Chapter 9

Towards an Aviation Fuel Through the Hydrothermal Liquefaction of Algae

S. Raikova1, C.D. Le2, J.L. Wagner1, V.P. Ting3 and C.J. Chuck3,    1Centre for Doctoral Training, Centre for Sustainable Chemical Technologies, Department of Chemical Engineering, University of Bath, Bath, United Kingdom,    2Department of Oil Refining and Petrochemistry, Hanoi University of Mining and Geology, Hanoi, Vietnam,    3Department of Chemical Engineering, University of Bath, Bath, United Kingdom

Abstract

The hydrothermal liquefaction (HTL) of algal biomass is a promising route to viable third-generation biofuels that improves on the energy efficiency of previously examined algal lipid production. The process is highly suitable for both microalgae and seaweed (macroalgae), however, there still needs to be improvement of the Energy Return on Energy Invested before the technology can be commercially adapted. This can be achieved through the integration of biomass cultivation with other services, such as environmental remediation, as well as the valorization of alternative products to form a closed-loop biorefinery. The HTL process utilizes water at sub-/near-critical conditions (200–374°C, 5–28 MPa) as both the reaction medium and solvent for a host of simultaneous reactions, converting algal biomass into a bio-crude oil, alongside a nutrient-rich aqueous phase, a solid char, and a number of gaseous products. The bio-crude oil produced is acidic, highly viscous, and contains high proportions of N and O. While the HTL oil contains fewer heteroatoms than pyrolysis oils, the bio-crude must still be catalytically upgraded to produce various fractions of hydrocarbon fuels, including aviation kerosene. However, the nitrogen content is substantially higher in algal bio-oils than alternative terrestrial oils, and as such, removal of the N-compounds in the bio-crude is one of the most significant challenges that must be overcome in order to advance HTL technology for biofuel production from algae.

Keywords

Microalgae; macroalgae; aviation kerosene; HTL

9.1 Introduction

There are numerous potential benefits of using microalgae as a biofuel feedstock, which address many of the issues associated with both first- and second-generation fuels [1]. Algae have faster growth rates, benefiting from higher photosynthetic efficiencies [2] (~3–8% compared to 0.5% for terrestrial plants), shorter growth cycles, and lower water demands compared to terrestrial plants [3,4], and as algae cultivation does not require arable land, algae can be cultivated without competing with food production. Additionally, the high CO2 fixation capacity of algae means that widespread algal production could potentially be used as a means of sequestering industrial CO2 emissions. These environmental benefits have clear links to potential economic and social benefits, particularly for developing nations [5]. The potential of microalgae and macroalgae as feedstocks for aviation fuels has already been covered in-depth in chapter ‘Feedstocks for Aviation Biofuels’. However, the large-scale implementation of algal technologies as a replacement for fossil fuel resources hinges on improvements in the efficiency of algal production and conversion.

9.2 Hydrothermal Processing of Algal Feedstocks

Most early investigations into algal biofuels focused on adapting existing biodiesel production technologies such as the use of microalgal lipids as a feedstock. An alternative to biodiesel production from algal lipids is thermochemical processing, which refers to processes such as gasification, pyrolysis, or hydrothermal liquefaction (HTL). These techniques utilize the entire algal biomass, including proteins and carbohydrates, to generate algal oils. However, although both gasification and pyrolysis are effective for processing of dry feedstocks such as lignocellulose, they are unappealing from an economic standpoint due to the high mass fraction of water in algal feedstock, and the need for energy-intensive drying [6].

In contrast, thermochemical processes conducted in the presence of water, such as hydrothermal gasification, hydrothermal carbonization, and HTL, are more suitable for wet feedstocks. HTL, in particular, has recently been examined for the processing of whole microalgal feedstocks, and can offer significant energetic cost savings, as lipid content and overall biomass productivity are often inversely related [6]. It utilizes water at sub-/near-critical conditions (200–374°C, 5–28 MPa) as both the reaction medium and solvent for a host of simultaneous reactions, converting algal biomass into a bio-crude oil, alongside a nutrient-rich aqueous phase, a solid char, and a number of gaseous products (Fig. 9.1). The bio-crude oil produced is acidic, highly viscous, and contains high proportions of N and O. Much like pyrolysis oil, this bio-crude can be catalytically upgraded to produce various fractions of hydrocarbon fuels, including aviation kerosene. However, the nitrogen content is substantially higher in algal bio-oils than alternative terrestrial oils, and as such, removal of the N-compounds in the bio-crude is one of the most significant challenges that must be overcome in order to advance HTL technology for biofuel production from microalgae.

image
Figure 9.1 Products of hydrothermal liquefaction.

HTL can be used to process biomass at a concentration of ca. 10–25 wt% in water, reducing the energy consumption of biomass preparation by 88% (with respect to using dry biomass) if a 16% slurry is generated using centrifugation and processed without further drying steps [7]. The mild temperatures used in HTL are well within the range of temperatures encountered in many conventional oil refinery operations [8], and as such, HTL processing for algal biomass can be made energy-efficient and easily scalable. Liu et al. evaluated the life cycle performance of lab-, pilot-, and full-scale scenarios, finding significant improvements in Greenhouse Gas emissions with respect to gasoline and corn ethanol, and a potential Energy Return on Energy Invested of ~2.5 for the full- scale scenario [9], subject to the optimization of a closed-loop system incorporating energy and nutrient recycling.

9.2.1 Reaction Mechanism

HTL employs water at near-critical conditions as a reaction medium and solvent. At near-critical conditions, the solvent properties of liquid water change substantially (including changes in dielectric constant, density, diffusivity, polarity, viscosity, H-bonding, and H+ donor capabilities), transforming it from a polar, highly H-bonded solvent to exhibiting behaviour more typical of a nonpolar organic solvent. In this form, it can act as a solvent for a range of organic reactions and facilitate the breakdown of biomass to bio-crude [10] (Fig. 9.2).

image
Figure 9.2 Hydrothermal liquefaction phase diagram [11].

The prevention of vapour formation at high pressure means the enthalpy associated with the phase change of water is largely reduced, giving vastly increased energy efficiencies for HTL processing over pyrolysis [12].

HTL is comprised of hundreds of simultaneous reactions, which are not well-characterized in the literature. In hot compressed liquid water (near the critical point of 374°C and 22.1 MPa), there are two competitive reactions: hydrolysis and repolymerization [13,14]. The former is more important and dominant at the early stages of the process, when the microalgae is decomposed and depolymerized into small compounds. These are highly reactive, and rapidly polymerize and form liquid bio-crude, gaseous, and solid products [1517]. At higher reaction temperatures and longer reaction times, repolymerization, condensation, and decomposition of substances from different phases may occur. This results in an increase of the solid and gas yields and reduction of the bio-crude yield [1820]. Dehydration reactions, leading to oxygen content reduction, also play a key role in improving the quality of the bio-crude.

9.3 Hydrothermal Liquefaction of Microalgae

Bio-crude yields and properties are dependent on the algal species and its composition. However, operational parameters, such as reaction temperature, heating rate, retention time, the initial biomass loading, and the presence of catalysts also have a significant effect on yields [14]. A wide range of operating conditions have been explored, with liquefaction typically carried out in high pressure batch reactors. Oil yields are generally calculated as a weight percentage of the feedstock weight, using either the dry (d) or the dry ash-free (daf) weight as a basis.

9.3.1 Effect of Biomass Composition on Bio-Crude Oil Production

It has been previously shown that bio-crude yields from HTL of microalgae can be strongly correlated to the biomass composition. The lipid fraction of the biomass is much more readily converted into bio-crude than other biomass components, with a study by Biller and Ross reporting that conversion efficiencies of different algal components to bio-crude oil were in the order lipids > proteins > carbohydrates [15]. This explains the particularly high yields obtained from lipid-rich microalgae. However, as HTL utilizes all biomass components, a high lipid content is not essential for obtaining good bio-crude yields, in contrast to algal biodiesel, which relies entirely on lipids. This effect has been demonstrated by Yu et al., who reported a yield of 39.0% bio-crude oil from an algae strain containing only 0.1% lipids [21]. Utilizing the entire algal biomass for processing lifts the constraints on high-lipid algal strain selection, and can significantly reduce cost and energy requirements of cultivation by using faster-growing strains or algal communities.

