solids, details on biomass slurry pumping are review by Elliott et al. (2015).
One of the main advantages of hydrothermal processing is its potential flexibility; this allows the use of feedstocks which are usually difficult to treat. This could include difficulties associated with high moisture content, nonhomogeneity, and bioactivity, as well as high ash content. Due to the versatile nature of the process, the feedstock can vary in composition and quality throughout operation without major implications on the process layout, although some consideration has to be given to the ease of pumping of the feedstock and quality of the products. Many types of wet biomass such as sewage sludge, AD press cakes, animal manures, food waste, and distillers' grain are available in large quantities. Biowastes are probably the most promising feedstocks for hydrothermal processes although significant effort has been directed at the utilization of future feedstocks such as algae which may become available in large quantities if this technology is favored for future oil production.
HTC has mainly been performed on lignocellulosic biomass; however there is growing interest in the processing of wet waste and biosolids such as municipal waste, sewage sludge, food waste and algae (Berge et al., 2011; He et al., 2013; Hoekman et al., 2013; Heilmann et al., 2010; Broch et al., 2013). One of the drawbacks of processing waste is the presence of heavy metals, pathogens, and biologically active compounds. The high temperatures employed in hydrothermal processing have the potential to destroy biologically active material rendering the products pathogen-free and the process streams sterile (Pham et al., 2013). This has been demonstrated for liquefaction of algae and pig manure (Pham et al., 2013). This has added benefits in improving the safety of products and may be beneficial for recovery and reuse of nutrients, pretreatment of process water, and discharging of effluents.
For low-moisture feedstocks such as lignocellulosics, recovery and reuse of the water for slurry preparation is imperative. This has been demonstrated to have beneficial effects on biocrude generation in HTL and hydrochar yields in HTC (Uddin et al., 2014; Zhu et al., 2015; Ramos-Tercero et al., 2015). For high-moisture biomass such as algae consideration must be given to avoid costs of processing excessive amount of water. Table 17.1 presents some common feedstocks processed by hydrothermal processing and their typical compositions.
Microalgae are particularly suited for continuous hydrothermal processing due to their small size (<100 μm) and ease of pumping. Algal biomass have become a particularly promising alternative resource for renewable fuels due to their higher photosynthetic efficiency and area-specific yields (Pienkos and Darzins, 2009; Ross et al., 2008). The ability to process microalgae whole and with high moisture content is one of the attractive features of HTL over say lipid extraction or pyrolysis. Microalgae can contain high levels of lipids which are beneficial in HTL, improving biocrude yields and quality. It can be argued that microalgae mass cultivation for bioenergy is still at the research and development stage and no industrial-scale cultivation facilities are currently in place. Therefore in the short to medium term, hydrothermal processing of biowastes and other wet biomass appears more likely compared to microalgae.
Lignocellulosics may require pretreatment such as grinding if pumped as a slurry; however some feedstock such as animal wastes and sludges can be processed in a similar manner to microalgae.
Particle size is less important in HTC, particularly when batch systems are being used; however HTL and HTG rely on pumping slurries and employ continuous systems. The influence of particle size on HTC has not been fully investigated; however it is likely to have considerable effects on the carbonization process. If particle sizes are too large, the inside of biomass particles are not fully contacted with water, and heat transfer is faster than mass transfer, leading to pyrolysis reactions rather than dehydration and decarboxylation.
Macroalgae, also known as seaweed, are a group of photosynthetic marine organisms. They generally have a low lipid content but are high in carbohydrates and have shown to be particularly suitable for HTG/CHG (Cherad et al., 2013). Macroalgae are generally of larger size than microalgae and so are more easily harvested although they require maceration in order to pump. An added complication with macroalgae is the presence of hydrocolloids such as alginates which become highly viscous in solution. Macroalgae are also high in ash and contain high chloride content which may promote chloride stress corrosion.

Table 17.1

Summary of feedstock being considered for hydrothermal processing

Feedstock (db)WoodsGrassesaMacroalgaeMicroalgaeManuresSewage sludgeDigestate
Moisture50–75%60–85%∼70∼99%∼90%∼90%∼90%
Ash3–86–915–357–2610–2020–5030–60
O%35–4540–5025–4025–3035–455020–40
N%0.5–31–43–75–93–63–81–4
HHV (MJ/kg)12–2015–2510–2025–3010–20∼14∼20
Size1–100,000 mm10–1000 mm1–10,000 mm1–100 μm1–10,000 μm1–100,000 μm1–10,000 μm
ReferencesUmeki et al. (2010), Wang et al. (2011)Wachendorf et al. (2009)Ross et al. (2008, 2009)Biller and Ross (2014)Wang et al. (2011), Vardon et al. (2011)Fonts et al. (2012)Reza et al. (2015), Smith et al. (2016)

image

a Includes unpublished data by the authors.

17.5. Product distribution and properties

Much of the research and development (R&D) reported in the literature on hydrothermal processing utilizes small batch reactors. The optimization of reaction conditions to maximize hydrochar, biocrude, and gasification efficiencies has been studied widely in batch reactors from a wide range of feedstocks (Garcia Alba et al., 2011; Brown et al., 2010; Jena et al., 2011). Each HT pathway has in common that an aqueous byproduct is produced alongside the primary products hydrochar, biocrude, and syngas.

17.5.1. Hydrothermal carbonization

The products from HTC include a solid biochar, noncondensible gases (mainly CO2), and water phase containing soluble organics and mineral matter and water (Hoekman et al., 2013). Applications of the hydrochar include use as a fuel directly or as supports for catalysts, adsorbents, soil enrichers, and functionalized materials with high energy storage capacity (Titirici and Antonietti, 2010; Sevilla et al., 2011; Sevilla and Fuertes, 2009a,b). HTC is performed at moderate heating rate within temperatures ranging from 180 to 250°C in water at pressures in the range 12–40 bar. The water is in the subcritical state to avoid gasification and the whole biomass must be submerged in liquid water to avoid pyrolysis (Funke and Ziegler, 2010). Typical yields of solid hydrochar range from 50 to 80 wt%, with 5–20% in the process water and 2–5% gas. The yields and properties of the resulting hydrochar are highly dependent upon operating conditions such as temperature and residence time and the feedstock biochemical composition (Funke and Ziegler, 2010). Fig. 17.8 describes the trends in yields and HHV during HTC.
Careful control of the process conditions and feedstock can result in highly functionalized carbon materials (Libra et al., 2011). Higher temperatures or longer residence times result in higher energy densification although yields of hydrochar are typically lower. A higher biomass to water loading will result in higher yields of hydrochar and lower yields of soluble hydrocarbons in the water (Stemann et al., 2013). Water can be recycled resulting in improved energy densification and increase yields of hydrochar (Stemann et al., 2013). The acidic nature of the process water due to high levels of organic acids can increase the reaction rates of HTC and is an added benefit of water recirculation (Berge et al., 2011). Typical yields of hydrochar are listed in Table 17.2 for a variety of typical feedstocks.
During HTC the biomass undergoes energy densification to a hydrochar which resembles a “lignite-like coal” with an energy density of 28 MJ/kg (Libra et al., 2011; Hoekman et al., 2013). As the temperature increases, yields of hydrochar will decrease, resulting in higher energy densification (Berge et al., 2011; Hoekman et al., 2013). Typically, 60–90% of the HHV of the biomass or waste is maintained within the hydrochar, representing approximately 75–80% of the original carbon (Ramke et al., 2009). Generally with lignocellulosic biomass, a biomass with a heating value of 16–18 MJ/kg can produce a biocoal with an energy density between 28 and 30 MJ/kg. Higher ash feedstock such as sewage sludge, MSW, and digestate produce more moderate energy densification due to the higher levels of inorganic material. The change in O/C ratio can be seen in the Van Krevelen diagram shown in Fig. 17.9. This indicates that the hydrochar produced has an energy density similar to that of a lignite coal. The morphology of the hydrochar particles is dependent on the feedstock composition and contains interesting carbon nanostructures (Titirici and Antonietti, 2010). The outer shell of this particle is thought to be more hydrophilic whereas the inner core is thought to be hydrophobic, this hydrophobicity increases with reaction (He et al., 2013). The hydrophobicity of the hydrochar reduces its water content and is likely to improve its handling and storage properties.
image
Figure 17.8 HTC yields and HHV of hydrochar as a function of temperature.
In addition to the increased energy density, the hydrochar also has different ash behavior when heated. Ash high in alkali metals, chloride, and sulfur can result in problems with slagging and fouling during combustion. An additional benefit of HTC is the improved combustion behavior of the resulting hydrochar due to removal of many of the problematic inorganic materials. Fig. 17.10 shows typical levels of reductions in inorganic species in biomass. Alkali metals such as potassium and sodium are significantly reduced together with high levels of chloride, sulfur, and a portion of the nitrogen and phosphorus. The impact of this is to increase ash melting temperatures and reduce corrosion. This change in ash behavior during heating is measured using an ash fusion test. An example for miscanthus is shown in Fig. 17.11, HTC is shown to increase the ash transition temperatures resulting in reduced problems with slagging.