Although the specifics of the effects of biomass composition remain unclear, a number of studies have examined the role of the different component fractions (eg, lipids, proteins, and carbohydrates) in bio-crude production via HTL. Most liquefaction processes under optimized conditions have resulted in bio-crude yields around 30–45% [15,2225], regardless of algae strain [12], although, notably, Li et al. obtained yields of 55% for Nannochloropsis sp. under HTL at 260°C for 60 min and at 25% algal loading, and 83% for Chlorella sp. (220°C, 90 min, 25% algal load) [26]. The two algal species have very different compositions (Nannochloropsis contains protein, lipids, and carbohydrates in an approximate ratio of 52:14:22, whereas the same components in Chlorella have a ratio of 9:60:13), but in both cases, bio-crude yield was enhanced with respect to initial lipid content, further reinforcing the advantages of hydrothermal processing over lipid extraction for algal biodiesel [22].

Higher heating values (HHV) of 25–35 MJ/kg are typical for bio-crude, and higher lipid levels in the biomass appear to correspond to higher HHV, although the two properties are not linearly related. Li et al. found, under optimal conditions, Nannochloropsis and Chlorella bio-crude HHVs were 31.5 and 34.2 MJ/kg, respectively, with a maximum HHV of 37 MJ/kg. Although this constitutes a significant increase with respect to the starting biomass (22.4 and 32.3 MJ/kg initially), it still falls short of the energy content of mineral crude (41–48 MJ/kg) [27]. Interestingly, higher bio-crude yields do not necessarily correspond to better oil properties; Nannochloropsis processed at 350°C yielded 34% oil with an HHV of 38.1 MJ/kg, while 46% oil was obtained at a processing temperature of 310°C, with a much lower HHV of 27.7 MJ/kg [25] (Table 9.1).

Table 9.1

Comparison of Documented Bio-Crude Yields, Compositions, and Energy Values

Feedstock Processing Conditions Oil Yield (%) C (%) H (%) N (%) O (%) S (%) HHV References
Spirulina Biomass only 42.26 5.86 3.47 47.26 1.15 20.4 [25]
Spirulina 300°C, 10–12 min, 20% TS, DCM extraction 32.6 68.9 8.9 6.5 14.9 0.86 33.2 [24]
Chlorella 350°C, 60 min, 10% TS 35.8 70.7 6.8 5.9 14.8 0.0 35.1 [23]
Nannochloropsis 350°C, 60 min, 10% TS 34.3 68.1 8.8 4.1 19.0 0.0 34.5 [23]
Spirulina 350°C, 60 min, 10% TS 29 73.3 9.2 7 10.4 0.0 36.8 [15]
Spirulina 220°C, 20 bar, 30 min, 25% TS 38 59.15 5.50 10.47 18.19 1.22 28.7 [25]
Spirulina 310°C, 115 bar, 30 min, 25% TS 38 71.29 8.01 7.66 16.82 0.81 35.2 [25]
Spirulina 350°C, 195 bar, 30 min, 25% TS 30 70.69 8.05 7.22 10.06 0.77 34.3 [25]

Image

TS, Total Solid Fraction.

The numerous simultaneous reactions occurring under HTL conditions lead to a bio-crude containing a diverse range of chemical compounds, the main constituents of which have been found to be C5–C16 cyclic nitrogen compounds, C15–C33 branched and unbranched hydrocarbons, branched oxygenates, aromatic compounds, and heterocycles [26]. Brown et al. carried out a detailed analysis of bio-crude produced by liquefaction of Nannochloropsis sp. Over 90 compounds were detected using GC–MS, although some lighter organics are likely to have been lost during sample extraction, and heavier compounds may not have been detected by GC. A brief overview of the major compounds is presented in Table 9.2.

Table 9.2

List of Major Bio-Crude Components Detected by GC–MS [28]

Major Compound Structure
Indole image
Methylindole image
1-Pentadecene image
Heptadecane image
Isomers of 2-phytene image
Myristic acid image
Phytane image
Palmitoleic acid image
Palmitic acid image
Oleic acid image
Stearic acid image
Cholest-4-ene image
Cholest-5-ene image
Cholesta-3,5,-diene image
Cholesterol image
Cholest-4-en-3-one image
Cholest-4,6-diene-3-one image

ImageImageImage

Elevated heteroatom (O and N) content with respect to mineral crude oil is typical of bio-crude oils [9,10,29]. Higher O and N levels give rise to undesirable fuel properties, such as high acidity and viscosity [24], and the increased diversity of chemical composition can negatively affect combustion performance, storage stability, and economic value [24,29]. The heteroatom content of the bio-crude can generally be attributed to the high O and N content of the starting biomass, although there is not always a direct correlation: Chlorella biomass containing 1.9% N was processed to bio-crude with a 0.3% N content, while Nannochloropsis (7.5% N initially) resulted in a bio-crude with 5.4% N. Chlorella bio-crude was also found to have a higher O content (11.5% compared to 9.5% for Nannochloropsis), although the O content of the biomass was 21.8%, compared to 29.1% for Nannochloropsis [26].

Additionally, the elevated N content of bio-crude of up to 11%, arising from the proteins in the starting biomass (compared to around 0.1, and seldom above 1%, for mineral crude) [30,31], limits its direct usability as a fuel, as it can lead to increased NOx emissions. Significantly, higher nitrogen levels may contribute to catalyst poisoning, making these bio-crude oils unsuitable for co-refining in existing refineries [12,31]. Hence, although bio-crude produced via HTL presents numerous advantages compared to lipid extraction, the product needs significant upgrading to obtain a suitable aviation fuel [24]. An alternative solution posed by Garcia Alba et al. suggests that proteins could be extracted prior to HTL processing and sold as a high-value co-product in a biorefinery paradigm [13]. However, this clearly has implications for bio-crude volumes, particularly when low-lipid algae such as Spirulina are used.

9.3.2 Effect of Microalgal Loading on Product Formation

One of the key benefits of wet biomass processing in comparison to lignocellulosic biomass or dry processing is that microalgae forms a slurry with water, which can be easily pumped and flowed through the reactor. Particularly for continuous HTL operation, determination of the optimal loading of microalgae in the slurry is crucial.

From the literature, many factors such as the scale of the reactor and economics need to be considered with the selection of the microalgal load used in HTL. Hence, there is no clear connection between the microalgal load (total solid fraction, TS) in the feed and the bio-crude yield. Studies of HTL of algae performed over a large range of microalgal loads (ranging from 4% to 50%) have apparently yielded contradictory results [6]. For example, Jena et al. concluded that an increase of the biomass fraction in the feed by 10% resulted in a jump of more than 20% in the bio-crude yield [32], though the opposite findings were reported by Garcia Alba et al. [13].

The optimum ratio of microalgae and water in the feed is presumably dependent on species and reactor design. A high microalgal loading would be economically beneficial, due to reduction of the water content of the feed, leading to a reduction in the energy required for heating the water, and a correspondingly higher volume efficiency and productivity of the HTL reactor [12]. However, there is expected to be an upper limit due to mass transfer limitations and the formation of blockages in continuous thermochemical conversion [33]. Continuous liquefaction studies have shown that while some practical operational issues of reactor blockage may occur at microalgal loads above 5 wt% in the feed [31], processing of microalgae slurries with microalgal loads of up to 35% (obtaining a bio-crude yield of up to 63.6%) are possible [34]. Another key consideration on the loading of microalgae is the cost of drying the original feedstock, which plays a significant role in dictating the most suitable loading achievable.

9.3.3 Effect of Holding Temperature on Microalgal Processing

The holding temperature has a remarkable influence on the yield and properties of HTL products. In general, an initial rise in the holding temperature triggers bio-crude production. After reaching a maximum value for the bio-crude yield, further increase in the holding temperature inhibits biomass liquefaction.

The optimization of holding temperature varies depending on microalgal species; however, in the majority of cases, compared to a bio-crude yield obtained at ca. 350°C, lower holding temperatures (~310–320°C) generated higher bio-crude yields. This effect was observed by Dote et al. [35], and subsequently by Minowa et al. [36], who observed that yields of bio-crude oil from the liquefaction of Dunaliella tertiolecta increased from 30.9% to 43.8% when increasing the holding temperature from 250°C to 300°C, but dropped to 42.6% when processing temperature was increased further to 340°C [36].

These findings are mirrored in the work of Gai et al. for each retention time examined, oil yields from Chlorella pyrenoidosa increased when holding temperature was increased from 260°C to 280°C, and proceeded to decline when reaction temperatures were brought to 320°C [37]. The authors suggested that with increasing temperatures, more secondary decomposition reactions may be triggered, resulting in the production of gases and char rather than oils. Regrettably, they were not able to satisfactorily confirm this using the gas yields obtained in their investigation, owing to the high experimental error.

In addition, the holding temperature has a strong influence on the bio-crude properties: increasing the temperature results in a decline in oxygen content and consequently increases the HHV of the oil [36,38]. The nitrogen content in turn appears to increase with holding temperature, likely due to the promotion of protein decomposition at high temperatures, although the opposite effect has been reported in other studies [21].