Table 17.2

Examples of different feedstocks processed under HTC conditions

Feedstock (db)LignocellulosicsAlgaeAgricultural residuesSewage sludgeDDGSa
Yield (% daf)a40–5020–4040–6060–7030–40
HHVa (MJ/kg)28–308–1219–261529
C%60–8030–5055–7233∼65
H%10–125–75–647
N%0–0.55–60.5–224
O%20–3020–3016–301820–25
Residence time (h)0.5–201–21–21–20.5–2
ReferencesOliveira et al. (2013), Uddin et al. (2014), Hoekman et al. (2013)Heilmann et al. (2010), Broch et al. (2013)Oliveira et al. (2013)He et al. (2013)Heilmann et al. (2011)

image

a DDGS, dried distillers' grains with solubles; daf, dry ash free basis; HHV, higher heating value.

image
Figure 17.9 Van Krevelen diagram showing hydrochars, biomass, lignite, and coals. Adapted from Smith, A.M., Singh, S., Ross, A.B., 2016. Fate of inorganic material during hydrothermal carbonisation of biomass: influence of feedstock on combustion behaviour of hydrochar. Fuel 169, 135–145.
image
Figure 17.10 Typical levels of extraction of inorganics during HTC. Modified from Smith, A.M., Singh, S., Ross, A.B., 2016. Fate of inorganic material during hydrothermal carbonisation of biomass: influence of feedstock on combustion behaviour of hydrochar. Fuel 169, 135–145.
image
Figure 17.11 Ash transition temperatures for hydrochar from miscanthus grass. Modified from Smith, A.M., Singh, S., Ross, A.B., 2016. Fate of inorganic material during hydrothermal carbonisation of biomass: influence of feedstock on combustion behaviour of hydrochar. Fuel 169, 135–145.
The HTC gaseous products are largely CO2 although there are also small amounts of CO, CH4, and H2 present. Increasing reaction severity will result in higher levels of decarboxylation and a corresponding increase in CO2.

17.5.2. Hydrothermal liquefaction

The products from HTL include a biocrude, a process water containing polar organics, and mineral matter, a solid residue low in carbon and a gaseous phase which is again largely CO2. Up to 80% of the carbon is fractionated into the biocrude, which is a hydrophobic mixture of many organic compounds, the composition of which is dependent on the biochemical composition of the feedstock. It is a highly viscous, black, tar-like material at room temperature. There has been limited research into the direct utilization of biocrude either as a heavy fuel or an additive for asphalt or a blending feedstock with other liquid fuels. Attention has concentrated on upgrading the biocrude catalytically, thermally in a hydrogen atmosphere, or hydrothermally to hydrodeoxygenate and crack the heavy fuel fractions. Biocrude differs significantly from pyrolysis oil in that it has a much lower water and oxygen content. Table 17.3 presents a comparison of oils from HTL and flash pyrolysis. Water content is typically 1–5% and O content around 7–20% compared to water content in pyrolysis oils of 20–40% and O of 25–45%.
Biocrude is more viscous that pyrolysis oil; however it is less dense. It contains a large number of different molecular compounds with a range of molecular weights (average MW ∼1000). The molecular composition is largely influenced by feedstock composition and operating conditions.

Table 17.3

Comparison of HTL biocrudes to flash pyrolysis bio-oils

HTLFlash pyrolysis
Elemental analysis (wet)
C (wt%)7358
H (wt%)86
O (wt%)1636
S (ppm)<4529
Moisture5.124.8
HHV (MJ/kg)35.722.6
Viscosity (cps)15,000 at 61°C59 at 40°C

image

Adapted from Elliott, D., 1993. Evaluation of wastewater treatment requirements for thermochemical biomass liquefaction, In: Bridgwater, A.V., Advances in Thermochemical Biomass Conversion, Springer Netherlands. pp. 1299–1313.

Several groups have investigated the composition of HTL biocrude from different feedstocks. Torri et al. described in detail the composition of biocrude derived from microalgae (Torri et al., 2011). Identified compounds from microalgae biocrude are attributed to their origin of either carbohydrate, protein, lipid, or algaenan. Lipids generally result in free fatty acids and alkanes, while carbohydrates form oxygenated compounds, phenols, alcohols, and cyclic ketones. Degradation products from protein include indoles, pyrroles, and pyrazines. The characterization of biocrudes is an area of ongoing investigation due to its complex composition and high molecular weight complicating its characterization. Although named biocrude, it is not a petroleum analogue due to the higher oxygen and nitrogen content in the oils. Particularly when processing feedstocks with a large protein content the nitrogen content is found to be undesirably high. This has been shown to be an issue in the liquefaction of microalgae due to their high protein content of up to 70%. This issue is not present for the liquefaction of lignocelluloses; however their liquefaction results in lower yields and higher viscosities. There have been several feedstocks under investigation as shown in Table 17.4 for the production of biocrude. Microalgae have been shown to result in the highest yields and HHV particularly when high-lipid-containing algae were utilized. The liquefaction of lignocelluloses generally results in lower yields but carbon efficiencies of >50% are still achieved.
Manures and sludges appear to be a promising feedstock for HTL with good energy recoveries but their investigation has been limited to date and not employed on continuous reaction systems. Continuous HTL has been performed on feedstocks such as corn stover, DDGS, forest residue, macroalgae, and microalgae. PNNL published a carbon balance on the liquefaction of lignocelluloses on their continuous bench scale systems. Overall carbon yield, including hydrotreatment of the biocrude product, was nearly 50% (Elliott et al., 2015).