The length of time the reaction is held at temperature also has a significant effect on product formation. It is expected that short retention times should be sufficient to degrade the biomass effectively, since hydrolysis and decomposition reactions are expected to be rapid under supercritical conditions [39]. However, there is still a relationship between holding temperature and holding time when bio-crude yield is considered. In many cases, to attain a high bio-crude yield, higher maximum temperatures require shorter holding times. For example, Karagoz et al. determined that at low temperatures (150°C) long holding times favoured the conversion of sawdust to oil, with the opposite trend observed at higher temperatures (250–280°C) [40]. Yu et al. also found that while the overall conversion of biomass and gas yield increased at higher temperature [41], a 120-min hold time was needed at 280°C to produce a bio-crude yield of 39.4%, whereas at 370°C, only 5 min was needed to attain a bio-crude yield of 49.4% [13].

While the holding time has an effect on the bio-crude yield, due to the influence of other operational parameters such as temperature, algae strain, and algal loading, it is difficult to quantify differences across all published studies [42]. There is no consensus on the optimum holding time for continuous operation in the literature. However, as hold time has such a significant effect on the oil yield and composition, this needs to be carefully optimized for continuous HTL reactions, requiring more detailed research and analysis of kinetic data in this area.

9.3.4 Effect of Heating Rate on the Crude Bio-Oil

While the effect of heating rate on the bio-crude yield is also dependent on the algal strain, maximum temperature, and algal load, it has recently been suggested that rapid heating rates and shorter reaction times may result in comparable, if not better, bio-oil yields than slow temperature ramps with longer retention times. For example, bio-crude yields of up to 66% were obtained during the conversion of Nannochloropsis using a holding time of only one minute. This was achieved by placing batch reactors into a sand bath with a temperature significantly above the required set-point temperature, enabling very fast heating rates [43]. This would support conversion by continuous HTL reaction. However, other studies report that while higher heating rates are beneficial for bulk fragmentation of biomass and inhibit char formation, they do not have a great effect on the bio-crude yield [42].

9.3.5 Effect of Catalysts on the HTL of Microalgae

The effect of both homogeneous catalysts, such as organic acids and organic bases, and heterogeneous catalysts, such as metallic oxides, on the HTL of microalgae has been investigated. Homogeneous organic catalysts are desirable since they are inexpensive, do not require separation prior to further upgrading, and reduce the degree of coking [44]. In many cases, HTL experiments have been conducted in the presence of inorganic salts such as sodium carbonate and other water-soluble species [15,44,45]. However, while these catalysts appear to selectively promote the decarboxylation of carbohydrates, they have been found to have a detrimental effect on oil formation from lipids and proteins [15]. One study on the conversion of albumin suggests that the nitrogen partitioning to the oil may be reduced with the presence of sodium carbonate [35], but this observation has not been reported in other similar studies [44].

While homogeneous catalysts are difficult to recycle, heterogeneous catalysts allow post-HTL separation and high reaction selectivity. Liquefaction in the presence of heterogeneous metal catalysts has been reported to give increased liquefaction yields compared to catalyst-free conditions. Furthermore, the presence of the catalyst had a significant influence on the physical properties (eg, the colour and apparent viscosity) of the product oils [45]. The highest oil yields (57%) were achieved with supported Pd catalysts; however, Ni- and Ru-based catalysts produced bio-crudes with the lowest N content [45]. Pt, Ni, and CoMo catalysts were also reported to enhance the degree of deoxygenation, but with reduced bio-crude yields [23].

9.3.6 Effect of Initial Pressure and Additional Reductive Gas

For batch HTL operation, high pressure is required to ensure the water remains in the liquid phase, reduces the enthalpy of phase change of water, and improves the efficiency of the HTL process. Several investigations have introduced additional pressure by charging the reactor with a gas prior to reaction. For batch HTL operation, bio-crude yield was reportedly not affected by the initial pressure [41]. The gases most commonly used are either inert gases (eg, nitrogen, argon) or reducing gases such as CO and H2. While reducing gases are added in order to reduce the oxygen content in the bio-crude, and stabilize the fragmented products of liquefaction, their hazardous nature, gas channelling, and maldistribution need to be carefully considered during the operation [42]. These problems can be overcome by using hydrogen donor solvents, such as tetralin and tetrahydrophenanthrene [42], which have been found to be more favourable than the parent gases in the hydrogenation of biomass fragments. Hydrogen donors (eg, tetralin) were reported by Wang et al. to be more effective than H2 for increasing the bio-crude yield and reducing the level of oxygenates in the bio-crude, which typically corresponds to an increase in HHV and overall improvements in bio-crude quality [46].

9.4 Macroalgae as a Feedstock: Prospects and Challenges

As an alternative to microalgae, seaweed (macroalgae) can be processed through HTL using a similar process to that described above for microalgal feedstocks [47]. While currently underdeveloped as a feedstock, macroalgae has enormous potential as a future feedstock for fuel production (see chapter: Feedstocks for Aviation Biofuels) [48]. Currently, the global consumption of seaweeds for fertilizers, cosmetics ingredients, foodstuffs, and phycocolloids is a multi-billion dollar industry [47]. The majority of macroalgae comes from a relatively small number of species with five genera, Laminaria (also called Saccharina for some species), Undaria, Porphyra, Eucheuma, and Gracilaria, representing 76% of the total tonnage of cultured macroalage [47].

Macroalgae tend to contain a high carbohydrate content and represent a potential source of biomass for bioenergy and chemicals [49]. According to Anastasakis and Ross, wet biomass with an assumed moisture content of 80% requires 10.4 MJ to evaporate the water, equivalent to 87% of the total calorific value of the seaweed (the HHV of the dry biomass of Laminaria saccharina was calculated to be 12 MJ/kg) [47]. Therefore, it has been suggested that only wet processes such as HTL will be energy efficient because of the high energy required to dry the algae [50]. Moreover, macroalgae from marine environments contain high alkali contents, which can generate issues such as slagging and fouling during combustion and potentially during pyrolysis as well [51].

In addition, macroalgae tend to contain higher carbohydrate and much lower lipid contents than microalgae, and, if from a marine origin, can also possess an ash content of up to 25%. This will substantially affect HTL reaction pathways; further research is needed in this area.

All studies to date have focused on batch systems [49], where macroalgal samples were harvested, treated by washing with fresh water, dried, and pulverized prior to the HTL stage [51]. Bio-crude yields up to 23% have been commonly reported from macroalgal HTL [19,50]. However, Bach et al. reported an oil yield of 79% from the HTL of L. saccharina at 350°C with a high heating rate (585°C per minutes) [52]. Neveux et al. found that HTL of Oedogonium spp. and Cladophora spp. (freshwater macroalgae) gave bio-crude yields of 26% and 20%, respectively, on a dry mass basis – higher than marine macroalgae Derbesia spp., which only yielded 20%, though this generated less solid residue than the freshwater macroalgae species [53].

Jin et al. carried out co-liquefaction experiments of microalgae (Spirulina platensis) and macroalgae (Entermorpha prolifera) at 340°C, and discovered optimal synergistic conversion effects for a 50/50 mixture of the two algae. While the fatty acid content of Spirulina strongly improved the conversion of Entermorpha, inorganic salts present in the latter feed did not considerably enhance bio-oil production [54]. In general, a greater amount of energy was recovered in the bio-crude produced, though it contained more nitrogen than the separately produced micro- and macroalgae bio-oils.

Macroalgal biomass has high potential for biofuel production, although there are still significant technological challenges to be overcome. These challenges include cost-effective cultivation and economic harvesting, and more efficient preprocessing of the feedstock. In addition to cleaning, dewatering, and milling steps [55], marine biomass requires desalination in order to prevent equipment corrosion, which can carry a particularly heavy environmental and energetic cost [56]. Additional challenges include developing a continuous HTL reaction system that can readily tolerate high ash content, and designing a biorefinery concept where the supply of aquatic biomass is substantial and secure, and where the aqueous phase can be used as an effective fertilizer. Lenstra et al. reasoned that as the current near-shore cultivation, harvest, and preprocessing costs for macroalgae range between €336 and €669 per tonne, substantial technological development is needed if a sustainable biofuel is to be produced from this feedstock [50,57]. Moreover, the seasonal nature of seaweed growth and culture must be considered for commercial-scale biofuel production, and the technologies of storage or preservation of macroalgae after harvesting need to be developed to enable year-round fuel production.

Similarly to microalgae, the batch experiments only give limited insights into a fully functioning industrial continuous process. Advancement and commercialization of HTL processing of macroalgae require the use of a continuous process with incorporated heat recovery, increasing the energy efficiency and allowing for high throughput. Particular engineering challenges on the processing side include the high salt content of macroalgae and the associated corrosion risk.