Table 17.4

Examples of different feedstocks processed under HTL conditions

Feedstock (db)LignocellulosicsMacroalgaeMicroalgaeManuresSewage sludgeDDGS
Bio-crude
Yield (% daf)29–509–2838–64304039
HHV (MJ/kg)3632–34353235
N%0.33–44–8465
O%126–85–18161815
Batch (b)/continuous (c)cc/bcbbc
ReferencesFrans and Jaap (2015), Zhu et al. (2014)Elliott et al. (2013a), Anastasakis and Ross (2011)Elliott et al. (2013b), Jazrawi et al. (2013)Vardon et al. (2011)Malins et al. (2015)Mørup et al. (2015)

image

Wet algae slurries have been converted into an upgradeable biocrude by HTL in a bench-scale continuous-flow system at PNNL (Elliott et al., 2013b). Direct biocrude recovery could be achieved without the use of a solvent and biomass trace components could be removed. High yields of biocrude (up to 82 wt% on a carbon basis) could be obtained with high slurry concentrations of up to 34 wt% of dry solids (Elliott et al., 2013b). Researchers at Aarhus University have processed DDGS in a continuous system for 24 h and were able to recover 60% carbon in the biocrude and a yield of 39 wt% (Mørup et al., 2015).
Continuous-flow HTL work by Jazrawi et al. (2013), Biller et al. (2015), and the batch work by Faeth et al. (2013) suggests that higher heating rates and lower residence times favor the production of biocrude. Fast heating rate batch systems have resulted in biocrude yields exceeding 60 wt% and energy recoveries of around 90% from microalgae. Fast heating rates have not been employed on other feedstocks to date but it is possible that similar effects can be seen. A drawback of the very fast heating and low residence times is a reduction in deoxygenation of the biocrude. Operating with longer residence times results in a reduced oxygen content and higher carbon content corresponding to an increased HHV. For example, in one study the change in reaction times from 10 to 90 min resulted in a reduction in oxygen content from 16% to 8% (Faeth et al., 2013). The quality of the biocrude is superior at the longer residence times and despite its lower yield, is more suitable for upgrading due to lower H2 demand for hydrodeoxygenation. The increased HHV at longer residence time results in a higher chemical energy recovery of the biocrude when processing algae.
The continuous HTL studies by Elliott et al. (2013b) and Morup et al. (2015) differ to those by Jazrawi et al. (2013) and Biller et al. (2015) as they did not employ the use of solvents to recover the biocrude. This has two advantages: firstly it avoids the use of toxic solvents which incur cost and can be harmful and secondly it adds undesired compounds to the biocrude and can therefore diminish its quality. Xu and Savage (2014) demonstrated in a batch HTL study the effect of using DCM for the recovery of biocrude (Xu and Savage, 2014). It was shown that additional biocrude is recovered from the aqueous phase compared to using no solvents. This accounted to about 8% of the total biocrude. There were remarkable differences between the biocrude recovered from the aqueous phase to the biocrude which was recovered directly without DCM. The O and N contents were found to be twice as high, adding additional N and O to the biocrude when recovered using DCM from all product phases, thereby diminishing its quality.

17.5.3. Hydrothermal gasification

HTG for the production of methane and hydrogen has received considerable attention in the literature and has been summarized in various reviews (Elliott, 2008, 2015; Kruse, 2009). A large amount of research in HTG has been on understanding the process chemistry using model compounds. There has been less research and development performed on HTG of biomass compared to HTC and HTL. HTG feedstocks under investigation include sawdust, swine manure, sugarcane bagasse, micro- and macroalgae, sewage sludge, and many more (Elliott, 2008; Yakaboylu et al., 2015). Examples of feedstocks processed under HTG conditions are presented in Table 17.5. In most cases the gasification efficiency is found to be high (>90%), particularly when catalysts are employed.

Table 17.5

Examples of feedstocks used for HTG

Feedstock (dry basis)LignocellulosesMacroalgaeMicroalgaeManuresSewage sludge
Carbon conversion (%)50–10072–9368–747660–100
CatalystRa NiNi/Al2O3, NaOHRu/CNiAC, Ra Ni
Temp (°C)300–410500400405400–650
Residence time (min)90303660–120
ReferencesWaldner and Vogel (2005)Cherad et al. (2013), Onwudili et al. (2013)Haiduc et al. (2009)Waldner (2007)He et al. (2014)

image

AC, Activated carbon; Ra Ni, Raney nickel.

There have been several papers investigating the use of lignocellulosic biomass feedstocks for HTG, for example Waldner and Vogel (2005) report a methane yield of 0.33 g of CH4 per g of wood. Schumacher et al. (2011) investigated the supercritical water gasification of macroalgae at 500°C and produced 12 and 13 g of H2/kg seaweed. More recently, Onwudili et al. (2013) investigated the catalytic SCWG of S. latissima producing 30 g of H2/kg seaweed in the presence of sodium hydroxide as a catalyst and noting a doubling of methane yield in the presence of nickel catalyst to 112 g CH4/kg seaweed (Onwudili et al., 2013). The main challenges associated with HTG of biomass slurries are the selection of suitable catalysts which are not deactivated as well as gas separation and scaling of reactor systems.

17.5.4. Composition of the process water

The water phase from all three hydrothermal processing conditions is a byproduct which has to be considered in the overall process. It is relatively high in C, N, and inorganics. The water phase contains soluble organic material such as sugars, organic acids, phenols, furfurals, and 5-HMF (Kumar and Gupta, 2009). The total organic carbon content has been reported to be in the range of 10–50 g/L, although this amount increases if water recycling is employed (Elliott, 1993). TOC and COD generally increase with temperature and reaction time up to 280°C but then reduce. The type of organic compounds in the process water also changes as the temperature increases and is sensitive to feedstock composition. The process waters from HTG are typically very low in organic carbon (200–1000 mg/L) demonstrating high gasification efficiencies. The remaining carbon compounds identified include acetaldehyde and acetic acid and very small amounts of phenols, oxalic acid, formic acid, and methanol (Waldner et al., 2007).
The types of sugars and polar compounds produced from HTC of woody biomass and microalgae are discussed by Hoekman et al. (2013) (Broch et al., 2013; Reza et al., 2014c). Lignocellulosic feedstocks produce higher levels of 5-HMF and furfural than other biomass. Levels of galactose, xylose, mannose, and fructose are highest at the lowest temperatures. As the temperature increases the levels of glucose and levoglucosan increase. The compounds in the process water can be classified as sugar-derived or lignin-derived. The main sugar derived components include furfural, 2-ethyl-5-methyl furan, and 2-hydroxy-3-methyl-2-cyclopentanon (Xiao et al., 2012). Furfural is a degradation product of cellulose. 5-HMF and furfural dominate in the process water at higher temperatures (Hoekman et al., 2013). High levels of volatile fatty acids (VFA) are produced such as butanoic acid, propanoic acid, acetic acid, and lactic acid. The level of VFA in the water reduces with increasing reaction time (Danso-Boateng et al., 2015). Lignin can depolymerize and undergo repolymerization resulting in the formation of phenolic fragments such as 2,6-dimethoxyl phenols, dimethoxy benzaldehyde, and esters.
Processing of feedstocks high in nitrogen, such as manures or algae, leads to increased nitrogen in the process water. Values of total nitrogen have been reported as 3000–8000 mg/L, mainly in the form of ammonium and nitrates. Organic nitrogen such as pyrroles and pyrrolidinones cyclic dipeptides, aromatic nitrogen compounds, amines, amides, and amino acids have been identified in the process water (Garcia Alba et al., 2013). The pH of process waters from lignocellulosic biomass range from 3 to 4, whereas with high protein feedstocks, such as microalgae, have been shown to have a higher pH. This is due to the effect of the N-compounds in the process water (Broch et al., 2013).
Table 17.6 lists some of the main chemicals identified in the process waters from the different hydrothermal routes and indicates the relative amounts under different conditions.
Algal biomass produces significantly higher levels of arabitol and mannose than lignocellulosic and reflects the different carbohydrate components in algae. The main acids produced from microalgae include malonic acid, lactic acid, succinic acid, and glutaric acid. The maximum amounts of sugars from microalgae were shown to be at 230°C, above which they degrade further and increase the levels of organic acids. The process water can also contain heavy metals and alkali salts (Reza et al., 2013; Smith et al., 2016).
Several approaches have been used to tackle the problem of the wastewater byproduct. For example in the case of microalgae HTL, PNNL applied catalytic hydrothermal gasification (CHG) effectively for HTL process water clean-up and fuel gas production from water-soluble organics, allowing the water to be considered for recycling of nutrients to the algae growth ponds (Elliott et al., 2013a). In the case of the process water from macroalgae HTL, CHG conversion of 99.2% of the carbon in the aqueous phase was achieved (Elliott et al., 2013b).