The cultivation of macroalgae as a sole biofuel feedstock may not be currently profitable [47]. However, macroalgae contains a diverse range of extractable biochemicals, which could potentially be used to add value to macroalgal fuels in a biorefinery paradigm. The vast majority of farmed macroalgae is currently utilized for human consumption (accounting for 83–90% of the global seaweed industry [55]), but other macroalgae-derived materials also have a significant market value. In particular, seaweed hydrocolloids (predominantly comprising agar, alginate, and carrageenan) account for a significant portion of the remaining value (estimated around $545 million annually) [58,59]. Other macroalgae derivatives are also utilized in the agricultural, chemical, and pharmaceutical industries [60]. Lammens et al. have also evaluated the potential of algal proteins as a source of high-value bio-based chemicals, such as N-methylpyrrolidinone, N-vinylpyrrolidinone, and acrylonitrile [61]. Overall, there is significant scope for extraction of higher-value materials from macroalgae prior to conversion to fuels, and the application of a biorefinery concept could increase the value of the seaweed biomass and improve the economics of production of bioenergy from macroalgae, thereby leading to a more rapid commercial realization of algal-derived biofuels.

9.5 Characteristics of Bio-Crude Oil From Hydrothermal Liquefaction of Algae

The bio-crude oil produced by the HTL of microalgae requires significant upgrading before it can enter the conventional fuel stream and be used to produce aviation kerosene. The high levels of nitrogen (typically between 4% and 8%) and sulphur (up to 1%), in particular, must be almost completely removed if the resulting fuels are to comply with emission standards. The crude bio-oils from HTL of algae can be blended with fossil oil and upgraded to conventional kerosene and other hydrocarbon fuels in a conventional refinery. While it is true that these facilities are capable of processing oils with high levels of sulphur, common industrial catalysts are not able to tolerate the high levels of oxygen and nitrogen seen in the bio-oils. Therefore, at least partial upgrading of the oils is required to reduce the oxygen and nitrogen concentrations to a level comparable to fossil crude oil before they can be sent to a refinery for fractionation.

Detailed chemical analysis of the oils is difficult, as they can contain up to several hundred compounds, including phenols, alkanes, alkenes, fatty acids, ketones, aldehydes, nitriles, amides, and nitrogen heterocycles such as indoles and pyridines [62,63]. In addition, due to the high boiling point range of the oils, most reports have only achieved partial characterization [6,64]. Elemental analysis is therefore one of the most reliable methods to achieve a like-for-like comparison of different crude oils and of the performance of different upgrading studies.

9.5.1 Upgrading of Bio-Crude Oil Produced by the HTL of Microalgae

Only a few studies have investigated the upgrading of HTL oils from microalgae (Table 9.3), with the majority of these studies conducted in batch and achieving only partial nitrogen removal. While the choice of catalyst was found to impact the physical properties of the reaction products, it had little impact on the denitrogenation performance itself [66]. Instead, the degree of denitrogenation appears to be mainly dependent on the reaction temperature, which suggests that these reactions are thermally controlled [66]. Unfortunately, the high temperatures also result in a significant reduction in the hydrogen content of the oils, presumably due to an increase in the aromatic content of the reduction product above desirable levels. Potentially, the poor performance of the selected catalysts under these conditions could lead to insufficient hydrogen being provided for complete hydrogenation. It has also been suggested that the accumulation of ammonia may be limiting the complete reduction of nitrogen compounds [69]. Nonetheless, more recent studies have demonstrated the full catalytic upgrading of bio-crude directly to suitable hydrocarbon fuels. Research has focused on both in situ heterogeneous catalysis and hydrotreatment of the bio-crude after separation [70]. Duan and Savage reported on the use of a 5% Pd/C catalyst for hydrotreatment of algal bio-crude oil, finding that the bio-crude HHV from Nannochloropsis was significantly improved from approximately 37 MJ/kg to 44 MJ/kg, with corresponding improvements in the kinematic viscosity, and significant reductions in N, O, and S content. However, these results were achieved over long retention times of up to 4 h and 80 wt% catalyst loadings, and the yield was substantially decreased [71]. Other studies have centred around supported Pd catalysts, though more recently, Li and Savage reported the upgrading of algal crude oil produced by HTL using HZSM-5, generating a paraffin-like oil composed of >95 wt% C and H, which retained 80% of the energy content of the initial bio-crude and was suitable for use as a liquid fuel [65].

Table 9.3

Summary of Bio-Crude Upgrading Studies Conducted to Date

Feedstock (Microalgae) Bio-Crude Quality (wt%) Catalyst Reaction Temperature Pressure Reaction Mode Reaction Time Product Quality (wt%) References Comments
N S O N S O
Bio-crude from HTL of Nannochloropsis at 350°C for 1 h 5.32 0.56 8.35 HZSM-5 (0–50 wt% loading) 400–500°C 4.35 MPa (hydrogen) Batch 0.5–4 h 1.6–2.71 Bdl 0.39–2.81 [65]  
Bio-crude from HTL of Nannochloropsis at 340°C for 4 h 4.80 0.48 8.07 Pt/C, Mo2C, HZSM-5 (5–20 wt%) 430–530°C 3.5 MPa (hydrogen) Batch in supercritical water (water/oil mass ratio of 4:5) 2–6 h 1.50–3.61 Bdl 0.13–5.31 [66] Temperature has biggest impact on oil properties
Bio-crude from HTL of Nannochloropsis at 320°C for 4 h 4.89 0.68 6.52 Pt/C (25 wt%), HCl, NaOH 400°C 3.4 MPa (hydrogen) Batch in supercritical water (water/oil mass ratio of 1:1) 4 h 2.17–2.79 Bdl 4.31–4.71 [67]  
Bio-crude from HTL of Chlorella sp. at 350°C for 1 h 7.3 Nd 7.8 Pt/γ-Al2O3 (0–40 wt%) HCOOH (0–88 wt%) 400°C 6 MPa (hydrogen) Batch in supercritical water (0–43 wt%) 1 h 2.4–5.8 N/A 4.7–17.9 [8] With formic acid, reacted with the crude algal oil resulting in increased yield
Bio-crude from HTL of Desmodesmus sp. at 200–375°C for 5–60 min 0.2–6.3 Nd 10–39 HZSM-5 (SiO2/Al2O3=280), zeolite:sample mass ratio of 20:1 600°C  Pyrolysis probe (heating rate: 20°C per second, pyrolysis time: 10 s)  0.02–0.14 N/A 0.08–0.30 [68] Elemental calculated from structure of compounds identified by GC
Bio-crudes from HTL of Nannochloropsis sp. at 344–362°C, in continuous reactor 4.0–4.7 0.3–0.5 5.3–8.0 Co-promoted MoS2 on fluorinated alumina support 125–170°C (1st quarter of reactor); 405°C (Main reactor) 13.6 MPa in hydrogen flow Bench-scale hydroprocessing system LHSV of 0.14–0.20 h−1 <0.05–0.16 <50 ppm 0.8–1.2 [34]  

Image

A number of studies have attempted to upgrade crude bio-oils produced by the liquefaction of Nannochloropsis sp. [66,67] or Chlorella sp. [8] in the presence of supercritical water. It was hoped that the reaction of water with hydrocarbons would generate additional hydrogen which, in turn, would promote the removal of heteroatoms from the bio-crude. However, while the presence of water did not appear to have a beneficial effect on the denitrogenation performance, it was found to increase the oxygen content of the reaction product [8]. The best results in the presence of supercritical water were obtained over Pt/C catalyst at 530°C with a reaction time of 6 h, with almost complete sulphur removal and a reduction in nitrogen content from 4.0% to 1.5% [66]. In contrast, upgrading of a liquefaction oil obtained from Nannochloropsis in the absence of water over an HZSM-5 catalyst, at 500°C, and with a reaction time of 4 h, resulted in a reduction in the nitrogen content from 5.3% to 1.6% [65].

Alternatively, Roussis et al. performed experiments to thermally treat HTL bio-crude oil over a temperature range of 350–450°C, with a residence time of 60 min [72]. While the total nitrogen content was not affected by the thermal treatment, the total oxygen content was significantly reduced from 5.7% in the bio-crude to 0.2% in the product oil treated at 400°C. Thermal treatment was also found to reduce the total acid content, as well as the boiling point range of the bio-crude oils, making them more volatile and less viscous. In addition, trace metals were partly removed from the bio-crude. Similar findings were also reported by Bai et al., who used a two-step process to hydrothermally treat the bio-crude oil [73].