Table 17.6

Typical components found in the process water

CompoundOriginCommentsHTCHTLHTG/CHG
GlucoseCellulosexxxx
Fructosexxxxx
Sucrosexxxxx
Acetic acidHemicellulosexxxxxxx
Formic acidxxxxxx
Lactic acidxxxxx
Butyric acidxxxxx
Propanoic acidxxxxx
XyloseHemicellulose cellulosexxxx
Mannosexxxx
Arabinosexxxx
Galactosexxxx
AcetoneCarbohydratesx
5-HMFxxxxx
Furfuralxxxx
2-Ethyl-5-methyl furanxx
2-Hydroxy-3-methyl cyclopentenonexx
Malonic acidCarbohydrateAlgal carbohydratesxxxx
Succinic acidxxxx
Glutaric acidxxxx
2,6-Dimethoxy phenolLigninLignocellulosic biomassxx
3-Methoxy-1,2-benzene diolxx
Alkyl benzenesxxx
Fatty acidsLipidsAlgae, food wastexx
Glycerolxxxx
Fatty acid amidesxx
AmidesProteinHigh in algae, food waste, manuresxxx
Pyrrolesx
Indolesxx
Pyridinesxx
Table Continued

image

CompoundOriginCommentsHTCHTLHTG/CHG
NH4+xxxxxx
PO43imagexxxx
TOCxxxxxx
pH 3–65–87–8

image

Relative amounts of species indicated as: xxx, high; xx, mid; x, low.

Anaerobic digestion of the aqueous phase from hydrothermal processing has been suggested in various reports; however experimental work has only been successfully carried out on the process water from HTC. Further research is required to evaluate the potential for HTL process water by AD and the inhibitory effect on methanogenesis. Wirth and Mumme (2014) have demonstrated that the process water can be treated using AD. Typical values of TOC are in the range 10–20 g/L and Erlach and Wirth have shown the biological methane potential (BMP) to be in the region of 0.5–1 L methane per gram of TOC (Erlach et al., 2011). BMP has been estimated from the Buswell equation for a range of feedstock such as sewage sludge and digestate (Danso-Boateng et al., 2015).
The potential for recycling nutrients from the HTL and HTG process water for algae cultivation has been attempted on a laboratory scale (Biller et al., 2012; Garcia Alba et al., 2013; Jena et al., 2011). The recycling of nutrients is regarded to be essential in microalgae biofuel systems since the energy, cost, and carbon emissions associated with nutrient production are very high. In most cases, it was necessary to significantly dilute the process waters to avoid growth inhibition from components such as phenols and metals such as nickel (Biller et al., 2012). Another study investigated the use of the HTL aqueous byproduct as a substrate for microbial growth under heterotrophic rather than phototrophic conditions. The microbial strains were shown to have a higher tolerance to the aqueous phase, leading to significantly reduced dilution requirements compared to microalgae (Nelson et al., 2013).

17.6. Development of technology and current research

17.6.1. Development of reactor systems

For all three of the hydrothermal processing routes, the majority of the literature published to date is performed on batch reactors. For the development of the technology, continuous systems will be required and this has been addressed in industry and academia in recent years. A number of industrial pilot-scale HTC plants are operating and described in Table 17.7. The capacities of the plants range from 6000 to 60,000 tons per year. The operating conditions used are analogous to those employed in batch studies reported in the literature. Therefore, despite there being limited data available on the full-scale plants operating, the laboratory batch reactor is broadly representative to what would be expected on a larger scale. HTC has the advantage that much lower operating temperatures and pressure (10–40 bar) are utilized, which reduces some of the engineering challenges associated with very high pressures as employed during HTL and HTG. HTC can either be run in batch, semibatch, or continuous mode while this is unlikely for HTL and HTG which will require fully continuous operations. One of the challenges with HTC reactor developments is the presence of large particles, these can lead to plugging in continuous plug flow type reactors and cause blocking in back-pressure regulators. There are only a limited number of continuous industrial or pilot-scale plants in operation and the development of reactor layouts is still ongoing. Whether indeed fully continuous plug flow or CSTR type reactors will prevail over batch or semicontinuous systems is currently unclear.
Contrary to HTC, where only limited reports on continuous processing are available, researchers have published increasing number of reports on the continuous operation of HTL. The first continuous HTL facilities were constructed in the 1970s and 1980s during the development of the HTU process, the PERC Albany Research Facility and the Lawrence Berkeley Laboratory Process with significant throughputs of 10–18 kg/h (dry solids) (Zhu et al., 2014). Table 17.8 presents a summary of continuous reactor systems in the recent literature; it can be seen that the capacities are still relatively low. The development of continuous HTL reactors is still in the early stages and further research needs to be performed to evaluate the best reactor designs, construction materials, and fluid dynamics. Elliott et al. (2015) describe the development of continuous HTL systems in their review in detail. To advance the technical maturity of HTL, PNNL identified specific challenges including reducing the risk of large-scale pumpability, reducing capital costs by moving away from a continuous stirred-tank reactor configuration to a scalable plug-flow reactor configuration, and understanding appropriate materials of construction for process design. A further study by PNNL on continuous HTL (Zhu et al., 2014) on forest residue and corn stover feedstocks demonstrated that nothing other than size reduction (formatting) was required. The process was configured in a PFR design which maintained a small CSTR as a preheater. Within the limits of the corrosion assessment testing, the study showed that stainless steels were suitable for HTL applications.

Table 17.7

Industrially operating HTC plants

ManufacturerStartedReactorConditionsCapacityWebsite
SunCoal2012Continuous200°C; 2 MPa; 6–12 h60,000 t/annumwww.suncoal.de
AVA–CO22010Batch220–230°C; 2.4 MPa; 5–10 h50,000 t/annumwww.ava–co2.com
TFC Engineering2012Continuous200–230°C; 2–2.5 MPa; 3–4 h10,000 t/annumwww.tfc-engineering.li
Terra Nova Energy2010Continuous200°C; 2–3.5 MPa; 4 h8000 t/annumwww.terranova-energy.com
Ingelia SL2010Continuous180–220°C; 1.7–2.4 MPa; 4–8 h6000 t/annumwww.ingeliahtc.com

image

Table 17.8

Summary of continuous HTL reactors in literature

InstitutionFeedstockReactorConditionsCapacityReference
PNNLMicro-, macroalgae, lingoncellulosicsCSTR/PFR350°C; 24–40 min36 L/day (5–35 wt% slurry)Elliott et al. (2013a)
Aarhus UniversityDDGSPFR350°C; 15 min13 L/day (20 wt% slurry)Mørup et al. (2015)
University of SydneyMicroalgaePFR250–350°C; 3–8 min700 L/day (10 wt% slurry)Jazrawi et al. (2013)
University of LeedsMicroalgaePFR350°C; 1–5 min60 L/day (10 wt% slurry)Biller et al. (2015)
University of IllinoisSwine manureCSTR305°C; 40–80 min48 kg slurry/dayOcfemia et al. (2006)

image

PFR, Continuous plug flow; CSTR, continuous stirred-tank reactor.

With funds from the Australian National Collaborative Research Infrastructure Strategy (NCRIS) a continuous flow pilot-scale hydrothermal processing unit became operational at the University of Sydney in 2012. The design of the unit is based on coiled stainless steel tubes submerged into a fluidized sand bath in a plug flow reactor type set-up.
Some noticeable outcomes of the study include: (1) more severe reaction conditions led to the highest yields, lowest oxygen but increased nitrogen contents in the biocrude; (2) higher solids loadings increase biocrude yields, as shown in PNNL studies, and reduce carbon losses within the system; (3) the “inverse scaling” effect of pumping and pressure control led to the conclusion that scaling up of the Sydney design should result in better controllability and reduced potential of formation of agglomerates and deposits, which can lead to blockages.
Researchers at Aarhus University investigated the commissioning of a PFR processing DDGS over 24 h using an induction heater with heating rates of 17°C/min. It was found that the mean residence time varied from the expected value and the system can therefore not be considered a plug flow type reactor but merely a tubular type reactor (Mørup et al., 2015).
Industrial installations of the HTL technology exist but currently limited information of capacities, reactor design, and operating conditions are available. Sapphire Energy Inc. (USA) and Muradel Ply Ltd. (Australia) are examples of companies working on liquid fuel production using HTL from microalgae and Steeper Energy (Demark) are investigating the use of lignocellulosic biomass. Muradel currently has a 3 t/day (30 wt% solids) continuous HTL facility in operation and is planning to increase the capacity to 6.5 t/day and a total output of 50 million L/annum by 2020, while Sapphire Energy is planning on constructing a plant in 2018 to produce ∼200 million L biocrude per annum (Energy, 2013).
There are only limited reports on hydrothermal gasification in larger than laboratory scale. At PNNL bench-scale systems were developed as early as 1989 (Elliott et al., 1989). This work was used to design a scaled-up mobile reactor system, designed for up to 0.5 t of wet feed per day (Elliott et al., 1994). PNNL demonstrated successfully the continuous gasification of biomass to methane-rich product gas. Initially problems occurred with plugging of the catalyst bed by biomass which was solved by a two-step process where biomass was liquefied in a CSTR before HTG. The design temperature of the mobile unit mounted on the back of a pick-up truck was 350°C (Duan and Savage, 2010c). The demonstration unit worked well, but some poisoning of the catalyst was observed.
A continuous HTG reactor has also been developed at the Paul Scherrer Institute (see Fig. 17.7 and references therein). This design aims to remove salts, particularly sulfates prior to the reactor bed. Sulfur has been shown to poison the active catalyst sites and this is one of the main issues of HTG.
At the Karlsruhe Institute of Technology the largest plant for hydrothermal biomass gasification was installed with a throughput of 100 kg/h slurry (∼20 wt% solids). After some “teething troubles” during the beginning of operation (plugging, problems concerning feeding in the biomass), various biomass feedstocks, eg, corn silage, were successfully converted to gas. The VERENA test facility is designed for a maximum of about 700°C and 35 MPa (Kruse, 2009).
Currently the development of HTG reactors is less advanced than for the HTC and HTL pathways and requires expanded process development to take the technology to a scale of industrial demonstration.

17.7. Lifecycle and techno-economic assessment

There has not been any comprehensive techno-economic analysis (TEA) published on hydrothermal carbonization to date; however initial assessments suggest the char material may be more costly to produce compared to torrefaction which represents the equivalent dry route to HTC (Hoekman et al., 2013). One study reports that HTC hydrochar pellets can be produced at around 10 €/GJ from biomass wastes if it is available at zero cost (Erlach et al., 2011). However one of the main advantages of HTC over torrefaction is its ability to utilize wet wastes, which are potentially available with a premium at the gate. Moreover, it has been reported that hydrochars are more amenable to pelletization compared to torrified chars (Reza et al., 2012). The pellets are more robust, which facilitates storage, handling, and transportation. To produce pellets from torrified biomass, binders are typically required to achieve robust pellets, incurring additional cost (Stelte et al., 2011). Economic competiveness could also arise due to solid fuel carbon taxes such as those implemented in Ireland of up to 50 €/t, which biomass-derived solid fuels would be exempt from. Consequently, the most economical applications of HTC appear to be situations in which biomass requires long-distance transport and long-term storage as well as its application for wet biowastes. Additional research and analysis is however required to fully quantify the cost and lifecycle implications of HTC hydrochar.
TEA and LCA studies on HTL are more comprehensive compared to HTC. Several studies have investigated the process including the upgrading of biocrude via hydrotreatment to transportation fuels. The National Alliance for Advanced Biofuels and Bioproducts (NAABB, USA) published a techno-economic analysis for HTL of algal biomass based on experimental results (Zhu et al., 2013; Jones t al., 2014). The primary cost driver was determined to be the algae feedstock cost. Key conclusions relate to the distribution of organic product between the biocrude phase and the aqueous byproduct phase and the need for more efficient biocrude recovery. The studies concluded that for the projected target case, a $4.49/gallon of gasoline equivalent (GGE) was the minimum fuel selling. This amounts to a conversion-only cost (excluding all feedstock-related costs) of $1.18/GGE. Improvement to a $3/GGE minimum fuel selling price would require a combination of improvements, such as upgrading at a central (larger scale) site, coprocessing with other available biomass types, and reducing the cost of the CHG water treatment.
A similar study by Ou et al. (2015) evaluated the techno-economic analysis of transportation fuels via HTL of defatted microalgae (Ou et al., 2015). Similar conclusions were drawn in regards to the minimum fuel selling price (MFSP), which was calculated as 679 $/m3. A sensitivity analysis revealed that the price is most sensitive to the HTL product yield, highlighting the importance of conversion performance. The feedstock cost was also a major uncertainty toward the MFSP which was also highlighted by the study at the PNNL.
A lifecycle analysis (LCA) was published based on HTL of microalgae to green diesel using data from a continuous pilot-scale plant. Three scenarios are analyzed, namely, lab-scale, pilot-scale, and full-scale systems (Liu et al., 2013). The energy return on investment, as well as the greenhouse gas emissions, was calculated and compared to lipid extracted algal biodiesel, petroleum fuels, corn-ethanol, and soy biodiesel. The study was based on data and expertise from Sapphire Energy Inc. (USA), for the production of green diesel using HTL. It was shown that pilot-scale facilities have lifecycle burdens on par with conventional biofuels. However, the results from extrapolating to full-scale facilities were more favorable; greenhouse gas emissions were lower compared to petroleum fuels and corn ethanol. The energy return on investment (EROI) was found to be between 1 and 3 with full-scale production facilities having an EROI of around 2.7. These EROI results are favorable compared to all conventional biofuels but not as high as petroleum-derived fuels.
PNNL also published a TEA on the HTL of woody biomass and upgrading to liquid fuel in a scenario of a 2000 dry metric t/day plant. It was found that using current state-of-technology resulted in an MFSP of $4.44/GGE and $2.52/GGE for the goal case. To advance to the goal case improvements to the loss of organics to the water phase should be implemented (Zhu et al., 2014).
Comprehensive LCA and TEA analysis on HTG is not available in the literature. HTG is still very much in its early development stage and scale-up issues are yet to be addressed in terms of overcoming reactor plugging/corrosion and process economics. However, the technology has demonstrated its economic competitiveness with other hydrogen production methods; Spritzer and Hong (2003) estimate the cost of hydrogen production from HTG of biomass to be about $3/GJ. Similarly Vogel and Waldner (2006) estimated the process economics for a supercritical system and concluded that using wood as a feedstock at $67/t could yield a gas product valued at $10.3/GJ and $6/GJ if zero-cost liquid manures are utilized.
A system analysis was carried out by Ro et al. using swine manure as a feedstock for subcritical gasification. The study suggests that catalytic hydrothermal gasification of wet manures is considerably more expensive compared to anaerobic digestion. However, the authors name several advantages of the technology compared to AD: destruction of pathogens and odorous compounds, production of relatively clean water and the production of valuable byproducts such as ammonia and phosphates for the fertilizer market. Additionally, the high rate of conversion of the organic matter into gas drastically decreases the land requirement for manure application and resulting costs in transportation and tipping fees. The net product gas in Ro et al.’s scenario would have a value $8/GJ (Ro et al., 2007).