These findings demonstrate that it is possible to process the HTL bio-crude in a similar way to heavy crude oil, continuously treating in a number of subsequent processes, including thermal treatment, hydrotreatment, and hydrocracking, to make fractions suitable for further processing towards production of different commercial products, such as gasoline, diesel, and jet fuel.

In a similar effort, promising results were achieved during the continuous hydrotreating of bio-oils produced by the HTL of various strains of Nannochloropsis over sulphided CoMo catalyst supported on fluorinated alumina [34]. Single-stage upgrading at 405°C with a space velocity of 0.20 h−1 and a pressure of 14 MPa in excess hydrogen produced oils with a nitrogen content ranging from 0.07 wt% to 0.25 wt%. An even lower nitrogen content could be achieved by pretreating the oil at temperatures between 125°C and 170°C in the first quarter of the reactor, before conversion at 405°C. While this study demonstrated that almost complete denitrogenation of microalgal crude oils can be achieved under continuous conditions, further studies are required to verify these findings and optimize the reaction conditions.

9.6 Continuous HTL Systems and Challenges in Advancing the Technology

The vast majority of the work to date has focused on batch systems for hydrothermal processing, which, while providing a wealth of information regarding optimal system conditions and chemical mechanisms, is of limited use in developing processing systems to deliver sufficient volumes of algal bio-crude for widespread use as an aviation fuel feedstock. Moreover, the use in most cases of small batch reactors has necessitated use of solvents for the recovery of the bio-crude oil product, leading to complications in determination of the oil yield and distortion of the compositional analysis and properties by the partitioning of solvent-extractable and water-soluble components.

While feeding wet biomass slurry at pressure is highly challenging, in order to advance HTL technology, it is necessary to move to continuous bench/pilot-scale HTL reaction systems. This would help to avoid the excessively long warm-up and cool-down periods that are problematic with batch reactors. In addition, heat exchange/recycling can be integrated into continuous systems, making the process simultaneously more economically feasible and more controllable. To date, only a small number of studies have examined the continuous HTL of biomass [31,34,74], with the majority of these systems not yet approaching a demonstration scale of operation.

Ocfemia et al. developed a small-scale continuous HTL reactor system, including a high-pressure slurry feeder, a process gas (eg, CO) feeder, a continuous stirred tank reactor (CSTR), and a vapour-liquid separation vessel, to process swine manure [74]. A CSTR was selected over a plug-flow reactor for ease of handling, to reduce plugging, and to allow good temperature control throughout the reaction vessel. At a processing temperature of 305°C and a residence time of 80 min, bio-crude oil was produced with an HHV of 31 MJ/kg and bio-crude yields ranging from 60% to 70.4%. The reactor was successfully operated for 16 h continuously with a capacity of up to 48 kg of manure slurry (eg, a solid content of 20%). Although no clogging or accumulation was observed using the CSTR, a pump durability issue was encountered after several experiments due to rapid wearing of the piston and sleeve.

Jazrawi et al. described a continuous system for biomass processing to bio-crude using Chlorella and Spirulina at relatively low algal loading (1–10%) [31]. The design of the HTL reactor differed from conventional continuous reactors in that the slurry was not stirred and the reactants were flowed through coiled stainless steel tubes submerged in a fluidized sand bath. Due to the low biomass loading of some samples, quantitative gravity separation of the bio-crude from the aqueous phase was challenging, and was achieved instead using DCM extraction of the oil phase from the reaction mixture. The group struggled to operate using algal loadings >10%, but nonetheless obtained maximum yields of 41.7% from Chlorella processed at 350°C, comparable to yields obtained under batch conditions. They also found that more severe operating conditions resulted in a decrease in bio-crude O content and a lower-molecular-weight bio-crude. Jazrawi et al. also encountered problems with pumping wet biomass slurry at high pressure [31]. However, they argued that it is more difficult to control the pressure in a smaller slurry system to maintain stable flows, as lower flow rates are accompanied by an increased risk of flow disruption and blockage due to greater incidence of particles damaging the valve seals of smaller pumps. The authors believed that this problem could be reduced if the HTL process were to be implemented at larger scale.

Elliott et al. developed a bench-scale continuous system for HTL of algal biomass in 2013 [34]. The system included the combination of a CSTR and a plug-flow reactor, with integrated modules for catalytic upgrading and sulphur stripping. This more complex hybrid HTL reactor configuration was developed as a direct result of plugging problems experienced previously with a plug-flow reactor system. The solids and bio-crude product were separated without the use of solvents through the use of an in-line filtration unit. The group was able to operate at far higher loadings than Jazrawi et al. (up to 35%), obtaining maximum bio-crude yields of 63.6% using a specially cultured Nannochloropsis strain. The group presented a proposal for a biorefinery, with water and nutrients in the post-HTL water recycled into the algal growth stage, although the model did not explore the potential of using the CO2-rich gaseous phase as the CO2 source to supplement algal cultivation [34].

9.7 Process Integration for an Advanced Biorefinery

Despite the many potential benefits of algal biofuel technology, for costs and sustainability benefits to be optimized, secondary value streams must be considered. Conventional crude oil refineries generate a wide range of products, including paraffin, lubricants, gases, sulphuric acid, petrochemicals, and feedstocks for plastic manufacture, alongside fuels. Similarly, algal fuel production has the potential to co-produce high-value chemicals, such as proteins, vitamins, and trace minerals in addition to liquid fuels [5]. However, few industries have the economic capacity to accommodate high volumes of co-products. The only industries on a similar scale to fuels are mining, agriculture, plastics, and environmental remediation.

9.7.1 Algae Cultivation and Waste Water Treatment

Significant value could be added by combining the microalgal production process with a secondary objective, such as waste water treatment, the recovery of metals from mining waste, or carbon sequestration from power plant effluents. Waste water treatment in particular could provide substantial quantities of algal biomass, without requiring the addition of costly nutrients. The environmental remediation industry has vastly expanded in recent years, with the remediation market in the United States alone generating an estimated $12.8 billion in 2010, of which waste water treatment represented just under 50% [75]. Wastewater treatment via algal remediation lends itself well to synergistic combination with biomass production, and numerous investigations into the cultivation of algae on industrial, municipal, and agricultural waste waters have already been carried out [76,77].

Global municipal waste water production amounts to approximately 300 billion m3, of which just over 50% is currently treated [78]. Assuming an average biomass yield of 1 g/L, and a liquefaction yield of 30%, complete conversion of the existing waste water treatment facilities could result in an annual bio-crude production of up to 90 billion L.

In addition, waste water treatment using microalgae has been proven to be highly effective in reducing the concentration of N and P pollutants, which are used as nutrients, in the effluent [79]. Compared to conventional methods such as chemical precipitation or the production of an activated sludge, forming waste products which are often disposed of by landfill, microalgal treatment provides a much more sustainable route as it allows the efficient recycling of these nutrients [77].

Metal recovery from mining waste could be another lucrative secondary function of microalgae cultivation. A number of studies have explored the concept of algal remediation, and although most studies to date have employed microalgae only as a nonliving adsorbent, remarkable metal concentrations were observed in the recovered biomass, which could potentially be retrieved from the liquefaction residue. For example, one gram of biomass produced from Oedogonium sp. adsorbed up to 145 mg of Pb(II) when exposed to a lead solution containing a metal concentration of 200 mg/L for 90 min, corresponding to a lead recovery of 35% [80]. Chromium absorption by a strain of the filamentous algae Spirogyra was somewhat lower, at only 14.7 mg/g, however the initial concentration of chromium in the treated solution was also substantially depleted, with a reduction of 5 mg/L observed [81]. More recent studies have also looked at the direct cultivation of algae on industrial waste water, with acid mine drainage representing another large-scale source of water pollution in need of continuous remediation [82]. The diatom Planothidium lanceolatum, for example, could be grown at respective Cd, Zn, and Cu concentrations of 0.1, 0.2, and 0.4 mg/L, without significant effects on photosynthesis [83]. Metal toxicity is a clear challenge, however, and must be addressed before living microalgae can be used for mining waste treatment [84].

9.7.2 The Biorefinery Concept and Nutrient Recovery

Within the biorefinery paradigm it is necessary to consider upstream sustainability factors, as well as final product quality. Despite the advantages conferred by HTL, cultivation of algal biomass is still a relatively energy-intensive process and requires high inputs of water, nutrients and CO2 [85]. The HTL paradigm goes some way to solving those issues, though the process must be optimized to render the process economically feasible. As well as optimizing bio-crude yields, it is also vital to maximize the carbon efficiency, retain efficient water and nutrient recycling, and ensure an inexpensive source of CO2. These issues can be at least partially addressed by encompassing nutrient recycling from the aqueous phase, as well as augmenting algal growth using the CO2 from gaseous products [13,32,86]. This nutrient-cycling concept has been discussed in recent literature, and has been described as ‘Environment-Enhancing Energy’ or E2-Energy/E2E by Zhou et al. at the University of Illinois [8789].