17.8. Conclusions

Overall hydrothermal processing is a promising technology for the treatment of wet wastes and wet biomass for the production of solid, liquid, or gaseous fuels. Hydrothermal processing has several advantages over dry thermochemical processing technologies as it can utilize wet feedstocks. The homogeneity of feedstocks is not vital to the technology allowing the use of mixed biomass and sources of biomass with varying composition. Hydrothermal carbonization is the least challenging pathway for hydrothermal processing due to its lower temperatures and pressures from an engineering standpoint. There are several industrial HTC plants currently in operation. Additional research however is required to assess the combustion properties, uses as biochar and storage and handling properties of hydrochar for wide-scale implementation of the technology. The process waters from HTC are high in inorganics and organic carbon and therefore require clean-up, most likely via AD before disposal. This aspect has not been investigated in enough detail, which is one of the reasons for the lack of comprehensive LCA and TEA studies of HTC.
HTL is particularly encouraging for the production of replacement drop-in transportation fuels for the aviation sector as viable alternative technologies are not developed. There have been significant developments in the HTL technology with several continuous pilot-scale facilities in operation. The main challenges associated with HTL are the pumpability of slurries to pressures of around 200 bar. Advanced pumping, wet pretreatment and reactor designs should be investigated to help solve this issue. The O and N contents in biocrudes are still undesirably high and additional development of in situ catalysis should be investigated. Post-HTL upgrading via hydrotreating appears promising but research and development of continuous upgrading facilities and catalyst development are required. To date microalgae have been a feedstock of choice in many hydrothermal processing studies of HTL and HTG due to their small size, ease of pumping, high lipid content, and potentially high area-specific growth yields. Further work should be carried out on biowastes and lignocellulosic feedstocks as the development of microalgae cultivation has not reached the scale required for industrial processing.
HTG has several advantages in the treatment of aqueous wastes due to high carbon gasification efficiencies, reduced land footprint and destruction of pathogens. However the technology is still very much at the research stage and requires continuing process development to take it to the industrial scale. Current bottlenecks include the capital cost of HTG systems, catalyst poisoning, material assessment, and pumping slurries to pressures of up to 300 bar.

References

Anastasakis K, Ross A.B. Hydrothermal liquefaction of the brown macro-alga Laminaria saccharina: effect of reaction conditions on product distribution and composition. Bioresource Technology. 2011;102(7):4876–4883.

Bai X, et al. Hydrothermal catalytic processing of pretreated algal oil: a catalyst screening study. Fuel. 2014;120(0):141–149.

Bandura A.V, Lvov S.N. The ionization constant of water over wide ranges of temperature and density. Journal of Physical and Chemical Reference Data. 2006;35(1):15–30.

Berge N.D, et al. Hydrothermal carbonization of municipal waste streams. Environmental Science & Technology. 2011;45(13):5696–5703.

Biller P, et al. Nutrient recycling of aqueous phase for microalgae cultivation from the hydrothermal liquefaction process. Algal Research. 2012;1(1):70–76.

Biller P, et al. Hydroprocessing of bio-crude from continuous hydrothermal liquefaction of microalgae. Fuel. 2015;159(0):197–205.

Biller P, Ross A.B. Potential yields and properties of oil from the hydrothermal liquefaction of microalgae with different biochemical content. Bioresource Technology. 2011;102(1):215–225.

Biller P, Ross A.B. Pyrolysis GC–MS as a novel analysis technique to determine the biochemical composition of microalgae. Algal Research. 2014;6(Part A(0)):91–97.

Broch A, et al. Analysis of solid and aqueous phase products from hydrothermal carbonization of whole and lipid-extracted algae. Energies. 2013;7(1):62–79.

Brown T.M, Duan P, Savage P.E. Hydrothermal liquefaction and gasification of Nannochloropsis sp.. Energy & Fuels. 2010;24(6):3639–3646.

Cherad R, et al. Macroalgae supercritical water gasification combined with nutrient recycling for microalgae cultivation. Environmental Progress & Sustainable Energy. 2013;32(4):902–909.

Cherad R, et al. Hydrogen production from the catalytic hydrothermal gasification of process water from microalgae liquefaction. Fuel. 2015;166:24–28.

Danso-Boateng E, et al. Hydrothermal carbonisation of sewage sludge: effect of process conditions on product characteristics and methane production. Bioresource Technology. 2015;177(0):318–327.

Dote Y, et al. Recovery of liquid fuel from hydrocarbon-rich microalgae by thermochemical liquefaction. Fuel. 1994;73(12):1855–1857.

Duan P, Savage P.E. Hydrothermal liquefaction of a microalga with heterogeneous catalysts. Industrial & Engineering Chemistry Research. 2010;50(1):52–61.

Duan P, Savage P.E. Catalytic hydrotreatment of crude algal bio-oil in supercritical water. Applied Catalysis B: Environmental. 2011;104(1–2):136–143.

Duan P, Savage P.E. Upgrading of crude algal bio-oil in supercritical water. Bioresource Technology. 2011;102(2):1899–1906.

Elliott D.C, et al. Low-temperature Conversion of High-moisture Biomass: Continuous Reactor System Results. vol. 90. NASA STI/Recon; 1989:14683 Technical Report N.

Elliott D.C, et al. Chemical processing in high-pressure aqueous environments. 4. Continuous-Flow reactor process development experiments for organics destruction. Industrial & Engineering Chemistry Research. 1994;33(3):566–574.

Elliott D.C, et al. Chemical processing in high-pressure aqueous environments. 7. Process development for catalytic gasification of wet biomass feedstocks. Industrial & Engineering Chemistry Research. 2004;43(9):1999–2004.

Elliott D.C, et al. Hydrothermal processing of macroalgal feedstocks in continuous-flow reactors. ACS Sustainable Chemistry & Engineering. 2013;2(2):207–215.

Elliott D.C, et al. Process development for hydrothermal liquefaction of algae feedstocks in a continuous-flow reactor. Algal Research. 2013;2(4):445–454.

Elliott D.C, et al. Hydrothermal liquefaction of biomass: developments from batch to continuous process. Bioresource Technology. 2015;178(0):147–156.

Elliott D.C. Evaluation of wastewater treatment requirements for thermochemical biomass liquefaction. In: Bridgwater A.V, ed. Advances in Thermochemical Biomass Conversion. Netherlands: Springer; 1993:1299–1313.

Elliott D.C. Catalytic hydrothermal gasification of biomass. Biofuels, Bioproducts, and Biorefining. 2008;2(3):254–265.

Elliott D.C. Hydrothermal processing. In: Thermochemical Processing of Biomass. John Wiley & Sons, Ltd; 2011:200–231.

Elliott D.C. Catalytic hydrothermal gasification. In: Biomass Power for the World. Pan Stanford; 2015:665–701.

Energy U.D.o. Sapphire Energy Integrated Algal Biorefinery (IABR). DOE/EE-0841; 2013.

Erlach B, Wirth B, Tsatsaronis G. Co-production of electricity, heat and biocoal pellets from biomass: a techno-economic comparison with wood pelletizing. In: Proceedings of World Renewable Energy Congress (Bioenergy Technology), Sweden. 2011.

Faeth J.L, Valdez P.J, Savage P.E. Fast hydrothermal liquefaction of Nannochloropsis sp. To produce biocrude. Energy & Fuels. 2013;27(3):1391–1398.

Fettig J, Ramke H.-G, Lehmann H.-J, Antonietti M. Machbarkeitsstudie zur Energiegewinnung aus organischen Siedlungsabfällen durch Hydrothermale Carbonisierung Abschlussbericht. Deutsche Bundesstiftung Umwelt; 2010.

Fonts I, et al. Sewage sludge pyrolysis for liquid production: a review. Renewable and Sustainable Energy Reviews. 2012;16(5):2781–2805.

Frans G, Jaap E.N. Biomass to Liquid Fuels via HTU, in Biomass Power for the World. Pan Stanford; 2015:631–664.

Funke A, Ziegler F. Hydrothermal carbonization of biomass: a summary and discussion of chemical mechanisms for process engineering. Biofuels, Bioproducts and Biorefining. 2010;4(2):160–177.

Garcia Alba L, et al. Hydrothermal treatment (HTT) of microalgae: evaluation of the process as conversion method in an algae biorefinery concept. Energy & Fuels. 2011;26(1):642–657.

Garcia Alba L, et al. Microalgae growth on the aqueous phase from hydrothermal liquefaction of the same microalgae. Chemical Engineering Journal. 2013;228(0):214–223.

Guo L.J, et al. Hydrogen production by biomass gasification in supercritical water: a systematic experimental and analytical study. Catalysis Today. 2007;129(3–4):275–286.

Haiduc A, et al. SunCHem: an integrated process for the hydrothermal production of methane from microalgae and CO2 mitigation. Journal of Applied Phycology. 2009;21(5):529–541.

He C, et al. Hydrothermal gasification of sewage sludge and model compounds for renewable hydrogen production: a review. Renewable and Sustainable Energy Reviews. 2014;39(0):1127–1142.

He C, Giannis A, Wang J.-Y. Conversion of sewage sludge to clean solid fuel using hydrothermal carbonization: hydrochar fuel characteristics and combustion behavior. Applied Energy. 2013;111:257–266.

Heilmann S.M, et al. Hydrothermal carbonization of microalgae. Biomass and Bioenergy. 2010;34(6):875–882.

Heilmann S.M, et al. Hydrothermal carbonization of distiller's grains. Biomass and Bioenergy. 2011;35(7):2526–2533.

Hoekman S.K, et al. Hydrothermal carbonization (HTC) of selected woody and herbaceous biomass feedstocks. Biomass Conversion and Biorefinery. 2013;3(2):113–126.

Inoue S, et al. Analysis of oil derived from liquefaction of Botryococcus brauniiBiomass and Bioenergy. 1994;6(4):269–274.

Jazrawi C, et al. Pilot plant testing of continuous hydrothermal liquefaction of microalgae. Algal Research. 2013;2(3):268–277.

Jena U, et al. Evaluation of microalgae cultivation using recovered aqueous co-product from thermochemical liquefaction of algal biomass. Bioresource Technology. 2011;102(3):3380–3387.

Jena U, Das K.C, Kastner J.R. Effect of operating conditions of thermochemical liquefaction on biocrude production from Spirulina platensisBioresource Technology. 2011;102(10):6221–6229.

Jones S.B, et al. Process Design and Economics for the Conversion of Algal Biomass to Hydrocarbons: Whole Algae Hydrothermal Liquefaction and Upgrading. 2014 (p. Medium: ED; Size: PDFN).

Kruse A, Dinjus E. Hot compressed water as reaction medium and reactant: properties and synthesis reactions. The Journal of Supercritical Fluids. 2007;39(3):362–380.

Kruse A, Dinjus E. Hot compressed water as reaction medium and reactant: 2. Degradation reactions. The Journal of Supercritical Fluids. 2007;41(3):361–379.

Kruse A, Funke A, Titirici M.-M. Hydrothermal conversion of biomass to fuels and energetic materials. Current Opinion in Chemical Biology. 2013;17(3):515–521.

Kruse A. Hydrothermal biomass gasification. The Journal of Supercritical Fluids. 2009;47(3):391–399.

Kumar S, Gupta R.B. Biocrude production from switchgrass using subcritical water. Energy & Fuels. 2009;23(10):5151–5159.

Libra J.A, et al. Hydrothermal carbonization of biomass residuals: a comparative review of the chemistry, processes and applications of wet and dry pyrolysis. Biofuels. 2011;2(1):71–106.

Liu X, et al. Pilot-scale data provide enhanced estimates of the life cycle energy and emissions profile of algae biofuels produced via hydrothermal liquefaction. Bioresource Technology. 2013;148(0):163–171.

Luo L, Dai L, Savage P.E. Catalytic hydrothermal liquefaction of soy protein concentrate. Energy and Fuels. 2015;29(5):3208–3214.

Luterbacher J.S, et al. Hydrothermal gasification of waste biomass: process design and life cycle asessment. Environmental Science & Technology. 2009;43(5):1578–1583.

Malins K, et al. Bio-oil from thermo-chemical hydro-liquefaction of wet sewage sludge. Bioresource Technology. 2015;187(0):23–29.

Mørup A, et al. Construction and commissioning of a continuous reactor for hydrothermal liquefaction. Industrial & Engineering Chemistry Research. 2015;54(22):5935–5947.

Müller J.B, Vogel F. Tar and coke formation during hydrothermal processing of glycerol and glucose. Influence of temperature, residence time and feed concentration. The Journal of Supercritical Fluids. 2012;70:126–136.

Nelson M, et al. Microbial utilization of aqueous co-products from hydrothermal liquefaction of microalgae Nannochloropsis oculataBioresource Technology. 2013;136(0):522–528.

Ocfemia, et al. Hydrothermal Processing of Swine Manure into Oil Using a Continuous Reactor System: Development and Testing. St. Joseph, MI, ETATS-UNIS: American Society of Agricultural Engineers; 2006:9.

Oliveira I, Blöhse D, Ramke H.-G. Hydrothermal carbonization of agricultural residues. Bioresource Technology. 2013;142(0):138–146.

Onwudili J.A, et al. Catalytic hydrothermal gasification of algae for hydrogen production: composition of reaction products and potential for nutrient recycling. Bioresource Technology. 2013;127(0):72–80.

Ou L, et al. Techno-economic analysis of transportation fuels from defatted microalgae via hydrothermal liquefaction and hydroprocessing. Biomass and Bioenergy. 2015;72(0):45–54.

Pastor-Villegas J, et al. Study of commercial wood charcoals for the preparation of carbon adsorbents. Journal of Analytical and Applied Pyrolysis. 2006;76(1–2):103–108.

Perry R.H, Green D.W. Perry's Chemical Engineers' Handbook. seventh ed. McGraw-Hill; 1997.

Peterson A.A, et al. Thermochemical biofuel production in hydrothermal media: a review of sub- and supercritical water technologies. Energy & Environmental Science. 2008;1(1):32–65.

Pham M, et al. Effects of hydrothermal liquefaction on the fate of bioactive contaminants in manure and algal feedstocks. Bioresource Technology. 2013;149(0):126–135.

Pienkos P.T, Darzins A. The promise and challenges of microalgal-derived biofuels. Biofuels, Bioproducts and Biorefining. 2009;3(4):431–440.

Ramke H.-G, et al. Hydrothermal carbonization of organic waste. In: Proc., Twelfth International Waste Management and Landfill Symposium, Sardinia, Italy. 2009.

Ramos-Tercero E.A, Bertucco A, Brilman D.W.F. Process water recycle in hydrothermal liquefaction of microalgae to enhance bio-oil yield. Energy & Fuels. 2015;29(4):2422–2430.

Reza M.T, et al. Pelletization of biochar from hydrothermally carbonized wood. Environmental Progress & Sustainable Energy. 2012;31(2):225–234.

Reza M.T, et al. Hydrothermal carbonization: fate of inorganics. Biomass and Bioenergy. 2013;49(0):86–94.

Reza M.T, et al. Evaluation of integrated anaerobic digestion and hydrothermal carbonization for bioenergy production. Journal of Visualized Experiments: JoVE. 2014;88:51734.