A reliable, low-cost water supply is critical to the success of biofuel production from microalgae to fulfil both economic and sustainability criteria [90]. Recirculating water is an obvious route to reducing overall water consumption. For nonthermochemical methods, this comes with a higher risk of infection and growth inhibition by bacteria, fungi, and viruses found in recycled water directly after biomass extraction. HTL confers the additional advantage of destroying biotic toxins (bacteria, viruses, and even prion proteins) in the bio-crude and aqueous phase [10], although nonliving and inorganic growth inhibitors remain.

The HTL route also concentrates trace mineral matter and nutrients, such as nitrogen, phosphorus, and potassium, as well as Fe, Ca, Mg, and polar organics in the aqueous phase [12,38,86]. This could present a route for simultaneous nutrient recycling and remediation of waste water from a synergistic process, which could not be discharged without further treatment.

Recovering phosphorus is crucial for continuing global agriculture; as well as being a dwindling resource in itself, phosphorus is extracted using fossil fuels, therefore peak oil and peak phosphorus may be intrinsically connected [91]. Apart from the impact on mineral resources, the vast quantities of nutrients needed for cultivation can severely affect the energy balance owing to the energy-intensive production [92]. As a result, nutrient provision for algal cultivation is a key sustainability concern [93], and nutrient recovery is a crucial step in making third-generation biofuel production viable [94].

In hydrothermal processing, high protein levels in the feedstock can lead to accumulation of light organics in the aqueous phase [26] (up to 50% of biomass carbon has been found to accumulate in process water) [86], leading to reduced bio-crude yields and poor carbon efficiency. Although this may be detrimental in some ways (in addition to carbon losses from the bio-crude, it has been noted that phenols present in the aqueous process water may inhibit algal growth) [95], some studies have found that elevated levels of total organic carbon (TOC) in algal growth media may actually supplement algal mixotrophic growth, leading to higher biomass yields [96].

Consequently, reusing HTL process water for algal growth may also improve carbon efficiency, with nutrient recovery from the aqueous phase simultaneously being used to treat process waste water, before release into waterways. Biller et al. carried out a comprehensive study in 2012 examining the impact on nutrient recycling potential. Studies on HTL of Chlorella vulgaris, Scenedesmus dimorphus, and Spirulina platensis found significantly higher nutrient levels in HTL process water than in standard growth media, 3N-BBM +V (Table 9.4) [86].

Table 9.4

Nutrient Content of HTL Process Water (Spirulina, 300°C) Compared to Standard Growth Media 3N-BBM +V [86]

Nutrient HTL Process Water 3N-BBM+V
Conc./ppm Conc./ppm
TOC 15123
Total N 8136 124
NH4+image 6295
PO43image 2159 135
K 1506 63
Acetate 7131
NO3image 194 547
Phenols 98

Image

The high levels of ammonium, phosphate, nitrates, and potassium ions are promising for potential nutrient recovery. However, elevated phenol levels could be problematic due to their inhibitory effects on algal growth [86], even at moderate concentrations (100–200 ppm) [97]. Additionally, high protein content in the starting biomass corresponds to elevated levels of nitrogen heterocycles (eg, pyrolidinones, piperidines, pyrroles, indoles) [15,23], thought to be produced via the Maillard reaction between amino acids and glucose/fructose [37]. This could, however, be beneficial if certain strains of algae are capable of utilizing organics in mixotrophic growth.

Several groups have already reported successful algal cultivation using HTL process water [32,84,86,89,98]. Jena et al. found that dilution was necessary to bring nutrient levels down to levels comparable to those in commercial growth media, but found that a 10-fold dilution was still too strong for algal growth to occur [32]. Consequently, Biller et al. performed tests using process water from the HTL of Spirulina at 300°C, diluted by a factor of 50, 100, and 400, for the cultivation of Spirulina, with algal growth measured using chlorophyll absorbance. The finding that no growth occurred at either 50× or 100× dilution (although at this point, nutrient levels are brought down significantly below typical concentrations encountered in growth media) suggests that the concentrations of phenols, fatty acids, and metals such as Ni were still above inhibitory levels. Finally, a 400× dilution was found to be effective, with Spirulina exhibiting very similar growth rates in diluted process water and 3N-BBM +V [86]. Notably, no growth was seen in pure distilled water, confirming the role of the process water as a nutrient source. For the algae Chlorogloeopsis, however, appreciable growth was seen at dilutions as low as 100×, while the cell counts at 400× dilution were actually increased by a third compared to those seen for standard growth media, suggesting that the organic carbon could supplement mixotrophic growth in this case. Growth does not seem to be inhibited by low concentrations of PO43image (for Chlorogloeopsis at 200× dilution). The study demonstrates the potential of HTL process water to be used as a growth medium, although the optimum dilution was different for each strain. Additionally, Pham et al. demonstrated that recycling into algal cultivation systems, coupled with granular activated carbon treatment, was an effective means of removing organic toxins and significantly reducing HTL waste water cytotoxicity for environmental release [98]. Another recent investigation has demonstrated that HTL process water from liquefaction of Spirulina at 350°C is capable of supplementing growth of a strain of algae cultivated in acid mine drainage. The algae demonstrated appreciable growth over 15 days, and was able to utilize 75% of the phosphate provided by the HTL process water (1:100 dilution in synthetic acid mine drainage) [84].

These results suggest that there is significant potential to vastly reduce the amount of additional nutrients required for algal growth by using recycled HTL process water as a growth supplement. If the algal strain at the site is found to exhibit mixotrophic growth, water-soluble organics in the HTL aqueous phase could be incorporated (improving carbon efficiency and biomass yields). Alternatively, water-soluble organics could be extracted prior to process water recirculation and sold as high-value commodity chemicals.

9.8 Conclusions and Future Direction

It is unlikely that the HTL of algae will give an aviation fuel alone, but it could produce a hydrocarbon-rich feedstock for upgrading into the standard fuel streams. This has the advantage of producing a bio-Jet A-1 drop-in replacement, rather than a fuel limited by blend level. Both seaweed and microalgae have potential as feedstocks to sufficiently supply the market, though currently both types of resources have specific issues that must be solved to produce an economic feedstock. There is huge potential for the commercialization of the HTL technology, where algae can be used as a feedstock. However, to advance the technical maturity of HTL, there are still a number of challenges that need to be addressed. Potential solutions could include reducing the risk of large-scale pumpability, reducing capital cost by moving away from a CSTR configuration to a scalable plug-flow reactor configuration, selection of more appropriate materials of construction for process design, upgrading bio-crude oil, and recycling of nutrients from the recovered by-products (P in the solids and N, K, and C in the aqueous phase) for process cost savings and improved sustainability.

References

1. Alam F, Date A, Rasjidin R, Mobin S, Moria H, Baqui A. Biofuel from algae-is it a viable alternative? Proc Eng. 2012;49:221–227.

2. Pirt J. The thermodynamic efficiency (quantum demand) and dynamics of photosynthetic growth. New Phytol. 1986;102:3–37.

3. Rodolfi L, Zittelli GC, Bassi N, et al. Microalgae for oil: strain selection, induction of lipid synthesis and outdoor mass cultivation in a low-cost photobioreactor. Biotechnol Bioeng. 2009;102:100–112.

4. Gouveia L. Microalgae as a Feedstock for Biodiesel Heidelberg: Springer; 2011.

5. Adenle AA, Haslam GE, Lee L. Global assessment of research and development for algae biofuel production and its potential role for sustainable development in developing countries. Energy Policy. 2013;61:182–195.

6. López Barreiro D, Prins W, Ronsse F, Brilman W. Hydrothermal liquefaction (HTL) of microalgae for biofuel production: state of the art review and future prospects. Biomass Bioenergy. 2013;53:113–127.

7. Xu L, Wim Brilman DWF, Withag JA, Brem G, Kersten S. Assessment of a dry and a wet route for the production of biofuels from microalgae: energy balance analysis. Bioresour Technol. 2011;102:5113–5122.

8. Duan P, Bai X, Xu Y, et al. Catalytic upgrading of crude algal oil using platinum/gamma alumina in supercritical water. Fuel. 2013;109:225–233.

9. Liu X, Saydah B, Eranki P, et al. Pilot-scale data provide enhanced estimates of the life cycle energy and emissions profile of algae biofuels produced via hydrothermal liquefaction. Bioresour Technol. 2013;148:163–171.

10. Peterson AA, Vogel F, Lachance RP, Fröling M, Antal Jr MJ, Tester JW. Thermochemical biofuel production in hydrothermal media: a review of sub-and supercritical water technologies. Energy Environ Sci. 2008;1:32–65.

11. J. Wagner, Production of transport fueld via the hydrothermal liquefaction of microalgae and subsequent upgrading reaction (PhD confirmation report), 2014.

12. Tian C, Li B, Liu Z, Zhang Y, Lu H. Hydrothermal liquefaction for algal biorefinery: a critical review. Renewable Sustainable Energy Rev. 2014;38:933–950.

13. Garcia Alba L, Torri C, Samorì C, et al. Hydrothermal treatment (HTT) of microalgae: evaluation of the process as conversion method in an algae biorefinery concept. Energy Fuels. 2012;26:642–657.

14. Zou S, Wu Y, Yang M, Li C, Tong J. Bio-oil production from sub- and supercritical water liquefaction of microalgae Dunaliella tertiolecta and related properties. Energy Environ Sci. 2010;3:1073.

15. Biller P, Ross AB. Potential yields and properties of oil from the hydrothermal liquefaction of microalgae with different biochemical content. Bioresour Technol. 2011;102:215–225.

16. Demirbas A. Mechanisms of liquefaction and pyrolysis reactions of biomass. Energy Convers Manage. 2000;41:633–646.

17. Yang YF, Feng CP, Inamori Y, Maekawa T. Analysis of energy conversion characteristics in liquefaction of algae. Resour Conserv Recycl. 2004;43:21–33.

18. Zou S, Wu Y, Yang M, Li C, Tong J. Thermochemical catalytic liquefaction of the marine microalgae Dunaliella tertiolecta and characterization of bio-oils. Energy Fuels. 2009;23:3753–3758.

19. Zhou D, Zhang L, Zhang S, Fu H, Chen J. Hydrothermal liquefaction of macroalgae Enteromorpha prolifera to bio-oil. Energy Fuels. 2010;24:4054–4061.

20. Anastasakis K, Ross AB. Hydrothermal liquefaction of the brown macro-alga Laminaria Saccharina: effect of reaction conditions on product distribution and composition. Bioresour Technol. 2011;102:4876–4883.

21. Yu G, Zhang Y, Schideman L, Funk T, Wang Z. Distributions of carbon and nitrogen in the products from hydrothermal liquefaction of low-lipid microalgae. Energy Environ Sci. 2011;4:4587–4595.

22. Yu G, Zhang Y, Schideman L, Funk TL, Wang Z. Hydrothermal liquefaction of low lipid content microalgae into bio-crude oil. Am Soc Agric Biol Eng. 2011;54:239–246.

23. Biller P, Riley R, Ross AB. Catalytic hydrothermal processing of microalgae: decomposition and upgrading of lipids. Bioresour Technol. 2011;102:4841–4848.

24. Vardon DR, Sharma BK, Scott J, et al. Chemical properties of biocrude oil from the hydrothermal liquefaction of Spirulina algae, swine manure, and digested anaerobic sludge. Bioresour Technol. 2011;102:8295–8303.

25. Toor SS, Reddy H, Deng S, et al. Hydrothermal liquefaction of Spirulina and Nannochloropsis salina under subcritical and supercritical water conditions. Bioresour Technol. 2013;131:413–419.

26. Li H, Liu Z, Zhang Y, et al. Conversion efficiency and oil quality of low-lipid high-protein and high-lipid low-protein microalgae via hydrothermal liquefaction. Bioresour Technol. 2014;154:322–329.

27. Speight JG. Handbook of Petroleum Analysis first ed. New York, NY: Wiley Interscience; 2001.

28. Brown TM, Duan P, Savage PE. Hydrothermal Liquefaction and Gasification of Microalga Nannochloropsis sp. Energy Fuels. 2010;24:3639–3646.

29. Huber GW, Iborra S, Corma A. Synthesis of transportation fuels from biomass: chemistry, catalysts, and engineering. Chem Rev. 2006;106:4044–4098.

30. Ball JS, Whisman ML, Wenger WJ. Nitrogen content of crude petroleums. Ind Eng Chem. 1951;43:2577–2581.

31. Jazrawi C, Biller P, Ross AB, Montoya A, Maschmeyer T, Haynes BS. Pilot plant testing of continuous hydrothermal liquefaction of microalgae. Algal Res. 2013;2:268–277.

32. Jena U, Vaidyanathan N, Chinnasamy S, Das KC. Evaluation of microalgae cultivation using recovered aqueous co-product from thermochemical liquefaction of algal biomass. Bioresour Technol. 2011;102:3380–3387.

33. Hirano A, Hon-Nami K, Kunito S, Hada M, Ogushi Y. Temperature effect on continuous gasification of microalgal biomass: theoretical yield of methanol production and its energy balance. Catal Today. 1998;45:399–404.

34. Elliott DC, Hart TR, Schmidt AJ, et al. Process development for hydrothermal liquefaction of algae feedstocks in a continuous-flow reactor. Algal Res. 2013;2:445–454.

35. Dote Y, Sawayama S, Inoue S, Minowa T, Yokoyama S. Recovery of liquid fuel from hydrocarbon-rich microalgae by thermochemical liquefaction. Fuel. 1994;73:1855–1857.

36. Minowa T, Yokoyama S, Kishimoto M, Okakura T. Oil production from algal cells of Dunaliella tertiolecta by direct thermochemical liquefaction. Fuel. 1995;74:1735–1738.

37. Gai C, Zhang Y, Chen W-T, Zhang P, Dong Y. Energy and nutrient recovery efficiencies in biocrude oil produced via hydrothermal liquefaction of Chlorella pyrenoidosa. RSC Adv. 2014;4:16958.

38. Ross AB, Biller P, Kubacki ML, Li H, Lea-Langton A, Jones JM. Hydrothermal processing of microalgae using alkali and organic acids. Fuel. 2010;89:2234–2243.

39. Sasaki M, Adschiri T, Arai K. Production of cellulose II from native cellulose by near- and supercritical water solubilization. J Agric Food Chem. 2003;51:5376–5381.

40. Karagoz S, Bhaskar T, Muto A, Sakata Y, Azhar Uddin MD. Low temperature hydrothermal treatment of biomass: effect of reaction parameters on products and boiling point distributions. Energy Fuels. 2004;18:234–241.

41. Yu G, Zhang Y, Schideman L, Funk TL, Wang W. Hydrothermal liquefaction of low lipid content microalgae into bio-crude oil. Trans ASABE. 2011;54:239–246.

42. Akhtar J, Amin NAS. A review on process conditions for optimum bio-oil yield in hydrothermal liquefaction of biomass. Renewable Sustainable Energy Rev. 2011;15:1615–1624.

43. Faeth JL, Valdez PJ, Savage PE. Fast hydrothermal liquefaction of Nannochloropsis sp to produce biocrude. Energy Fuels. 2013;27:1391–1398.

44. Yeh TM, Dickinson JG, Franck A, Linic S, Thompson Jr LT, Savage PE. Hydrothermal catalytic production of fuels and chemicals from aquatic biomass. J Chem Technol Biotechnol. 2013;88:13–24.

45. Han JX, Duan JZ, Chen P, Lou H, Zheng XM, Hong HP. Nanostructured molybdenum carbides supported on carbon nanotubes as efficient catalysts for one-step hydrodeoxygenation and isomerization of vegetable oils. Green Chem. 2011;13:2561–2568.

46. Wang G, Li W, Li B, Chen H. Direct liquefaction of sawdust under syngas. Fuel. 2007;86:1587–1593.

47. Milledge JJ, Smith B, Dyer PW, Harvey P. Macroalgae-derived biofuel: a review of methods of energy extraction from seaweed biomass. Energies. 2014;7:7194–7222.

48. Lundquist T, Woertz I, Quinn N, Benemann J. A Realistic Technology and Engineering Assessment of Algae Biofuel Production Berkley, CA: Energy Biosciences Institute, University of California; 2010.

49. Biller P, Ross AB. Hydrothermal processing of algal biomass for the production of biofuels and chemicals. Biofuels. 2012;3:603–623.

50. Murphy F, Devlin G, Deverell R, McDonnell K. Biofuel production in Ireland—an approach to 2020 targets with a focus on algal biomass. Energies. 2013;6:6391–6412.

51. Anastasakis K, Ross AB. Hydrothermal liquefaction of four brown macro-algae commonly found on the UK coasts: an energetic analysis of the process and comparison with bio-chemical conversion methods. Fuel. 2015;139:546–553.

52. Bach QV, Sillero MV, Tran KQ, Skjermo J. Fast hydrothermal liquefaction of a norwegian macro-alga: screening tests. Algal Res. 2014;6:271–276.

53. Neveux N, Yuen AKL, Jazrawi C, et al. Bioresour Technol. 2014;155:334–341.

54. Jin B, Duan P, Xu Y, Wang F. Biocrude yield and productivity from the hydrothermal liquefaction of marine and freshwater green macroalgae. Bioresour Technol. 2013;149:103–110.

55. G. Roesijadi, S.B. Jones, Y. Zhu, Macroalgae as a biomass feedstock: a preliminary analysis, pacific northwest national laboratory. Report no.: PNNL-19944. Sponsored by the US Department of Energy; September 2010, 2010.

56. Bruton T, Lyons H, Lerat Y, Stanley M, Rasmussen MB. A Review of the Potential of Marine Algae as a Source of Biofuel in Ireland Dublin: Sustainable Energy Ireland; 2009.

57. W.J. Lenstra, J.H. Reith, J.W.V. Hal. Proceedings of the 4th International Algae Congress, Amsterdam, The Netherlands, 1–2 December 2010.

58. D.J. McHugh, A guide to the seaweed industry, food and agricultural organization: 106, 2003.

59. R. Morchio, C. Cáceres, Macroalgae current state in Latin America. International Workshop on Sustainable Bioenergy from Algae. Berlin, Germany, 2009.

60. Jung KA, Lim S-R, Kim Y, Park JM. Potentials of macroalgae as feedstocks for biorefinery. Bioresour Technol. 2013;135:182–190.

61. Lammens TM, Franssen MCR, Scott EL, Sanders JPM. Availability of protein-derived amino acids as feedstock for the production of bio-based chemicals. Biomass Bioenerg. 2012;44:168–181.

62. Hu Z, Zheng Y, Yan F, Xiao B, Liu S. Bio-oil production through pyrolysis of Blue-Green Algae Blooms (BGAB): product distribution and bio-oil characterization. Energy. 2013;52:119–125.

63. Jena U, Das KC, Kastner JR. Effect of operating conditions of thermochemical liquefaction on biocrude production from Spirulina platensis. Bioresour Technol. 2011;102:6221–6229.

64. Harman-Ware AE, Morgan T, Wilson M, et al. Microalgae as a renewable fuel source: fast pyrolysis of Scenedesmus sp. Renewable Energy. 2013;60:625–632.

65. Li Z, Savage PE. Feedstocks for fuels and chemicals from algae: treatment of crude bio-oil over HZSM-5. Algal Res. 2013;2:154–163.

66. Duan P, Savage PE. Catalytic treatment of crude algal bio-oil in supercritical water: optimization studies. Energy Environ Sci. 2011;4:1447.

67. Duan P, Savage PE. Upgrading of crude algal bio-oil in supercritical water. Bioresour Technol. 2011;102:1899–1906.

68. Torri C, Fabbri D, Garcia-Alba L, Brilman DWF. Upgrading of oils derived from hydrothermal treatment of microalgae by catalytic cracking over H-ZSM-5: a comparative Py–GC–MS study. J Anal Appl Pyrolysis. 2013;101:28–34.

69. Yuan P-Q, Cheng Z-M, Zhang X-Y, Yuan W-K. Catalytic denitrogenation of hydrocarbons through partial oxidation in supercritical water. Fuel. 2006;85:367–373.

70. Duan P, Savage PE. Hydrothermal liquefaction of a microalga with heterogeneous catalysts. Ind Eng Chem Res. 2011;50:52–61.

71. Duan P, Savage PE. Catalytic hydrotreatment of crude algal bio-oil in supercritical water. Appl Catal B Environ. 2011;104:136–143.

72. Roussis SG, Cranford R, Sytkovetskiy N. Thermal treatment of crude algae oils prepared under hydrothermal extraction conditions. Energy Fuels. 2012;26:5294–5299.

73. Bai X, Duan P, Xu Y, Zhang A, Savage PE. Hydrothermal catalytic processing of pretreated algal oil: a catalyst screening study. Fuel. 2014;120:141–149.

74. Ocfemia KS, Zhang Y, Funk T. Hydrothermal processing of swine manure into oil using a continuous reactor system: development and testing. Am Soc Agric Biol Eng. 2006;49:533–541.

75. United States International Trade Commission, Environmental and Related Services—Investigation No. 332-53, USITC Publication 4389, 2013.

76. Chen WT, Zhang Y, Zhang J, et al. Hydrothermal liquefaction of mixed-culture algal biomass from wastewater treatment system into bio-crude oil. Bioresour Technol. 2014;152:130–139.

77. Pittman JK, Dean AP, Osundeko O. The potential of sustainable algal biofuel production using wastewater resources. Bioresour Technol. 2011;102:17–25.

78. Aquastat, Food and Agriculture Organization of the United Nations, 2014.

79. Wang L, Min M, Li Y, et al. Cultivation of green algae Chlorella sp in different wastewaters from municipal wastewater treatment plant. Appl Biochem Biotechnol. 2010;162:1174–1186.

80. Gupta VK, Rastogi A. Biosorption of lead(II) from aqueous solutions by non-living algal biomass Oedogonium sp and Nostoc sp.—a comparative study. Colloids Surf B Biointerfaces. 2008;64:170–178.

81. Gupta VK, Shrivastava AK, Jain N. Biosorption of chromium (VI) from aqueous solutions by green algae Spirogyra species. Water Res. 2001;35:4079–4085.

82. Johnson DB, Hallberg KB. Acid mine drainage remediation options: a review. Sci Total Environ. 2005;338:3–14.

83. Sbihi K, Cherifi O, El gharmali A, Oudra B, Aziz F. Accumulation and toxicological effects of cadmium, copper and zinc on the growth and photosynthesis of the freshwater diatom Planothidium lanceolatum (Brébisson) lange-bertalot: a laboratory study. J Mater Environ Sci. 2012;3:497–506.

84. Raikova S, Smith-Baedorf H, Bransgrove R, et al. Assessing hydrothermal liquefaction for the production of bio-oil and enhanced metal recovery from microalgae cultivated on acid mine drainage. Fuel Process Technol. 2016;142:219–227.

85. Aitken D, Antizar-Ladislao B. Achieving a green solution: limitations and focus points for sustainable algal fuels. Energies. 2012;5:1613–1647.

86. Biller P, Ross AB, Skill SC, et al. Nutrient recycling of aqueous phase for microalgae cultivation from the hydrothermal liquefaction process. Algal Res. 2012;1:70–76.

87. Zhou Y, Schideman L, Yu G, Zhang Y. A synergistic combination of algal wastewater treatment and hydrothermal biofuel production maximized by nutrient and carbon recycling. Energy Environ Sci. 2013;6:3765.

88. Y. Zhou, L. Schideman, Y. Zhang, G. Yu, In: Proceedings of the Water Environment Federation, 2011, pp. 7268–7282(15).

89. Zhou Y, Schideman L, Zhang Y, Yu G, Wang Z, Pham M. Resolving bottlenecks in current algal wastewater treatment paradigms: a synergistic combination of low-lipid algal wastewater treatment and hydrothermal liquefaction for large-scale biofuel production. Proc Water Environ Fed. 2011;6:347–361.

90. Slade R, Bauen A. Micro-algae cultivation for biofuels: cost, energy balance, environmental impacts and future prospects. Biomass Bioenergy. 2013;53:29–38.

91. Rhodes CJ. Peak phosphorus—peak food? The need to close the phosphorus cycle. Sci Prog. 2013;96:109–152 (44).

92. Bagnoud-Velásquez M, Schmid-Staiger U, Peng G, Vogel F, Ludwig C. First developments towards closing the nutrient cycle in a biofuel production process. Algal Res. 2015;8:76–82.

93. National Research Council of the National Academies. Sustainable Development of Algal Biofuels in the United States Washington, DC: The National Academies Press; 2012.

94. Chisti Y. Constraints to commercialization of algal fuels. J Biotechnol. 2013;167:201–214.

95. Nakai S, Inoue Y, Hosomi M. Growth inhibition effects and inducement modes by plant-producing phenols. Water Res. 2001;35:1855–1859.

96. Bhatnagar A, Chinnasamy S, Singh M, Das KC. Renewable biomass production by mixotrophic algae in the presence of various carbon sources and wastewaters. Appl Energy. 2011;88:3425–3431.

97. Scragg AH. The effect of phenol on the growth of Chlorella vulgaris and Chlorella VT-1. Enzyme Microb Technol. 2006;39(4):796–799.

98. Pham M, Schideman L, Scott J, Rajagopalan N, Plewa MJ. Chemical and biological characterization of wastewater generated from hydrothermal liquefaction of Spirulinau. Environ Sci Technol. 2013;47(4):2131–2138.

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

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