Reza M.T, et al. Hydrothermal carbonization of biomass for energy and crop production. Applied Bioenergy. 2014;1(1).

Reza M.T, et al. Hydrothermal carbonization of loblolly pine: reaction chemistry and water balance. Biomass Conversion and Biorefinery. 2014;4(4):311–321.

Reza M.T, Mumme J, Ebert A. Characterization of hydrochar obtained from hydrothermal carbonization of wheat straw digestate. Biomass Conversion and Biorefinery. 2015:1–11.

Ro K.S, et al. Catalytic wet gasification of municipal and animal wastes. Industrial & Engineering Chemistry Research. 2007;46(26):8839–8845.

Ross A.B, et al. Classification of macroalgae as fuel and its thermochemical behaviour. Bioresource Technology. 2008;99(14):6494–6504.

Ross A.B, et al. Investigation of the pyrolysis behaviour of brown algae before and after pre-treatment using PY-GC/MS and TGA. Journal of Analytical and Applied Pyrolysis. 2009;85(1–2):3–10.

Ross A.B, et al. Hydrothermal processing of microalgae using alkali and organic acids. Fuel. 2010;89(9):2234–2243.

Schubert M, et al. Continuous salt precipitation and separation from supercritical water. Part 3: interesting effects in processing type 2 salt mixtures. The Journal of Supercritical Fluids. 2012;61:44–54.

Schubert M, Regler J.W, Vogel F. Continuous salt precipitation and separation from supercritical water. Part 1: type 1 salts. The Journal of Supercritical Fluids. 2010;52(1):99–112.

Schubert M, Regler J.W, Vogel F. Continuous salt precipitation and separation from supercritical water. Part 2. Type 2 salts and mixtures of two salts. The Journal of Supercritical Fluids. 2010;52(1):113–124.

Schumacher M, et al. Hydrothermal conversion of seaweeds in a batch autoclave. The Journal of Supercritical Fluids. 2011;58(1):131–135.

Sevilla M, Fuertes A.B. Chemical and structural properties of carbonaceous products obtained by hydrothermal carbonization of saccharides. Chemistry – A European Journal. 2009;15(16):4195–4203.

Sevilla M, Fuertes A.B. The production of carbon materials by hydrothermal carbonization of cellulose. Carbon. 2009;47(9):2281–2289.

Sevilla M, Fuertes A, Mokaya R. High density hydrogen storage in superactivated carbons from hydrothermally carbonized renewable organic materials. Energy & Environmental Science. 2011;4(4):1400–1410.

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.

Specht F.B.H. Die Anwendung hoher Drucke bei chemischen Vorgängen und eine Nachbildung des Entstehungsprozesses der Steinkohle. Halle an der Saale: Verlag Wilhelm Knapp; 1913:58.

Spritzer M.H, Hong G.T. Supercritical water partial oxidation. In: Proceedings of the 2002 US DOE Hydrogen Program Review. 2003 NREL/CP-570–30535.

Stelte W, et al. Pelletizing properties of torrefied spruce. Biomass and Bioenergy. 2011;35(11):4690–4698.

Stemann J, Putschew A, Ziegler F. Hydrothermal carbonization: process water characterization and effects of water recirculation. Bioresource Technology. 2013;143:139–146.

Stucki S, et al. Catalytic gasification of algae in supercritical water for biofuel production and carbon capture. Energy & Environmental Science. 2009;2(5):535–541.

Titirici M.-M, Antonietti M. Chemistry and materials options of sustainable carbon materials made by hydrothermal carbonization. Chemical Society Reviews. 2010;39(1):103–116.

Titirici M.-M, Thomas A, Antonietti M. Back in the black: hydrothermal carbonization of plant material as an efficient chemical process to treat the CO2 problem? New Journal of Chemistry. 2007;31(6):787–789.

Torri C, et al. Hydrothermal Treatment (HTT) of Microalgae: detailed molecular characterization of HTT oil in view of HTT mechanism elucidation. Energy & Fuels. 2011;26(1):658–671.

Uddin M.H, et al. Effects of water recycling in hydrothermal carbonization of loblolly pine. Environmental Progress & Sustainable Energy. 2014;33(4):1309–1315.

Uematsu M, Frank E.U. Static dielectric constant of water and steam. Journal of Physical and Chemical Reference Data. 1980;9(4):1291–1306.

Umeki K, et al. High temperature steam-only gasification of woody biomass. Applied Energy. 2010;87(3):791–798.

Vardon D.R, et al. Chemical properties of biocrude oil from the hydrothermal liquefaction of Spirulina algae, swine manure, and digested anaerobic sludge. Bioresource Technology. 2011;102(17):8295–8303.

Vogel F, Waldner M.H. Catalytic hydrothermal gasification of woody biomass at high feed concentration. Science in Thermal and Chemical Biomass Conversion. 2006:1001–1012 Bridgwater AV and Boocock DGB.

Wachendorf M, et al. Utilization of semi-natural grassland through integrated generation of solid fuel and biogas from biomass. I. Effects of hydrothermal conditioning and mechanical dehydration on mass flows of organic and mineral plant compounds, and nutrient balances. Grass and Forage Science. 2009;64(2):132–143.

Wagner W, Pruss A. The IAPWS formulation 1995 for the thermodynamic properties of ordinary water substance for general and scientific use. Journal of Physical and Chemical Reference Data. 2002;31(2):387–535.

Waldner M.H, Vogel F. Renewable production of methane from woody biomass by catalytic hydrothermal gasification. Industrial & Engineering Chemistry Research. 2005;44(13):4543–4551.

Waldner M.H, Krumeich F, Vogel F. Synthetic natural gas by hydrothermal gasification of biomass: selection procedure towards a stable catalyst and its sodium sulfate tolerance. The Journal of Supercritical Fluids. 2007;43(1):91–105.

Waldner M.H. Catalytic Hydrothermal Gasification of Biomass for the Production of Synthetic Natural Gas (Ph.D. thesis). Zürich: ETH/PSI; 2007.

Wang L, Shahbazi A, Hanna M.A. Characterization of corn stover, distiller grains and cattle manure for thermochemical conversion. Biomass and Bioenergy. 2011;35(1):171–178.

Wirth B, Mumme J. Anaerobic digestion of waste water from hydrothermal carbonization of corn silage. Applied Bioenergy. 2014;1(1).

Xiao L.-P, et al. Hydrothermal carbonization of lignocellulosic biomass. Bioresource Technology. 2012;118(0):619–623.

Xu L, et al. Assessment of a dry and a wet route for the production of biofuels from microalgae: energy balance analysis. Bioresource Technology. 2011;102(8):5113–5122.

Xu D, Savage P.E. Characterization of biocrudes recovered with and without solvent after hydrothermal liquefaction of algae. Algal Research. 2014;6(Part A(0)):1–7.

Yakaboylu O, et al. Supercritical water gasification of biomass: a literature and technology overview. Energies. 2015;8(2):859.

Zhang L, Xu C, Champagne P. Overview of recent advances in thermo-chemical conversion of biomass. Energy Conversion and Management. 2010;51(5):969–982.

Zhu Y, et al. Development of hydrothermal liquefaction and upgrading technologies for lipid-extracted algae conversion to liquid fuels. Algal Research. 2013;2(4):455–464.

Zhu Y, et al. Techno-economic analysis of liquid fuel production from woody biomass via hydrothermal liquefaction (HTL) and upgrading. Applied Energy. 2014;129(0):384–394.

Zhu Z, et al. Subcritical hydrothermal liquefaction of barley straw in fresh water and recycled aqueous phase. In: World Sustainable Energy Days Next 2014. Springer; 2015:121–128.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset