Next Article in Journal
Occurrence and Removal of Microplastics in Wastewater Treatment Plants: Perspectives on Shape, Type, and Density
Previous Article in Journal
Surface Modification of PVDF and PTFE Hollow Fiber Membranes for Enhanced Nitrogen Removal in a Membrane-Aerated Biofilm Reactor
Previous Article in Special Issue
Analysis on Operation and Water Quality Characteristics of Centralized Wastewater Treatment Plants of Industrial Parks in Yellow River Basin, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrothermal Carbonization Technology for Wastewater Treatment under the “Dual Carbon” Goals: Current Status, Trends, and Challenges

1
Beijing Key Laboratory of Farmland Soil Pollution Prevention and Remediation, Key Laboratory of Plant-Soil Interactions of Ministry of Education, College of Resources and Environmental Sciences, China Agricultural University, Beijing 100193, China
2
Agronomy Department, Faculty of Agriculture, Assiut University, Assiut 71526, Egypt
3
Department of Biology, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
*
Author to whom correspondence should be addressed.
Water 2024, 16(12), 1749; https://doi.org/10.3390/w16121749
Submission received: 29 May 2024 / Revised: 18 June 2024 / Accepted: 19 June 2024 / Published: 20 June 2024
(This article belongs to the Special Issue Sustainable Wastewater Treatment and the Circular Economy)

Abstract

:
Hydrothermal carbonization (HTC) technology transforms organic biomass components, such as cellulose and lignin, into valuable carbon materials, gases and inorganic salts through hydrolysis, degradation and polymerization, with significant advantages over traditional methods by reducing energy consumption, lowering pollutant emissions and enhancing carbonization efficiency. In the context of global climate change, HTC plays a critical role in water environment management by addressing industrial, agricultural, and domestic wastewater challenges. The application of HTC extends to wastewater treatment, where hydrochar effectively adsorbs heavy metals, organic compounds, and anions, thereby improving water quality. However, challenges remain, such as optimizing the process for diverse raw materials, managing economic costs, and addressing environmental and social impacts. Future research and policy support are essential for advancing HTC technology. By enhancing reaction mechanisms, developing catalysts, and promoting international cooperation, HTC can significantly contribute towards achieving carbon neutrality goals and fostering sustainable development.

1. Introduction

As global climate change intensifies, countries around the world, including China, have set carbon reduction targets [1,2,3,4,5]. As the world’s largest emitter of carbon, China has formulated a dual-carbon policy, namely achieving carbon peak and carbon neutrality, which not only demonstrates its responsibility as a major nation but also introduces unprecedented challenges and opportunities across various industries [6,7,8,9,10,11]. To achieve these targets, scientists need to work with industries, policymakers and community stakeholders to reduce net emissions, and concomitantly develop and implement carbon drawdown strategies [12,13,14,15,16,17]. The emerging field of carbon-based materials boasts new opportunities [15,18,19,20]. In this context, hydrothermal carbonization, as an efficient and eco-friendly technology for resource utilization, holds broad prospects for application in wastewater treatment [21,22].
In general, (hydro)thermal techniques can be divided into two main categories: thermal treatments (operating under inert atmosphere and dry conditions) and hydrothermal treatments (operating in a closed pressurized system and under wet conditions). Each category can be further divided into carbonization, liquefaction, and gasification treatments, based on the operating temperature and phase partitioning of the products (Figure 1). In this paper, the hydrothermal carbonization treatment will be mainly introduced [23].
Hydrothermal carbonization (HTC), a recently prominent carbonization technology, is increasingly recognized for its applications in environmental protection [21,24]. This process utilizes water as a thermal medium in a sealed high-temperature, high-pressure environment to transform waste into useful substances such as carbon materials, gases, and inorganic salts [25,26,27]. Compared to traditional carbonization methods, HTC offers significant advantages, including lower energy consumption, reduced pollutant emissions, higher carbonization efficiency, and greater value-added products [22,27].
As urban modernization progresses rapidly, living standards in cities have significantly improved. Concurrently, China’s water environments are increasingly stressed. The discharge of industrial, agricultural, and domestic wastewater has progressively increased, exacerbating water pollution [28,29,30]. In addition, the development of the pharmaceutical industry, animal husbandry and the unreasonable discharge of wastewater, is causing antibiotic pollution and it is particularly prominent in the water environment [31]. The “China Ecological Environment Statistical Yearbook (2018–2023)” published by the Ministry of Ecology and Environment reports high pollutant levels in discharged wastewater, heavily impacting water environments. Industrial and agricultural wastewater discharges have significantly polluted rivers and lakes, harming the ecosystem [19,20,28,32,33]. Consequently, effectively addressing water environmental issues has become an urgent problem. Figure 2 is a statistical chart of the number of papers on the application of hydrochar in wastewater treatment from 2014 to 2024 in June. In 2014–2019, the number of papers in this area is small, and the number of papers after 2020 increases significantly, indicating that in recent years, people have conducted more and more research on the application of hydrochar for wastewater treatment, which reflects the development of this technology, the increasing attention to water environment problems, and the acceleration of the practical application and transformation of research results. In Figure 3, it can see the main keywords used in articles on applications and use the search term, on the scientific network. The terms “hydrochar yield”, “higher heating value”, “mechanism”, “adsorption” and “functional group” are more controllable terms, which may indicate that the adsorption application of hydrochar is related to the mechanism of hydrochar, the type of functional groups, and the yield and calorific value of hydrochar also have indispensable value for its application.
In response to global climate challenges and to achieve the “Dual Carbon” goals of peaking and neutralizing carbon emissions, continuous innovation in wastewater management technologies is essential. HTC, as an emerging treatment technology, offers significant advantages and broad application prospects. This paper explores the current applications of HTC in wastewater treatment, including its effectiveness against various pollutants, technological processes, and engineering examples. We also analyze the principles and historical development of this technology, addressing current challenges, research directions, and policy recommendations. Through this study, we aim to advance the development and application of HTC, thereby significantly contributing to water environment management and protection. We believe that with ongoing technological advancements and policy improvements, HTC will play a crucial role in achieving the “Dual Carbon” goals and promoting sustainable development.

2. HTC Technology

2.1. Basic Principle

HTC is a process that converts biomass into carbon-rich materials using a high-temperature, high-pressure aqueous environment [24,26,27]. This technology operates in the absence of oxygen, preventing combustion and oxidation, thus preserving the carbon content of the raw materials. It not only transforms waste into valuable resources but also reduces environmental pollution. The carbon materials produced take the advantages of low cost, excellent chemical stability, large surface area and abundant pores [34].
HTC is a thermochemical conversion technique that processes biomass with water at specific temperatures (180–250 °C) and pressures (2–10 MPa) into high-value carbon products [32,35], along with significant liquid and minor gaseous byproducts, primarily CO2 [36,37]. This technology effectively converts fibrous materials found in animal manure and crop residues, including cellulose, hemicellulose, and lignin [27]. HTC involves the decomposition of macromolecules into smaller molecules, which then re-polymerize [37]. The process includes hydrolysis, dehydration, decarboxylation, condensation, and aromatization [38,39]. Cellulose, for example, undergoes hydrolysis where C-O and C-C bonds break, leading to the formation of intermediates like 5-hydroxymethylfurfural, followed by dehydration and aromatization to form hydrophobic microsphere cores and hydrophilic shells through aldol condensation [40]. Similarly, hemicellulose and lignin undergo specific breakdown and reformation processes to produce high-quality, high energy-dense hydrochar [41,42]. HTC’s strength lies in the precise control of reaction parameters (temperature, pressure, feedstock particle size, and duration) enabling the production of superior quality carbons with wider application potential [43,44]. In Figure 4, the main mechanism of hydrochar formation is highlighted by describing the HTC reaction pathways from lignocellulosic biomass.
HTC traces its origins to the early 19th century when the formation mechanisms of coal began to be studied. In Figure 5, the history of HTC technology is described. In 1913, the German chemist Friedrich Bergius and his colleagues conducted carbonization of cellulose under hydrothermal conditions at temperatures ranging from 250 to 310 °C, resulting in a black carbonaceous material with a significantly reduced oxygen-to-carbon (O/C) atomic ratio compared to the raw material. Subsequently, researchers expanded the range of feedstocks to other biomass materials and systematically studied HTC processes.
In the 1970s, HTC technology began to garner attention from scientists and researchers, who explored the transformation of biomass into carbonaceous products using high-temperature, high-pressure water. These studies primarily focused on the thermal decomposition reactions of lignocellulosic biomass [46].
In the 1990s, HTC technology was increasingly applied to waste management and energy conversion. Researchers conducted successful studies on various types of biomasses, including food waste, agricultural residues, and other discarded plant materials, achieving promising results. Since the 21st century, with the increasing prominence of energy and environmental concerns, HTC has emerged as a viable method for waste management and energy conversion. As technology has advanced, there has been a deeper understanding and research into its applications and the value of its products.

2.2. Main Types and Methods of Hydrochar

HTC, as a green and sustainable method for carbon material production, has garnered extensive attention and research in China. Based on various process conditions and operational principles, this technology can be categorized into the following main types:

2.2.1. Ordinary HTC

In a closed system, biomass is mixed with water and carbonized under specific temperature and pressure conditions, resulting in a product that is black, solid, rich in oxygenated functional groups, and primarily composed of carbon. The standard HTC process is relatively simple and efficiently utilizes biomass resources.

2.2.2. HTC Using Catalysts

During the resource recovery process from waste, optimizing energy consumption is essential to accommodate variations in biomass feedstocks and enhance the yield and quality of hydrochar [47]. In the reaction, the rate of biomass dehydration and carbon product formation is generally low, particularly when using plant biomass, which results in hydrochar with lower carbon content, surface area, porosity, and thermal stability, limiting its practical applications [48]. Introducing catalysts can improve the efficiency and quality of the carbonization process. Catalysts facilitate the degradation and carbonization of organic materials in biomass, yielding carbon materials with more complex structures and superior properties. Common catalysts include salts, acids, metal oxides, molecular sieves, and hydroxides [47]. In Table 1, some studies are described to better express the effect of using catalysts on HTC technology.
Petrovic et al. [49] utilized grape residue to prepare hydrochar by KOH modification under the condition of 220℃ and 1 h of hydrothermal temperature. Obtained results showed that the KOH treatment increased the sorption capacity of hydrochar from 27.8 mg/g up to 137 mg/g at pH 5. Adsorption of lead on either of the materials was achieved through the ion-exchange mechanism, chemisorption and Pb2+-π interaction. The Sips isotherm model gave the best fit with the experimental data obtained for Pb2+ sorption using activated hydrochar. Similarly, Jiang et al. [50] prepared hydrochar by using Sedum Alfredii Hance as raw material under KOH modification at 180–270 °C for 5 h. Research showed that the adsorption capacity of hydrochar for Cd(II) was greatly enhanced after KOH modification, due to improved specific surface areas and pore structure. The maximum Cd(II) adsorption capacity of modified hydrochar was 25.69 mg/g. Additionally, Qian et al. [51] utilized bamboo sawdust as raw material to prepare hydrochar maintained at 473 K for 24 h and treated with a NaOH modification. The adsorption capacity of hydrochar to methylene blue after the modified treatment is 655.76 mg/g. The high adsorption capacity of hydrochar for methylene blue indicates that alkaline treatment followed by hydrothermal carbonization in acidic media has potential applications for producing highly efficient MB adsorbents for wastewater treatment.
Huang et al. [52] utilized sludge as raw material under hydrothermal conditions at 180 °C for 2 h, with a citric acid dosage ratio (citric acid to dry sludge mass ratio) of 0.1. The produced hydrochar featured a maximum surface area of 59.95 m2/g, rich in oxygen-containing functional groups, and exhibited the highest equilibrium adsorption capacity for Pb(II). Wang et al. [53] activated hydrochar produced from sawdust at temperatures between 200 and 250 °C and a heating rate of 4 °C/min using acetic acid, which reduced the energy consumption of the HTC reaction. The resultant hydrochar had a higher calorific value and energy density, making it suitable for biofuel applications. Chen et al. [54] used bamboo powder as raw material and activated the product in citric acid solution under hydrothermal conditions at 180 °C for 6 h. Citric acid is an efficient HTC additive to improve the carbonization degree, to optimize the pore structure distribution, and to introduce more oxygen-containing functional groups (OFGs) to hydrochar, which offer an appropriate precursor to be activated by KOH to improve the specific surface area (SSA) level and electrochemical performance of the activated porous carbon.
Huang et al. [55] utilized raw sewage sludge as raw material and modified hydrochar with calcium acetate (CaAc2) and sodium acetate (NaAc) at 160–250 °C for 30 min. This modification increased the surface area and pore volume compared to unmodified samples and achieved a phenol removal rate of 65.7%. The addition of Ca acetate reduces the nitrogen retention in the hydrochar due to enhanced protein hydrolysis and deamination, leading to more nitrogen transformation into the NH4+-N in the aqueous product. The addition of Na acetate also slightly enhances protein hydrolysis, thus increasing the polypeptide-N in the aqueous product. Mumme et al. [56] conducted HTC of cow dung and corn digestate at 190–270 °C using natural zeolite as a catalyst. Adding zeolite enhances the carbon, hydrogen, and ash content, improves energy recovery rates, increases the surface area and pore volume of the hydrochar, and significantly boosts the recovery rates of nitrogen and sulfur. Lang et al. [57] activated hydrochar from pig manure with CaO at temperatures of 180, 200, and 220 °C for 10 h, and significantly increased the pH and yield of the hydrochar, although the recovery rates of carbon and nitrogen were slightly decreased. Additionally, the addition of CaO facilitates the transformation of phosphorus from non-apatite inorganic phosphorus (NAIP) to apatite phosphorus (AP), with nearly all the phosphorus in the pig manure recovered as AP in the hydrochar through CaO-assisted HTC.

2.2.3. Multi-Step HTC

During the HTC process, raw materials are mixed with a suitable solvent such as water in a sealed container. The reaction conditions are manipulated through multiple stages of varying temperatures and durations to produce hydrochar with unique porous structures and surface characteristics, enhancing its utility across various fields. Specifically, the two-step method involves an initial stage under low temperatures and high pressure to degrade organic matter in biomass into smaller organic molecules. The second stage operates at high temperatures and lower pressure to further carbonize these molecules into carbon materials. This multi-step approach enhances carbonization efficiency, optimizes the structure of the carbon materials, and reduces energy loss.
Zhang et al. [58] synthesized a novel arsenic adsorption material, iron-modified hydrochar, by using different iron species, i.e., FeCl3·6H2O, FeSO4·7H2O, and Fe(NO3)3·9H2O. These hydrochars were prepared through a one-step hydrothermal carbonization process at 220 °C. The results indicated that compared with FeCl3·6H2O and FeSO4·7H2O, hydrochar modified with Fe(NO3)3·9H2O exhibited a maximum iron retention rate of 84.2% and a maximum arsenic adsorption capacity of 11.19 mg/g. Jiao et al. [59] proposed an effective strategy for synergistic production of high value-added xylooligosaccharides (XOS) and humic-like acid (HLA) from vinegar residue based on a two-step hydrothermal pretreatment. During the first-step hydrothermal pretreatment (170 & DEG; C, 50 min), 29.1% of XOS (X-2-X-6) was obtained. The XOS yield was further improved to 36.2% with endoxylanase hydrolysis, thereby increasing the value of (X-2-X-4)/XOS from 0.8 to 1.0. Subsequently, the second-step hydrothermal pretreatment was investigated to produce HLA from the solid residue of the first-step hydrothermal pretreatment. The highest HLA yield was 15.3% in the presence of 0.6 mol/L of KOH at 210 °C for 13 h. In addition, 31.7% of hydrochar by-product was obtained. The mass balance results showed that 1000 g of vinegar residue produced 67.9 g of XOS, 91.6 g of HLA, and 189.5 g of hydrochar. Tan et al. [60] presented a novel approach to synthesize nitrogen-doped porous carbon materials via a three-step fabrication process using citric acid as the carbon source and urea as the nitrogen source. Firstly, hydrochar was synthesized by a microwave-assisted hydrothermal method using citric acid and urea and as the reactants. The hydrochar was then subjected to high-temperature carbonization in an Ar atmosphere followed by KOH activation, resulting in nitrogen-doped porous carbon materials. The as-prepared porous carbon possesses a high BET surface area of 2397 m2/g and an average pore size of 1.8 nm. Such N-rich porous carbon shows outstanding capacitive performance (365 F/g at 0.5 A/g), good rate capacitive behavior, and excellent cycling stability, indicating a great potential for supercapacitors.

2.2.4. HTC Used in the Hydrothermal Process of Liquid Circulation

During the HTC process, the liquid by-products are recycled to enhance the biomass conversion rate and the production efficiency of carbon materials. The recycling of the process liquid significantly improves the resource recovery efficiency of the hydrothermal fluid, maximizing the value of the products and enhancing water resource utilization. The use of hydrothermal fluid as a recyclable medium not only reduces water consumption, but also offers benefits such as the reuse of thermal energy, increasing the efficiency of energy recovery, and reducing the costs of HTC processing [61]. Furthermore, recycling hydrothermal fluid significantly reduces the volume of wastewater produced, decreases the costs of wastewater treatment and contributes to environmental protection [62].
Stemann et al. [62] conducted 19 cycles of process water under hydrothermal conditions at 220 °C for 4 h. During the first five cycles, the total organic carbon (TOC) concentration was increased, and then stabilized. The accumulation of organic acids in the liquid phase enhances the calorific value and carbon concentration of the solids, thereby increases energy density. Uddin et al. [63] used Pinus taeda at hydrothermal temperatures of 200–260 °C with a residence time of 5 min and a water-to-biomass dry basis mass ratio of 5. They recycled the process at 200 and 230 °C for 9 cycles and 260 °C for 5 cycles. The biomass carbon yield was 5–10% higher than the initial cycle at each investigated temperature, with the higher heating value (HHV) remaining essentially unchanged. The TOC concentration in the aqueous phase was gradually concentrated with each cycle, reaching equilibrium.

2.2.5. Co-HTC

Co-HTC involves mixing two or more types of biomasses and processing them together. This method leverages the differences in properties and densities of the materials, resulting in a synergistic effect that significantly enhances combustion. This synergy allows the mixture to burn more completely, releasing greater amounts of heat and making more efficient use of waste biomass materials [47,64,65]. Additionally, co-carbonization can lower the carbonization temperature required for individual materials, thus improving resource utilization. The co-treatment extends the temperature interval between hydrolysis and pyrolysis, reducing the activation energy required for pyrolysis. Furthermore, it significantly improves ignition and combustion characteristics, shifting the peak combustion rate and its temperature range to higher temperatures, thereby lowering the activation energy of the combustion reaction [47]. In Table 2, several research are carried out to better understand co-HTC technology.
Zhu et al. [64] mixed food waste digestate (FWD) with wood waste (WW) as feedstock for co-HTC at 250 °C for 2 h. The results showed that the combination strategy enhanced the fuel properties of the hydrochar. In particular, the ash content of hydrochar was reduced to a minimum of 6.3%, thereby increasing the heating value by nearly two folds. The comprehensive combustion and combustion stability indices were improved with the maximum values of 3.98 × 10−9 %2°C−5 and 4.22 × 102 %°C−3, respectively. Wang et al. [65] used food waste (FW) and woody sawdust (WS) as feedstock for conducting co-HTC at 180–260 °C for 1 h with a stirrer speed of 100 r/min. Results suggested that hydrochar yield consistently decreased with an increase in both the FW ratio and HTC temperature. The C retention from 260 °C hydrochar was low (approximately 65%), but more microsphere structures were formed due to the enhanced carbonization degree of the hydrochar. Li et al. [66] mixed livestock manure (SM) with corn cob (CC) as feedstock for co-HTC at 180–260 °C for 2 h. Compared with HTC of SM, the addition of CC could effectively reduce the ash content, enhance the N recovery to 38.95–47.61% (SM:CC = 1:1), and increase the surface porous structure, making the Co-hydrochars suitable as fertilizers. Under the optimal hydrothermal conditions of 240 °C, 2 h, and mixing ratio of 1:1, the hydrochar yield was as high as 36.72%, and the TNC (6.341%), N recovery rate (47.61%) and P recovery rate (86.41%) were all suitable for fertilizer. Sharma and Dubey [67] mixed food waste (FW) and yard waste (YW) for co-HTC at 220–260 °C for 1–4 h. They found that the calorific value of blended raw feedstock was 13.5 MJ/kg which increased to 27.6 MJ/kg after Co-HTC at 220 °C for 1 h. The energy yield and fuel ratio calculated was 45% and 0.65, respectively. The hydrochar produced demonstrated a stable combustion profile as compared to reactive combustion profile for raw samples. Shen et al. [68] used corn straw and chlorella as raw materials for co-HTC at 240 °C for 1 h. They observed that moderate reaction conditions favored nitrogen enrichment and increased porosity in the hydrochar. Under these conditions, the optimal nitrogen content and surface area of the product were 3.50% and 5.91 m2/g, respectively.

2.2.6. HTC at Different Temperatures

In the HTC process, the temperature significantly impacts the properties and structure of the carbonized products. Variations in hydrothermal temperatures yield diverse products, and studying these differences is crucial for developing high-performance carbon materials for practical applications. Higher temperatures (above 300 °C) facilitate the effective degradation of organic materials in biomass, resulting in high-purity carbon materials with greater potential application value. Conversely, lower temperatures (below 300 °C) are more conducive to preserving carbon. In Table 3, some studies were performed to better understand the effects of different temperatures on HTC technology.
Oxidized hydrothermal biochar was prepared by hydrothermal carbonization of Spartina alterniflora biomass (240 °C for 4 h) and subsequent oxidization (240 °C for 10 min) under air. Oxidized hydrochar achieved a Fe(III) reducing capacity of 2.15 mmol/g at pH 2.0 with 120 h, which is 1.2 times higher than un-oxidized hydrochar. Low temperature oxidization increases the contents of carboxyl and carbonyl groups on the hydrochar surface. This study reveals that low temperature oxidization is an effective way to improve and restore the abiotic reducing ability of hydrochar [69]. The hydrochar yield and carbon retention in processing the livestock manure were decreased when increasing the reaction temperatures between 180 and 240 °C for 1 h [70]. Nguyen et al. [71] performed the reaction with grape marc at 180–260 °C for 30 min to prepare the hydrochar slurries for ignition experiments. It has shown that the 260 °C solid hydrochar exhibited the shortest ignition delay time (0.2 s) and the lowest ignition temperature (179 °C).
Gao et al. [72] used eucalyptus bark as the raw material in the reaction for 2–10 h in the range of 220–300 °C to obtain the hydrochar. With the increase in temperature, the yield of hydrochar decreased slightly from 46.4% at 220 °C to 40.0% at 300 °C. The atomic ratio of O/C and H/C decreased from 1.69 and 0.80 to 0.83 and 0.23, respectively, and the oxygen-containing functional groups decreased with increasing temperature. Peng et al. [73] used glucose as the raw material and kept it at 160–220 °C for 1–12 h to make the hydrochar. The experimental study showed that the adhesion of hydrochar was different: the lower the temperature, the greater the adhesion. Hydrochar is mainly composed of furan domains and aromatic clusters, and its surface is rich in oxygen-containing functional groups. The increase in the reaction temperature (160–220 °C) enhanced the aromatization degree of the hydrochar. Li et al. [74] prepared hydrochar with chicken, dairy, and swine manures at 200–350 °C for 2 h. The results showed that high temperature resulted in low yield and decreased H/C, O/C and volatiles of hydrochar. The high temperature results in a higher fixed carbon, fuel ratio and calorific value of the hydrochar, which indicates that hydrochar from animal manures can be used as a substitute for fuel. The carbon retention rate decreases with the increase in temperature. High temperature improves the aromatics of hydrochar and enhances its thermal oxidation resistance.
In practical applications, the choice of carbonization methods and operational conditions varies based on specific needs and circumstances. For example, in waste management, HTC can transform waste into hydrochar and steam, extracting valuable substances and energy. In the field of biomass energy conversion, this technology can process biomass into high-purity, high energy-dense hydrochar, which can be further utilized in the production of fuels, chemicals, and electrode materials. In Table 4, the advantages of the methods for the different types of HTC technology are described.
In HTC, the conversion of biomass into carbon materials involves several critical steps. In Figure 6, the general steps of HTC technology are described. Initially, biomass constituents such as cellulose, sugars, and lignin are mixed with water to create a slurry. This mixture is then transferred to a sealed reactor or autoclave, where it is heated to a predetermined temperature for a specific duration. Under these high temperature and pressure conditions, the biomass undergoes a series of chemical reactions including hydrolysis, degradation, and polymerization, leading to the formation of carbon materials. Once the reaction is completed, the reactor is cooled to room temperature, and the solid carbonaceous product is removed, washed, and dried. Finally, to enhance the properties of the carbon material, further post-treatment steps such as activation or surface modification may be employed.

3. Application of HTC in Water Environment Management

Water environment management is essential for global sustainable development [29,30,33]. HTC technology represents an emerging environmentally friendly approach for revolutionizing global water environmental management. Primarily, this technology leverages the heat and pressure conditions in aqueous environments to transform organic materials into carbon materials with outstanding properties. As a novel green technology, HTC holds broad prospects for application in water environmental governance. This paper will explore in detail the applications of HTC in managing water environments, including wastewater treatment, surface water restoration, groundwater purification, and other areas of use.

3.1. Wastewater Treatment

HTC products are widely used in wastewater treatment, such as heavy metals removal, organic pollutants, and anionic wastewater. Hydrochar offers several advantages in wastewater treatment, including high efficiency, versatility, low cost, and simple operation, thus realizing significant environmental and economic benefits through the “waste-to-wealth” approach. Due to its high functional group content, porosity, and surface charge, hydrochar can effectively adsorb heavy metals, organic compounds, and anions in wastewater [75]. By modifying hydrothermal conditions and adding modifiers, the quantity and types of oxygen-containing functional groups on the surface of hydrochar can be increased, thereby further optimizing the availability of active sites. In Figure 7, HTC’s liquid and solid processes for treating wastewater are described.

3.1.1. Use of Hydrochar for Heavy Metal Ions

For heavy metal ions, the adsorption mechanisms of hydrochar primarily include electrostatic attraction, ion exchange, complexation, redox reactions, and coprecipitation as chemical adsorption processes, along with physical adsorption [76,77,78]. Redox adsorption can facilitate chemical reactions between various anions and mineral components on the surface of hydrochar and heavy metal ions, resulting in precipitation [77]. In addition to chemical adsorption, physical adsorption is also an important mechanism for hydrochar adsorption of heavy metal ions. Physical adsorption is mainly due to the van der Waals force or hydrophobic force of heavy metal ions with the hydrochar surface. When dealing with heavy metal pollution, appropriate hydrochar materials and optimized preparation conditions can be selected according to the specific application scenarios and the characteristics of the target heavy metal ions to improve the adsorption efficiency. At the same time, the adsorption performance of hydrochar can be further improved by means of surface modification and doping with other functional materials. In Figure 8, the main Ce(III) adsorption mechanism onto hydrochar surface is described. In Table 5, some studies show the effect of hydrochar on the adsorption of heavy metals.
Amine-modified hydrochar synthesized from black liquor using hexamethylenediamine (HMDA) at hydrothermal temperatures of 150, 180, 200, and 220 °C, and a duration of 0.5 h exhibited a maximum adsorption capacity of Cr(VI) removal of 741.74 mg/g at 45 °C and pH 2 [80]. Hydrochar microspheres produced from glucose at 180 °C for 48 h enables highly selective separation of U(VI). At 333.15 K and pH 4.5, the maximum adsorption capacity for U(VI) was 408.36 mg [81]. Similarly, hydrochar microspheres synthesized from sucrose at 190 °C for 11 h, activated with KOH demonstrated a maximum adsorption capacity of 704.2 mg/g for methylene blue. The adsorption process was endothermic [82]. Phosphoric acid-modified hydrochar produced from banana peels to at 230 °C over 2 h was evaluated for adsorption capabilities of Pb2+ [83]. The evaluation revealed lead adsorption capacities of 359 mg/g and 193 mg/g for dehydrated and fresh banana peel hydrochar, respectively. Moreover, hydrochar produced from willow twigs via HTC at 300 °C for 30 min used as an adsorbent to remove Cu2+ and Cd2+ from aqueous solutions exhibited adsorption capacities of 34 mg/g (0.313 mmol/g) for Cd2+ and 31 mg/g (0.503 mmol/g) for Cu2+ [84]. Furthermore, hydrochar from pine sawdust (SD) at 260 °C for 2 h, was then activated with H2O2 to enhance the Pb2+ removal capacity [85]. The maximum adsorption capacity of Pb2+ reached 92.80 mg/g at pH 5.0 and 298 K, which is more than 42 times higher than that of the original hydrochar (2.20 mg/g). Hydrochar from synthesized Pseudomonas aeruginosa shells at 200 °C for 20 min was found to be more effective in adsorbing Pb2+ and Cd2+ compared to biochar. The adsorption kinetics followed a pseudo-second-order model, describable by the Langmuir isotherm [86]. After reaction at 180–220 °C for 9 h, citric acid catalyzed arecanut husk to produce hydrochar. The results showed that compared with the parent biomass, the mass yield of hydrochar was 58.7%, and the fixed carbon increased from 17% to 39.7%, compared with Zn2+, Cr6+ and Ni2+, in the aqueous solution with the initial concentration of 100 mg/L. Hydrochar at a dose of 0.1% shows a maximum adsorption rate of Pb2+ (79.86 mg/g), and hydrochar can be used to remove Pb2+ from wastewater as it shows a maximum removal efficiency of 95.08% at 25 mg/L [87]. Additionally, sulfide-modified magnetic hydrochar (MHC-S4) was prepared by simultaneously supporting the synthesis of nanoparticles with Fe3O4 and grafting a sulfide-containing group onto a carboniferous water prepared by reacting with pinecone at 200 °C for 5 h. MHC-S4 can effectively adsorb Cd (II)/Pb (II) over a wide pH range and achieve rapid absorption equilibrium within 25 min. In addition, MHC-S4 has excellent adsorption properties in a single system, and the maximum single-layer adsorption capacity of Cd (II) and Pb (II) is 62.49 and 149.33 mg/g, respectively [88].

3.1.2. Use of Hydrochar for Organic Pollutants

Hydrochar serves as an effective adsorbent for organic pollutants such as fuels and antibiotics in wastewater. The adsorption mechanisms of hydrochar involve both physical and chemical interactions, including electrostatic interactions, hydrophobic effects, hydrogen bonding, pore filling, partitioning, and interactions between aromatics and cations [76,77,89]. Physical adsorption utilizes the surface characteristics of hydrochar, such as porosity and surface area, to adsorb pollutants on its surface or within its micropores [90]. Chemical adsorption involves the formation of chemical bonds between the adsorbent’s functional groups, such as hydroxyl, carboxyl, and amino groups, and the adsorbates, facilitating effective removal of heavy metal ions through stable chelation or ionic and covalent bonds. The dominant mechanism depends on the structural and physicochemical properties of the organic pollutants and the biomass charcoal. Modifications and preparations of hydrochar alter its physicochemical properties, enhancing its adsorption performance for organic compounds [91]. In Figure 9, the adsorption mechanism of hydrochar for organic pollutants was emphasized by describing the mechanism of TC adsorption of hydrochar, modified-hydrochar, and hydrochar-derived ACs. In Table 6, some studies show the effect of hydrochar on the adsorption of organic pollutants.
Hydrochar from synthesized municipal sludge at 600 °C for 2 h with Zn/Fe modification demonstrated maximum adsorption capacities of 145.0 mg/g for tetracycline (TC) and 74.2 mg/g for ciprofloxacin (CIP) [93]. Moreover, hydrochar from produced bamboo shavings at 200 °C for 3 h, assisted by acid treatment, exhibited maximum adsorption capacities of 90.51 mg/g for Congo Red and 72.93 mg/g for 2-naphthol [94]. Tian et al. [91] evaluated the adsorption effects of KOH-modified and unmodified hydrochar produced from montmorillonite and rice husks at 180 °C for 16 h on estrogens in water. They found that a 1% mass ratio of KOH to hydrochar maximized the adsorption efficiency. Li et al. [95] prepared hydrochar from rice straw at temperatures ranging from 160 °C to 200 °C and durations from 40 to 70 min. At 298 K and an initial concentration of 0.5 mg/mL, the maximum adsorption capacities for Congo Red, berberine hydrochloride, 2-naphthol, Zn2+, and Cu2+ were 222.1 mg/g, 174.0 mg/g, 48.7 mg/g, 112.8 mg/g, and 144.9 mg/g, respectively. Furthermore, hydrochar produced from coffee husks at 210 °C for 243 h achieves an adsorption capacity of 34.85 mg/g for methylene blue, predominantly through physical adsorption [96]. Zbair et al. [97] synthesized hydrochar from Moroccan nut shells (ANS) at temperatures of 180 °C and 200 °C for 6 h. Additionally, hydrochar produced from fructose at 180 °C for 2 h modified with phloroglucinol showed an adsorption capacity of 274.7 mg/g for tetracycline in water [98].

3.1.3. Use of Hydrochar for Inorganic Anions

For inorganic anions, hydrochar typically undergoes chemical adsorption through chemically formed bonds on the surface [77], exhibiting good adsorption efficiency for anions such as phosphate, arsenate, and fluoride in wastewater. For example, the hydroxyl group on the surface of the hydrochar can form hydrogen bonds with the phosphate ions or undergo coordination interactions to remove the phosphate from the wastewater. Similarly, the carboxyl group on the surface can coordinate with the metal ions in arsenic acid to form a stable complex to effectively remove arsenic acid. For fluoride ions, metal oxides such as aluminum and iron on the surface of hydrochar can exchange ions with fluoride ions to reduce the concentration of fluoride ions in water. In addition, good hydrochar chemical stability and corrosion resistance can be in a wider range of pH to maintain good adsorption performance, which makes the hydrochar become the ideal material for dealing with the wastewater containing inorganic anions. In Table 7, some studies show the effect of hydrochar on the adsorption of inorganic anions.
Anoxygenically digested cattail hydrochar was produced by reacting with pure water, acetic acid assisted, and sodium hydroxide assisted HTC at 200–300 °C for 1 h. The study showed that when the HTC temperature increased from 200 °C to 300 °C, the yield using water, acetic acid and sodium hydroxide decreased from 40.2% to 31.6%, 37.5% to 28.3% and 35.7% to 22.7%, respectively. The adsorption capacities of NH4+-N and PO43−-P of these hydrochar are 92.6–122.4 mg/g and 1.6–15.8 mg/g, respectively, which fit the Freundlich model well [99]. Hydrothermal modification reduces the number of acidic functional groups, which is beneficial for anion adsorption. Hydrochar synthesized from chicken feathers at 150–170 °C after 1–3 h has a strong affinity for phosphate, and has a good adsorption efficiency for phosphorus removal under acidic conditions [100]. The adsorption isotherms for hydrochar suited the Langmuir model better, with the maximum adsorption capacity (qm) of hydrochar being 21.70 mg/g at 30 °C. Hydrochar synthesized from microalgae at 348 K for 40 min modified with magnesium exhibited a strong affinity for phosphate, with a maximum adsorption capacity of 89.61 mg/g [25]. Additionally, the magnesium-containing hydrochar adsorbed phosphate from water mainly through ion exchange. Magnetic hydrochar modified with Fe3O4 was synthesized from sodium alginate at 210 °C for 5 h [101]. The results showed that at a dosage of 2 g/L, the removal rates of arsenic and fluoride were around 85%, with maximum adsorption capacities of 26.06 mg/g and 15.64 mg/g, respectively, in individual systems.
Anoxygenically digested cattail hydrochar was produced by reacting with pure water, acetic acid assisted, and sodium hydroxide assisted HTC at 200–300 °C for 1 h. The study showed that when the HTC temperature increased from 200 °C to 300 °C, the yield using water, acetic acid and sodium hydroxide decreased from 40.2% to 31.6%, 37.5% to 28.3% and 35.7% to 22.7%, respectively. The adsorption capacities of NH4+-N and PO43−-P of these hydrochar are 92.6–122.4 mg/g and 1.6–15.8 mg/g, respectively, which fit the Freundlich model well [99]. Hydrothermal modification reduces the number of acidic functional groups, which is beneficial for anion adsorption. Hydrochar synthesized from chicken feathers at 150–170 °C after 1–3 h has a strong affinity for phosphate, and has a good adsorption efficiency for phosphorus removal under acidic conditions [100]. The adsorption isotherms for hydrochar suited the Langmuir model better, with the maximum adsorption capacity (qm) of hydrochar being 21.70 mg/g at 30 °C. Hydrochar synthesized from microalgae at 348 K for 40 min modified with magnesium exhibited a strong affinity for phosphate, with a maximum adsorption capacity of 89.61 mg/g [25]. Additionally, the magnesium-containing hydrochar adsorbed phosphate from water mainly through ion exchange. Magnetic hydrochar modified with Fe3O4 was synthesized from sodium alginate at 210 °C for 5 h [101]. The results showed that at a dosage of 2 g/L, the removal rates of arsenic and fluoride were around 85%, with maximum adsorption capacities of 26.06 mg/g and 15.64 mg/g, respectively, in individual systems.

3.2. Surface Water Restoration

Surface water restoration is a primary focus in current water environment management [102], with HTC technology playing a significant role, particularly in sediment improvement and water quality enhancement in rivers and lakes. Among them, surface water restoration refers to the treatment and restoration of polluted or damaged surface water bodies through a series of hydrothermal carbonization technical measures and management measures to achieve the purpose of improving water quality and restoring water ecosystem functions. By thermally decomposing and carbonizing sediments and deposits, this technology effectively reduces nutrient release from sediments, thereby controlling eutrophication trends in lakes. Additionally, the hydrochar produced exhibits high adsorptive properties, enabling the selective absorption of harmful substances in the water, significantly enhancing water quality and purifying the water body. Following treatment with HTC technology, rivers and lakes gradually restore their natural ecosystems, providing favorable habitats for diverse aquatic life forms and yielding substantial economic and environmental benefits [103]. In Table 8, studies through surface water restoration highlight the role of hydrochar.
Using Prunus serrulata bark as raw material, hydrochar was prepared by a reaction at 200 °C for 6 h, and the product has high efficiency and low precursor material cost, which can remove atrazine from river water. When the acidic pH = 3 is favorable for adsorption, the ideal adsorbent dose is 0.8 g/L, and kinetic studies show that the concentration does not affect the system equilibrium time, which is reached at 240 min. The Langmuir model presented the greatest compliance to the isotherm data and indicated a higher affinity between atrazine and hydrochar, reaching a maximum adsorption capacity of 63.35 mg/g [104]. Additionally, extending the reaction time proved to be beneficial for improving the quality of the hydrothermal biochar. Water hyacinth was employed as a feedstock to produce hydrochar at 240 °C and various durations from 0.5 to 24.0 h [105]. The results revealed that the specific surface area of the hydrochar was initially increased and then decreased over time. This process not only utilizes the invasive water hyacinth but also achieves ecological restoration. Hydrochar prepared from glucose at 180 °C for 10 h and composited with FeCl3·6H2O and MnCl2·4H2O to successfully load MnFe2O4 on its surface could rupture algal cells and disrupt their photosynthetic systems, demonstrating its capability to treat organic matter [106]. Water hyacinth was modified with MgCl2·6H2O and AlCl3·6H2O, and reacted at 150℃ for 24 h to obtain Mg/Al-layered double hydroxides modified water hyacinth hydrochar (MgAl@WH) with layered surface and many adsorption active sites. The maximum adsorption capacity is 311.0 mg/g, and the adsorption efficiency is about 98.0% after 10 cycles [107]. Moreover, research on algal raw materials offers a potential solution for managing eutrophication in lakes and rivers.
Simultaneously, an appropriate amount of hydrochar can enhance the water and soil environment, thereby promoting plant growth and facilitating ecological restoration. Numerous studies have demonstrated that the carbon and magnesium content in hydrochar can influence the uptake of critical nutrients, such as ammonium salts and phosphates [108]. Additionally, hydrochar can enhance sulfur levels in the soil and decrease the available nitrogen, and thereby enhance soybean growth [109]. However, excessive application of hydrochar may inhibit crop growth, particularly during the seed germination and seedling stages [109,110,111].

3.3. Groundwater Purification

Groundwater, as a crucial water resource, is the focus of intense protection and management efforts. Using HTC technology in groundwater purification aims to remove heavy metals and organic pollutants, reduce their bioavailability, and thus inhibit microbial growth and reproduction, ensuring groundwater purity and safety. HTC technology is utilized to prepare hydrochar with diverse adsorption capabilities for various pollutants. Employed as an adsorbent with specific particle sizes, it facilitates efficient underground transport. The impact of carbon-based materials on pollutant removal depends on their physicochemical properties, which are determined by the type of raw material and the pyrolysis conditions used for production. The increase in additives can change the physicochemical properties of carbon-based materials, including biochar and hydrochar, to remove different types of pollutants, which is helpful to develop remediation technologies for contaminated groundwater [112]. This enables the removal of organic substances and heavy metal ions from groundwater, safeguarding water source safety and promoting sustainable groundwater use. In Table 8, studies through groundwater purification highlight the role of hydrochar. Hydrochar prepared from switchgrass at 300 °C for 30 min was utilized to remove U (VI) contaminants from groundwater [113]. The removal of pathogenic rotavirus (RV) and human adenovirus (HAdV) was investigated by hydrothermal treatment of swine feces (2 h at 180 °C and 7 h at 230 °C) under two conditions, using a 10 cm bed height sand column and with or without an aqueous charcoal supplement (1.5%, w/w). The results showed that the removal efficiency of the two viruses in the improved hydration column was >3 logarithm (complete removal). Although different HTC conditions lead to different characteristics of 180 HTC and 230 HTC, there is no significant difference in virus removal performance between the two hydrates. Hydrochar derived from fecal waste can be used as a capacity grade virus adsorbent [114]. Prepared by anaerobic digestion of swine manure at 180 °C for 45 min, hydrochar has been shown to stimulate the growth of denitrifies without toluene degradability, including Candidatus Competibacter and Ferrovibrio, and increase NO3-N removal, which was mainly attributed to the partial denitrification [112]. Following ball milling treatment, it exhibited excellent underground transport properties, making it effective for removing TCE from groundwater.
Table 8. Research on the ecological protection of hydrochar.
Table 8. Research on the ecological protection of hydrochar.
Raw MaterialsHydrothermal ConditionsModification TreatmentProductType of ProtectionReferences
Prunus serrulata bark200 °C, and 6 h——The maximum adsorption capacity for atrazine in the river is 63.35 mg/g.surface water restoration[104]
Water hyacinth240 °C, and 0.5–24.0 h——With the increase in residence time, the higher heating value in all hydrochar products was 16.83 MJ/kg to 20.63 MJ/kg.surface water restoration[105]
Glucose180 °C, 10 hFeCl3·6H2O, MnCl2·4H2O The product can break algal cells and destroy the photosynthetic system of algal cells, and deal with organic substances.surface water restoration[106]
Water hyacinth150 °C, 24 hMg/Al-layered double hydroxide modificationThe maximum adsorption capacity of the product for mordant brown (anionic dye) was 311.0 mg/g.surface water restoration[107]
Switchgrass300 °C, 30 min——The hydrochar-formed permeable reaction wall demonstrated rapid removal of U (VI).groundwater purification[113]
Swine feces180 °C, 2 h; 230 °C, 7 hH2SO4The removal efficiency of pathogenic RV and HAdV in the hydrated modified column was >3 logarithms (complete removal).groundwater purification[114]
Anaerobic digestion of swine manure180 °C, 45 min——Hydrochar enhance toluene removal from groundwater, stimulate the growth of denitrifiers without toluene degradability, and increase NO3-N removal.groundwater purification[112]

4. Research Advances in the Context of the “Dual Carbon” Goals

4.1. Influence of HTC Technology on Carbon Emission

Currently, HTC technology is a research hotspot under the “dual carbon” goals. This technology utilizes high-temperature and high-pressure hydrothermal conditions to transform organic matter into carbonaceous materials [21,22,23,24]. Compared to traditional carbonization methods, HTC offers advantages such as lower energy consumption, minimal pollution, and enhanced performance of carbon materials. More importantly, HTC can convert carbon in organic wastes into stable carbonaceous materials and steam, significantly reduce carbon emissions from the transformation of organic matter into carbon materials and avoid greenhouse gas emissions. This process effectively sequesters carbon from biomass, which is crucial for achieving the “dual carbon” goals. Additionally, HTC technology can recycle and utilize the energy from organic wastes, further enhancing energy efficiency.

4.2. Comparison with Other Environmental Governance Technologies

HTC technology exhibits unique advantages in environmental management. Compared to traditional environmental management technologies, it is more eco-friendly and can significantly reduce secondary pollutants during the treatment of organic wastes, which is fully aligned with current green and low-carbon development concepts [25,26,27]. Additionally, this technology facilitates resource utilization by converting organic wastes into high-value products such as activated carbon and carbon black [32]. This not only enhances the economic value of wastes but also yields significant social benefits. Notably, HTC processes organic wastes under high temperatures and pressures, achieving high processing efficiency and energy recovery rates, which underscores its effectiveness in environmental management. These advantages underscore the broad potential of HTC technology in environmental management.
Traditional sewage treatment technologies, such as the activated sludge method and biofilm method, can effectively remove organic matter and nutrients in wastewater, but the treatment process may produce secondary pollution, such as odor, noise and chemical residue. The HTC technology can significantly reduce these secondary pollutants, which is fully in line with the current concept of green and low-carbon development. Anaerobic fermentation technology is mainly used to treat high concentration of organic wastewater, and the biogas produced is a kind of clean energy. However, for some wastewater with low concentrations and a complex composition, the treatment effect of anaerobic fermentation technology may not be ideal. HTC technology can deal with organic waste of various components to achieve high processing efficiency and energy recovery. In Table 9, different water environmental treatment technologies are compared in terms of advantages, Shortcoming and energy consumption.
However, the HTC technique also has some limitations. High-temperature and high-pressure conditions of HTC technology may increase equipment costs compared to conventional wastewater treatment technologies, and the procurement and processing of biomass or carbon-containing waste may pose environmental challenges. In order to fully exploit the advantages of HTC technology and mitigate its weaknesses, further research and practice are needed. This includes research on technology optimization, reduced equipment costs, and a sustainable supply of biomass or carbon-containing waste.

4.3. Experiments with HTC Technology

Using the Nanjing Qiaobei Sewage Treatment Plant as an example, Wang et al. [115] applied HTC technology to sludge treatment, combined with physicochemical pretreatment, namely Fenton oxidation (FO) and HTC technology to utilize SSL, and found that moderate oxidation could improve the output and performance of the produced hydrochar. The results showed that the pretreatment affected the surface structure and organic composition of SSL and promoted the carbonization of intermediate products. Moreover, compared with the yield of hydrochar (50.7%) obtained via the direct HTC treatment, the yield of the hydrochar obtained using the combined process increased to 55.0–65.2% (depending on the pretreatment condition). Hydrochar properties were enhanced using the combined process, and the energy density of hydrochar slightly decreased after pretreatment (1.2–13.1%); however, the energy yields increased by 0.6–30.1% due to the enhanced hydrochar yield. The carbonization degree of hydrochar was improved; the carbon in the feedstock distributed to the hydrochar increased from 33.40% to 46.09%. The treated sludge not only was effectively reduced but also enabled the recovery and reuse of energy and materials, thereby enhancing the waste’s resource utilization. Furthermore, it adsorbs heavy metal ions in wastewater, contributing to water environmental protection. The application of this technology not only reduces carbon emissions but also effectively addresses challenges in sludge treatment, achieving resource recycling. This provides a valuable reference for urban sludge disposal and resource utilization in China.
During the research process, several successful experiments have provided valuable insights. For example, Shi et al. [116] utilized HTC technology to process large particles (>2 cm) of kitchen waste, successfully converting them into carbon materials. This not only addressed wastes disposal issues but also created a new supply source for the carbon materials market. This experiment demonstrates that technological innovation and application can facilitate a win-win situation for economic development and environmental protection.
Experimental studies serve as an effective means to verify the application effects of HTC technology. Scholars have explored the application of HTC technology in various domains, including agricultural wastes, urban trash, and industrial wastewater [117]. These experimental studies demonstrate that while HTC technology is effective in practical applications, it also faces several challenges that need to be addressed. Challenges include the pretreatment and post-treatment of biomass or carbon-containing wastes and scaling up and industrializing the equipment.

5. Challenges and Countermeasures

5.1. Technical Challenges

Although the principle of HTC technology is relatively simple, its practical application requires precise control of parameters such as temperature, pressure, and time to achieve optimal carbonization effects [26,118,119]. Additionally, the adaptability of this technology to various raw materials poses a technical challenge, as organic wastes of different origins and properties may require tailored processing conditions. Therefore, further optimizing the technological process and enhancing carbonization efficiency are critical challenges that need to be addressed [118,119]. Currently, this technology remains in the research and development stage and requires further optimization and enhancements to achieve more efficient and stable production. Furthermore, the widespread adoption and application of this technology must overcome several obstacles, including challenges related to equipment design, manufacturing, operation, and maintenance.

5.2. Economic Costs

Economic cost is a significant challenge facing the application of innovative technology [120,121,122]. While the long-term economic benefits of HTC technology are substantial, the initial investment is considerable, encompassing equipment purchases, installation, commissioning, and personnel training. Additionally, operational energy consumption and maintenance costs are non-negligible. These factors make it prohibitive for economically disadvantaged areas, potentially limiting its adoption in certain regions. Therefore, it is essential to explore strategies to reduce production costs and enhance economic efficiency, enabling broader adoption of this technology by more regions and enterprises.

5.3. Environmental and Social Impact

Environmental and social impact will be closely related to innovative technology application [123,124]. When used to process organic wastes, HTC technology can reduce greenhouse gas emissions and enhance resource utilization efficiency. However, it may also generate by-products and pollutants, such as carbonization residues and wastewater. Improper management of these can lead to secondary environmental pollution. Therefore, in promoting this technology, it is crucial to fully assess its environmental and social impacts and establish comprehensive mechanisms for managing by-products and pollutants to ensure real societal benefits and welfare.

5.4. Policy and Management

The policy and management play a crucial role in the application of innovative technology [125,126]. Governments should establish policies and regulations to encourage and support the research, development, and application of this technology. Additionally, it is essential to establish regulatory mechanisms, define technical standards, and set regulatory requirements to ensure compliance with national and international standards and prevent potential environmental and social risks, thereby providing a robust legal framework for the technology’s application.

6. Conclusions and Perspective

6.1. Summary of the Main Research Results

In the research and application of HTC technology, we have achieved remarkable results. Compared with other wastewater treatment techniques, this technique has been successfully applied to various wastewater treatments, showing effective effects and stability. Moreover, HTC technology has obvious advantages in improving the efficiency of resource recovery extracted from wastewater. It not only transforms organic substances into valuable materials but also can produce high-value carbon materials suitable for different fields, realizing the recycling and reuse of resources.
The application of HTC technology in water environment management provides a new idea and method for water environment management in the world. HTC technology can not only effectively deal with water environment pollution and improve the quality of water environment, but also has obvious advantages in improving the efficiency of resource recovery, reducing the pressure of the resource shortage in China. In addition, the application of HTC technology can also promote the development of the environmental protection industry and promote the realization of a green economy.

6.2. Suggestions for Future Research Directions

Although HTC technology has achieved remarkable results in water environment management, there are still some aspects that need further research and improvement. In my perspective, the study of the HTC reaction mechanism is crucial, which helps to improve the reaction efficiency and optimize the process parameters, so as to ensure the product quality. Therefore, it is necessary to organize more in-depth research for a comprehensive understanding and grasp of the HTC reaction mechanism. In addition, the application scope of HTC technology also needs to be further expanded to adapt to various water quality and flow conditions. To this end, the HTC technology is needed to be optimized and improved to make it suitable for more kinds of wastewater treatment, including industrial wastewater, urban sewage, etc. Finally, enhanced integration with other wastewater treatment technologies is the key to developing more efficient and environmentally friendly wastewater treatment systems. This can not only improve processing efficiency, reduce operating costs, but also better cope with complex water quality problems.
To effectively promote and apply HTC technology, decision makers should provide necessary support and guidance. On the one hand, decision makers need to increase their investment in HTC technology research and development, and encourage the cooperation between enterprises, universities and research institutions, to jointly promote the development of HTC technology. This can not only accelerate the process of technology research and development, but also promote technological innovation and upgrading. On the other hand, decision makers need to develop environmental policies to encourage enterprises to adopt environmentally friendly wastewater treatment technologies and improve wastewater treatment standards. This can not only reduce environmental pollution, but also promote sustainable economic development. Finally, the establishment of a set of comprehensive standards for the sewage treatment industry is crucial to standardize its development and promote the wide application of HTC technologies. Through these measures, it can effectively promote the development and application of HTC technology and provide strong technical support for solving the water environment problems in China.
In general, the research and application of HTC technology in organic wastewater treatment and resource recovery provides a new solution for water environment management in China and has broad application prospects and important research value. In the future, we should continue to study HTC technology, optimize its treatment effect, reduce its operating cost, so as to better cope with the challenges of water environment management, promote the development of China’s water environment protection cause, and make contributions to the construction of a beautiful China.

Author Contributions

G.L.: investigation, data curation, writing—original draft; Q.X.: writing—review and editing, visualization; S.F.A.-E.: writing—review and editing; M.A.A.: writing—review and editing; T.Z.: writing—review and editing, supervision, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

The research was sustained by the grant from the Undergraduate Research Program of China Agricultural University, the National Key Research and Development Program of China “Intergovernmental Cooperation in International Science and Technology Innovation” [Grant number 2023YFE0104700], the National Natural Science Foundation of China [Grant Number 31401944].

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Zhu, Z.; Zhang, X.; Dong, H.; Wang, S.; Reis, S.; Li, Y.; Gu, B. Integrated livestock sector nitrogen pollution abatement measures could generate net benefits for human and ecosystem health in China. Nat. Food 2022, 3, 161–168. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, Y.; Su, M.; Wang, Y.; Cui, S.; Meng, F.; Yue, W.; Liu, Y.; Xu, C.; Yang, Z. Food production in China requires intensified measures to be consistent with national and provincial environmental boundaries. Nat. Food 2020, 1, 572–582. [Google Scholar] [CrossRef]
  3. Springmann, M.; Clark, M.; Croz, D.M.D.; Wiebe, K.; Bodirsky, B.L.; Lassaletta, L.; de Vries, W.; Vermeulen, S.J.; Herrero, M.; Carlson, K.M.; et al. Options for keeping the food system within environmental limits. Nature 2018, 562, 519–525. [Google Scholar] [CrossRef] [PubMed]
  4. Raihan, A.; Voumik, L.C.; Mohajan, B.; Rahman, M.S.; Zaman, M.R. Economy-energy-environment nexus: The potential of agricultural value-added toward achieving China’s dream of carbon neutrality. Carbon Res. 2023, 2, 43. [Google Scholar] [CrossRef]
  5. Qiu, L.; Yu, R.; Hu, F.; Zhou, H.; Hu, H. How can China’s medical manufacturing listed firms improve their technological innovation efficiency? An analysis based on a three-stage DEA model and corporate governance configurations. Technol. Forecast. Soc. Chang. 2023, 194, 122684. [Google Scholar] [CrossRef]
  6. Jin, S.; Zhang, B.; Wu, B.; Han, D.; Hu, Y.; Ren, C.; Zhang, C.; Wei, X.; Wu, Y.; Mol, A.P.J.; et al. Decoupling livestock and crop production at the household level in China. Nat. Sustain. 2021, 4, 48–55. [Google Scholar] [CrossRef]
  7. Li, F.; Cheng, S.; Yu, H.; Yang, D. Waste from livestock and poultry breeding and its potential assessment of biogas energy in rural China. J. Clean. Prod. 2016, 126, 451–460. [Google Scholar] [CrossRef]
  8. Jing, Q.; Luo, W.; Bai, H.; Xu, H. A method for city-level energy-related CO2 emission estimation. J. Environ. Sci. 2018, 38, 4879–4886. [Google Scholar]
  9. Wang, Y.; Quan, S.; Tang, X.; Hosono, T.; Hao, Y.; Tian, J.; Pang, Z. Organic and inorganic carbon sinks reduce long-term deep carbon emissions in the continental collision margin of the southern Tibetan plateau: Implications for Cenozoic climate cooling. J. Geophys. Res.-Solid Earth 2024, 129, e2024JB028802. [Google Scholar] [CrossRef]
  10. Zhao, S.; Zhang, L.; An, H.; Peng, L.; Zhou, H.; Hu, F. Has China’s low-carbon strategy pushed forward the digital transformation of manufacturing enterprises? Evidence from the low-carbon city pilot policy. Environ. Impact Assess. Rev. 2023, 102, 107184. [Google Scholar] [CrossRef]
  11. Zhao, S.; Zhang, L.; Peng, L.; Zhou, H.; Hu, F. Enterprise pollution reduction through digital transformation? Evidence from Chinese manufacturing enterprises. Technol. Soc. 2024, 77, 102520. [Google Scholar] [CrossRef]
  12. Li, H.; Zhang, T.; Shaheen, S.M.; Abdelrahman, H.; Ali, E.F.; Bolan, N.S.; Li, G.; Rinklebe, J. Microbial inoculants and struvite improved organic matter humification and stabilized phosphorus during swine manure composting: Multivariate and multiscale investigations. Bioresour. Technol. 2022, 351, 126976. [Google Scholar] [CrossRef] [PubMed]
  13. Shang, M.; Luo, J. The Tapio decoupling principle and key strategies for changing factors of Chinese urban carbon footprint based on cloud computing. Int. J. Environ. Res. Public Health 2021, 18, 2101. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, C.; Chen, J.; He, B.; Liu, J. Digitalization and carbon emission reduction technology R&D in a Stackelberg model. Appl. Econ. Lett. 2024, 1–6. [Google Scholar] [CrossRef]
  15. Wu, F.; Li, F.; Zhao, X.; Bolan, N.S.; Fu, P.; Lam, S.S.; Mašek, O.; Ong, H.C.; Pan, B.; Qiu, X.; et al. Meet the challenges in the “carbon age”. Carbon Res. 2022, 1, 1. [Google Scholar] [CrossRef]
  16. Zhang, T.; Wu, X.; Shaheen, S.M.; Abdelrahman, H.; Ali, E.F.; Bolan, N.S.; Ok, Y.S.; Li, G.; Tsang, D.C.W.; Rinklebe, J. Improving the humification and phosphorus flow during swine manure composting: A trial for enhancing the beneficial applications of hazardous biowastes. J. Hazard. Mater. 2022, 425, 127906. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, T.; Li, H.; Yan, T.; Shaheen, S.M.; Niu, Y.; Zhang, Y.; Abdelrahman, H.; Ali, E.F.; Bolan, N.S.; Rinklebe, J. Organic matter stabilization and phosphorus activation during vegetable waste composting: Multivariate and multiscale investigation. Sci. Total Environ. 2023, 891, 164608. [Google Scholar] [CrossRef] [PubMed]
  18. Fang, C.; Zhang, T.; Li, P.; Jiang, R.; Wu, S.; Nie, H.; Wang, Y. Phosphorus recovery from biogas fermentation liquid by Ca-Mg loaded biochar. J. Environ. Sci. 2015, 29, 106–114. [Google Scholar] [CrossRef]
  19. He, X.; Zhang, T.; Ren, H.; Li, G.; Ding, L.; Pawlowski, L. Phosphorus recovery from biogas slurry by ultrasound/H2O2 digestion coupled with HFO/biochar adsorption process. Waste Manag. 2017, 60, 219–229. [Google Scholar] [CrossRef]
  20. Zhang, T.; Xu, H.; Li, H.; He, X.; Shi, Y.; Kruse, A. Microwave digestion-assisted HFO/biochar adsorption to recover phosphorus from swine manure. Sci. Total Environ. 2018, 621, 1512–1526. [Google Scholar] [CrossRef]
  21. He, X.; Zhang, T.; Xue, Q.; Zhou, Y.; Wang, H.; Bolan, N.S.; Jiang, R.; Tsang, D.C.W. Enhanced adsorption of Cu(II) and Zn(II) from aqueous solution by polyethyleneimine modified straw hydrochar. Sci. Total Environ. 2021, 778, 146116. [Google Scholar] [CrossRef]
  22. Zhang, T.; Wu, X.; Shaheen, S.M.; Zhao, Q.; Liu, X.; Rinklebe, J.; Ren, H. Ammonium nitrogen recovery from digestate by hydrothermal pretreatment followed by activated hydrochar sorption. Chem. Eng. J. 2020, 379, 122254. [Google Scholar] [CrossRef]
  23. Huang, R.; Fang, C.; Lu, X.; Jiang, R.; Tang, Y. Transformation of phosphorus during (hydro)thermal treatments of solid biowastes: Reaction mechanisms and implications for P reclamation and recycling. Environ. Sci. Technol. 2017, 51, 10284–10298. [Google Scholar] [CrossRef] [PubMed]
  24. Shi, Y.; Zhang, T.; Ren, H.; Kruse, A.; Cui, R. Polyethylene imine modified hydrochar adsorption for chromium (VI) and nickel (II) removal from aqueous solution. Bioresour. Technol. 2018, 247, 370–379. [Google Scholar] [CrossRef]
  25. Deng, Y.; Zhang, T.; Sharma, B.K.; Nie, H. Optimization and mechanism studies on cell disruption and phosphorus recovery from microalgae with magnesium modified hydrochar in assisted hydrothermal system. Sci. Total Environ. 2019, 646, 1140–1154. [Google Scholar] [CrossRef] [PubMed]
  26. Deng, Y.; Zhang, T.; Clark, J.; Aminabhavi, T.; Kruse, A.; Tsang, D.C.W.; Sharma, B.K.; Zhang, F.; Ren, H. Mechanisms and modelling of phosphorus solid–liquid transformation during the hydrothermal processing of swine manure. Green Chem. 2020, 22, 5628–5638. [Google Scholar] [CrossRef]
  27. Wang, T.; Zhai, Y.; Zhu, Y.; Li, C.; Zeng, G. A review of the hydrothermal carbonization of biomass waste for hydrochar formation: Process conditions, fundamentals, and physicochemical properties. Renew. Sust. Energy Rev. 2018, 90, 223–247. [Google Scholar] [CrossRef]
  28. Zhang, T.; Ding, L.; Ren, H.; Guo, Z.; Tan, J. Thermodynamic modeling of ferric phosphate precipitation for phosphorus removal and recovery from wastewater. J. Hazard. Mater. 2010, 176, 444–450. [Google Scholar] [CrossRef]
  29. Zhang, T.; Li, Q.; Ding, L.; Ren, H.; Xu, K.; Wu, Y.; Sheng, D. Modeling assessment for ammonium nitrogen recovery from wastewater by chemical precipitation. J. Environ. Sci. 2011, 23, 881–890. [Google Scholar] [CrossRef]
  30. Fang, C.; Zhang, T.; Jiang, R.; Ohtake, H. Phosphate enhance recovery from wastewater by mechanism analysis and optimization of struvite settleability in fluidized bed reactor. Sci. Rep. 2016, 6, 32215. [Google Scholar] [CrossRef]
  31. Xu, X.; Feng, F.; Wan, Z.; Wang, Y.; Yu, M.; Han, X.; Wu, G.; Xing, W. Rapid electron transfer reinforced by interfacial Co-O bonding in MOF/COF hybrids for highly efficient degrade tetracycline by activating peroxymonosulfate. Colloid Surf. A-Physicochem. Eng. Asp. 2024, 689, 133686. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Zhang, T. Biowaste valorization to produce advance carbon materials-hydrochar for potential application of Cr (VI) and Cd (II) adsorption in wastewater: A review. Water 2022, 14, 3675. [Google Scholar] [CrossRef]
  33. Zhang, T.; Li, P.; Fang, C.; Jiang, R. Phosphate recovery from animal manure wastewater by struvite crystallization and CO2 degasification reactor. Ecol. Chem. Eng. S. 2014, 21, 89–99. [Google Scholar] [CrossRef]
  34. Wang, Y.; Liu, Z.; Huang, W.; Lu, J.; Luo, S.; Czech, B.; Li, T.; Wang, H. Capture-reduction mechanism for promoting Cr(VI) removal by sulfidated microscale zerovalent iron/sulfur-doped graphene-like biochar composite. Carbon Res. 2023, 2, 11. [Google Scholar] [CrossRef]
  35. Mumme, J.; Eckervogt, L.; Pielert, J.; Diakité, M.; Rupp, F.; Kern, J. Hydrothermal carbonization of anaerobically digested maize silage. Bioresour. Technol. 2011, 102, 9255–9260. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, L.; Chang, Y.; Li, A. Hydrothermal carbonization for energy-efficient processing of sewage sludge: A review. Renew. Sust. Energ. Rev. 2019, 108, 423–440. [Google Scholar] [CrossRef]
  37. Wu, Y.; Li, W.; Wu, Q.; Liu, S. Preparation, properties and applications of hydrochar. Prog. Chem. 2016, 28, 121–130. [Google Scholar]
  38. Funke, A.; Ziegler, F. Hydrothermal carbonization of biomass: A summary and discussion of chemical mechanisms for process engineering. Biofuels Bioprod. Biorefining 2010, 4, 160–177. [Google Scholar] [CrossRef]
  39. Kruse, A.; Funke, A.; Titirici, M.M. Hydrothermal conversion of biomass to fuels and energetic materials. Curr. Opin. Chem. Biol. 2013, 17, 515–521. [Google Scholar] [CrossRef]
  40. Sevilla, M.; Fuertes, A.B. The production of carbon materials by hydrothermal carbonization of cellulose. Carbon 2009, 47, 2281–2289. [Google Scholar] [CrossRef]
  41. Kang, S.; Li, X.; Fan, J. Characterization of hydrochars produced by hydrothermal carbonization of lignin, cellulose, d-xylose, and wood meal. Ind. Eng. Chem. Res. 2012, 51, 9023–9031. [Google Scholar] [CrossRef]
  42. Liu, W.; Jiang, H.; Yu, H. Thermochemical conversion of lignin to functional materials: A review and future directions. Green Chem. 2015, 17, 4888–4907. [Google Scholar] [CrossRef]
  43. Liu, Q.; Zhang, G.; Kong, G.; Liu, M.; Cao, T.; Guo, Z.; Zhang, X.; Han, L. Valorizing manure waste into green coal-like hydrochar: Parameters study, physicochemical characteristics, combustion behaviors and kinetics. Renew. Energy 2023, 216, 119103. [Google Scholar] [CrossRef]
  44. Yan, W.; Hoekman, S.K.; Broch, A.; Coronella, C.J. Effect of hydrothermal carbonization reaction parameters on the properties of hydrochar and pellets. Environ. Prog. Sustain. Energy 2014, 33, 676–680. [Google Scholar] [CrossRef]
  45. Reza, M.T.; Andert, J.; Wirth, B.; Busch, D.; Pielert, J.; Lynam, J.G.; Mumme, J. Hydrothermal carbonization of biomass for energy and crop production. Appl. Bioenergy 2014, 1, 11–29. [Google Scholar] [CrossRef]
  46. Lynam, J.G.; Reza, M.T.; Yan, W.; Vásquez, V.R.; Coronella, C.J. Hydrothermal carbonization of various lignocellulosic biomass. Biomass Convers. Biorefinery 2015, 5, 173–181. [Google Scholar] [CrossRef]
  47. Djandja, O.S.; Liew, R.K.; Liu, C.; Liang, J.; Yuan, H.; He, W.; Feng, Y.; Lougou, B.G.; Duan, P.-G.; Lu, X.; et al. Catalytic hydrothermal carbonization of wet organic solid waste: A review. Sci. Total Environ. 2023, 873, 162119. [Google Scholar] [CrossRef]
  48. Zhao, H.; Lu, X.; Wang, Y.; Sun, B.; Wu, X.; Lu, H. Effects of additives on sucrose-derived activated carbon microspheres synthesized by hydrothermal carbonization. J. Mater. Sci. 2017, 52, 10787–10799. [Google Scholar] [CrossRef]
  49. Petrovic, J.T.; Stojanovic, M.D.; Milojkovic, J.V.; Petrovic, M.S.; Sostaric, T.D.; Lausevic, M.D.; Mihajlovic, M.L. Alkali modified hydrochar of grape pomace as a perspective adsorbent of Pb2+ from aqueous solution. J. Environ. Manag. 2016, 182, 292–300. [Google Scholar] [CrossRef]
  50. Jiang, L.; Li, K.; Xia, L.; Gao, J.; Tang, L.; Jia, Y. KOH-modified hydrochar produced from Cd/Zn hyperaccumulator Sedum alfredii Hance for aqueous Cd(II) removal: Behavior and mechanism. J. Environ. Chem. Eng. 2023, 11, 110925. [Google Scholar] [CrossRef]
  51. Qian, W.; Luo, X.; Wang, X.; Guo, M.; Li, B. Removal of methylene blue from aqueous solution by modified bamboo hydrochar. Ecotoxicol. Environ. Saf. 2018, 157, 300–306. [Google Scholar] [CrossRef] [PubMed]
  52. Huang, Y.; Shen, D.; Wang, Z. Preparation of citric acid-sewage sludge hydrochar and its adsorption performance for Pb(II) in aqueous solution. Polymers 2022, 14, 968. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, T.; Zhai, Y.; Zhu, Y.; Peng, C.; Xu, B.; Wang, T.; Li, C.; Zeng, G. Acetic acid and sodium hydroxide-aided hydrothermal carbonization of woody biomass for enhanced pelletization and fuel properties. Energy Fuels 2017, 31, 12200–12208. [Google Scholar] [CrossRef]
  54. Chen, F.; Zhang, Y.; Zheng, M.; Xiao, Y.; Hu, H.; Liang, Y.; Liu, Y.; Dong, H. Preparation of high-performance porous carbon materials by citric acid-assisted hydrothermal carbonization of bamboo and their application in electrode materials. Energy Fuels 2022, 36, 9303–9312. [Google Scholar] [CrossRef]
  55. Huang, J.; Wang, Z.; Qiao, Y.; Wang, B.; Yu, Y.; Xu, M. Transformation of nitrogen during hydrothermal carbonization of sewage sludge: Effects of temperature and Na/Ca acetates addition. Proc. Combust. Inst. 2021, 38, 4335–4344. [Google Scholar] [CrossRef]
  56. Mumme, J.; Titirici, M.M.; Pfeiffer, A.; Lueder, U.; Reza, M.T.; Masek, O. Hydrothermal carbonization of digestate in the presence of zeolite: Process efficiency and composite properties. ACS Sustain. Chem. Eng. 2015, 3, 2967–2974. [Google Scholar] [CrossRef]
  57. Lang, Q.; Zhang, B.; Liu, Z.; Jiao, W.; Xia, Y.; Chen, Z.; Li, D.; Ma, J.; Gai, C. Properties of hydrochars derived from swine manure by CaO assisted hydrothermal carbonization. J. Environ. Manag. 2019, 233, 440–446. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, Y.; Chen, D.; Xing, Y.; Liu, B.; Zhou, Y.; Lu, P. Role of nitrate in the production of iron-modified hydrochar for arsenic removal. Water Air Soil Pollut. 2024, 235, 233. [Google Scholar] [CrossRef]
  59. Jiao, N.; Zhu, Y.; Li, H.; Yu, Y.; Xu, Y.; Zhu, J. Two-step hydrothermal pretreatments for co-producing xylooligosaccharides and humic-like acid from vinegar residue. Fermentation 2023, 9, 589. [Google Scholar] [CrossRef]
  60. Tan, J.; Chen, H.; Gao, Y.; Li, H. Nitrogen-doped porous carbon derived from citric acid and urea with outstanding supercapacitance performance. Electrochim. Acta 2015, 178, 144–152. [Google Scholar] [CrossRef]
  61. Leng, S.; Leng, L.; Chen, L.; Chen, J.; Chen, J.; Zhou, W. The effect of aqueous phase recirculation on hydrothermal liquefaction/carbonization of biomass: A review. Bioresour. Technol. 2020, 318, 124081. [Google Scholar] [CrossRef] [PubMed]
  62. Stemann, J.; Putschew, A.; Ziegler, F. Hydrothermal carbonization: Process water characterization and effects of water recirculation. Bioresour. Technol. 2013, 143, 139–146. [Google Scholar] [CrossRef] [PubMed]
  63. Uddin, M.H.; Reza, M.T.; Lynam, J.G.; Coronella, C.J. Effects of water recycling in hydrothermal carbonization of loblolly pine. Environ. Prog. Sustain. Energy 2014, 33, 1309–1315. [Google Scholar] [CrossRef]
  64. Zhu, X.; He, M.; Xu, Z.; Luo, Z.; Gao, B.; Ruan, R.; Wang, C.-H.; Wong, K.-H.; Tsang, D. Combined acid pretreatment and co-hydrothermal carbonization to enhance energy recovery from food waste digestate. Energy Conv. Manag. 2022, 266, 115855. [Google Scholar] [CrossRef]
  65. Wang, T.; Si, B.; Gong, Z.; Zhai, Y.; Cao, M.; Peng, C. Co-hydrothermal carbonization of food waste-woody sawdust blend: Interaction effects on the hydrochar properties and nutrients characteristics. Bioresour. Technol. 2020, 316, 123900. [Google Scholar] [CrossRef]
  66. Li, C.; Cai, R.; Hasan, A.; Lu, X.; Yang, X.; Zhang, Y. Fertility assessment and nutrient conversion of hydrochars derived from co-hydrothermal carbonization between livestock manure and corn cob. J. Environ. Chem. Eng. 2023, 11, 109166. [Google Scholar] [CrossRef]
  67. Sharma, H.B.; Dubey, B.K. Co-hydrothermal carbonization of food waste with yard waste for solid biofuel production: Hydrochar characterization and its pelletization. Waste Manag. 2020, 118, 521–533. [Google Scholar] [CrossRef] [PubMed]
  68. Shen, Q.; Zhu, X.Q.; Peng, Y.; Xu, M.; Huang, Y.; Xia, A.; Zhu, X.; Liao, Q. Structure evolution characteristic of hydrochar and nitrogen transformation mechanism during co-hydrothermal carbonization process of microalgae and biomass. Energy 2024, 295, 131028. [Google Scholar] [CrossRef]
  69. Xu, Y.; Lou, Z.; Yi, P.; Chen, J.; Ma, X.; Wang, Y.; Li, M.; Chen, W.; Liu, Q.; Zhou, J.; et al. Improving abiotic reducing ability of hydrothermal biochar by low temperature oxidation under air. Bioresour. Technol. 2014, 172, 212–218. [Google Scholar] [CrossRef]
  70. Zhou, S.; Han, L.; Yang, Z.; Ma, Q. Influence of hydrothermal carbonization temperature on combustion characteristics of livestock and poultry manures. Trans. Chin. Soc. Agric. Eng. 2017, 33, 233–240. [Google Scholar]
  71. Nguyen, D.; Zhao, W.X.; Mäkelä, M.; Alwahabi, Z.T.; Kwong, C.W. Effect of hydrothermal carbonisation temperature on the ignition properties of grape marc hydrochar fuels. Fuel 2022, 313, 122668. [Google Scholar] [CrossRef]
  72. Gao, P.; Zhou, Y.; Meng, F.; Zhang, Y.; Liu, Z.; Zhang, W.; Xue, G. Preparation and characterization of hydrochar from waste eucalyptus bark by hydrothermal carbonization. Energy 2016, 97, 238–245. [Google Scholar] [CrossRef]
  73. Peng, J.; Kang, X.; Zhao, S.; Zhao, P.; Ragauskas, A.J.; Si, C.; Xu, T.; Song, X. Growth mechanism of glucose-based hydrochar under the effects of acid and temperature regulation. J. Colloid Interface Sci. 2023, 630, 654–665. [Google Scholar] [CrossRef] [PubMed]
  74. Li, F.; Jiang, Z.; Ji, W.; Chen, Y.; Ma, J.; Gui, X.; Zhao, J.; Zhou, C. Effects of hydrothermal carbonization temperature on carbon retention, stability, and properties of animal manure-derived hydrochar. Int. J. Agric. Biol. Eng. 2022, 15, 124–131. [Google Scholar] [CrossRef]
  75. Maniscalco, M.P.; Volpe, M.; Messineo, A. Hydrothermal carbonization as a valuable tool for energy and environmental applications: A review. Energies 2020, 13, 4098. [Google Scholar] [CrossRef]
  76. Chen, H.; Gao, Y.; Li, J.; Fang, Z.; Bolan, N.; Bhatnagar, A.; Gao, B.; Hou, D.; Wang, S.; Song, H.; et al. Engineered biochar for environmental decontamination in aquatic and soil systems: A review. Carbon Res. 2022, 1, 4. [Google Scholar] [CrossRef]
  77. Liu, Z.; Wang, Z.; Chen, H.; Cai, T.; Liu, Z. Hydrochar and pyrochar for sorption of pollutants in wastewater and exhaust gas: A critical review. Environ. Pollut. 2021, 268, 115910. [Google Scholar] [CrossRef] [PubMed]
  78. Li, D.; Cui, H.; Cheng, Y.; Xue, L.; Wang, B.; He, H.; Hua, Y.; Chu, Q.; Feng, Y.; Yang, L. Chemical aging of hydrochar improves the Cd2+ adsorption capacity from aqueous solution. Environ. Pollut. 2021, 287, 117562. [Google Scholar] [CrossRef] [PubMed]
  79. Dos Reis, G.S.; Schnorr, C.E.; Dotto, G.L.; Vieillard, J.; Netto, M.S.; Silva, L.F.O.; De Brum, I.A.S.; Thyrel, M.; Lima, É.C.; Lassi, U. Wood waste-based functionalized natural hydrochar for the effective removal of Ce(III) ions from aqueous solution. Environ. Sci. Pollut. Res. 2023, 30, 64067–64077. [Google Scholar] [CrossRef]
  80. Tian, Y.; Yin, Y.; Liu, H.; Zhou, H. One-step hydrothermal carbonization of amine modified black liquor and lignin for efficient Cr(VI) adsorption. J. Water Process. Eng. 2022, 46, 102583. [Google Scholar] [CrossRef]
  81. Song, Q.; Ma, L.; Liu, J.; Bai, C.; Geng, J.; Wang, H.; Li, B.; Wang, L.; Li, S. Preparation and adsorption performance of 5-azacytosine-functionalized hydrothermal carbon for selective solid-phase extraction of uranium. J. Colloid Interface Sci. 2012, 386, 291–299. [Google Scholar] [CrossRef]
  82. Bedin, K.C.; Martins, A.C.; Cazetta, A.L.; Pezoti, O.; Almeida, V.C. KOH-activated carbon prepared from sucrose spherical carbon: Adsorption equilibrium, kinetic and thermodynamic studies for methylene blue removal. Chem. Eng. J. 2016, 286, 476–484. [Google Scholar] [CrossRef]
  83. Zhou, N.; Chen, H.; Xi, J.; Yao, D.; Zhou, Z.; Tian, Y.; Lu, X. Biochars with excellent Pb(II) adsorption property produced from fresh and dehydrated banana peels via hydrothermal carbonization. Bioresour. Technol. 2017, 232, 204–210. [Google Scholar] [CrossRef]
  84. Regmi, P.; Moscoso, J.L.G.; Kumar, S.; Cao, X.; Mao, J.; Schafran, G. Removal of copper and cadmium from aqueous solution using switchgrass biochar produced via hydrothermal carbonization process. J. Environ. Manag. 2012, 109, 61–69. [Google Scholar] [CrossRef] [PubMed]
  85. Xia, Y.; Yang, T.; Zhu, N.; Li, D.; Chen, Z.; Lang, Q.; Liu, Z.; Jiao, W. Enhanced adsorption of Pb(II) onto modified hydrochar: Modeling and mechanism analysis. Bioresour. Technol. 2019, 288, 121593. [Google Scholar] [CrossRef]
  86. Elaigwu, S.E.; Rocher, V.; Kyriakou, G.; Greenway, G.M. Removal of Pb2+ and Cd2+ from aqueous solution using chars from pyrolysis and microwave-assisted hydrothermal carbonization of prosopis africana shell. J. Ind. Eng. Chem. 2014, 20, 3467–3473. [Google Scholar] [CrossRef]
  87. Ramesh, S.; Sundararaju, P.; Banu, K.S.P.; Karthikeyan, S.; Doraiswamy, U.; Soundarapandian, K. Hydrothermal carbonization of arecanut husk biomass: Fuel properties and sorption of metals. Environ. Sci. Pollut. Res. 2019, 26, 3751–3761. [Google Scholar] [CrossRef] [PubMed]
  88. Zhang, Y.; Qu, J.; Yuan, Y.; Song, H.; Liu, Y.; Wang, S.; Tao, Y.; Zhao, Y.; Li, Z. Simultaneous scavenging of Cd(II) and Pb(II) from water by sulfide-modified magnetic pinecone-derived hydrochar. J. Clean Prod. 2022, 341, 130758. [Google Scholar] [CrossRef]
  89. Zhang, Z.; Zhu, Z.; Shen, B.; Liu, L. Insights into biochar and hydrochar production and applications: A review. Energy 2019, 171, 581–598. [Google Scholar] [CrossRef]
  90. Xia, M.; Chen, Z.; Li, Y.; Li, C.; Ahmad, N.M.; Cheema, W.A.; Zhu, S. Removal of Hg(II) in aqueous solutions through physical and chemical adsorption principles. RSC Adv. 2019, 9, 20941–20953. [Google Scholar] [CrossRef]
  91. Tian, S.-R.; Liu, Y.-G.; Liu, S.-B.; Zeng, G.-M.; Jiang, L.-H.; Tan, X.-F.; Huang, X.-X.; Yin, Z.-H.; Liu, N.; Li, J. Hydrothermal synthesis of montmorillonite/hydrochar nanocomposites and application for 17β-estradiol and 17α-ethynylestradiol removal. RSC Adv. 2018, 8, 4273–4283. [Google Scholar] [CrossRef]
  92. Ngoc, D.M.; Hieu, N.C.; Trung, N.H.; Chien, H.H.; Thi, N.Q.; Hai, N.D.; Chao, H. Tetracycline removal from water by adsorption on hydrochar and hydrochar-derived activated carbon: Performance, mechanism, and cost calculation. Sustainability 2023, 15, 4412. [Google Scholar] [CrossRef]
  93. Ma, Y.; Li, M.; Li, P.; Yang, L.; Wu, L.; Gao, F.; Qi, X.; Zhang, Z. Hydrothermal synthesis of magnetic sludge biochar for tetracycline and ciprofloxacin adsorptive removal. Bioresour. Technol. 2021, 319, 124199. [Google Scholar] [CrossRef] [PubMed]
  94. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X.; Wang, H.; Tsend, N. Hydrochars from bamboo sawdust through acid assisted and two-stage hydrothermal carbonization for removal of two organics from aqueous solution. Bioresour. Technol. 2018, 261, 257–264. [Google Scholar] [CrossRef] [PubMed]
  95. Li, Y.; Tsend, N.; Li, T.; Liu, H.; Yang, R.; Gai, X.; Wang, H.; Shan, S. Microwave assisted hydrothermal preparation of rice straw hydrochars for adsorption of organics and heavy metals. Bioresour. Technol. 2019, 273, 136–143. [Google Scholar] [CrossRef] [PubMed]
  96. Ronix, A.; Pezoti, O.; Souza, L.S.; Souza, I.P.A.F.; Bedin, K.C.; Souza, P.S.C.; Silva, T.L.; Melo, S.A.R.; Cazetta, A.L.; Almeida, V.C. Hydrothermal carbonization of coffee husk: Optimization of experimental parameters and adsorption of methylene blue dye. J. Environ. Chem. Eng. 2017, 5, 4841–4849. [Google Scholar] [CrossRef]
  97. Zbair, M.; Bottlinger, M.; Ainassaari, K.; Ojala, S.; Stein, O.; Keiski, R.L.; Bensitel, M.; Brahmi, R. Hydrothermal carbonization of argan nut shell: Functional mesoporous carbon with excellent performance in the adsorption of bisphenol a and diuron. Waste Biomass Valorization 2020, 11, 1565–1584. [Google Scholar] [CrossRef]
  98. Bai, C.-X.; Shen, F.; Qi, X.-H. Preparation of porous carbon directly from hydrothermal carbonization of fructose and phloroglucinol for adsorption of tetracycline. Chin. Chem. Lett. 2017, 28, 960–962. [Google Scholar] [CrossRef]
  99. Ghimire, S.; Wang, L.; Zhang, B.; Li, X.; Shahbazi, A. Production and modification of hydrochar from anaerobically digested cattail for adsorbing ammonium and phosphorous in wastewater. Water Sci. Technol. 2021, 84, 1678–1692. [Google Scholar] [CrossRef]
  100. Akter, J.; Islam, M.A.; Kibria, K.Q.; Islam, M.A. Adsorption of phosphate ions on chicken feather hydrochar and hydrochar-soil mixtures. Water Air Soil Pollut. 2021, 232, 413. [Google Scholar] [CrossRef]
  101. Fu, J.; Jin, J.; Ji, H.; Liu, Y.; Fu, X. Adsorption characteristics of arsenic and fluoride in water by magnetic hydrothermal carbon. J. Jiangsu Univ. Sci. Technol. (Nat. Sci.) 2019, 40, 423–430. [Google Scholar]
  102. Ge, X.; Chen, X.; Liu, M.; Wang, C.; Zhang, Y.; Wang, Y.; Tran, H.T.; Joseph, S.; Zhang, T. Toward a better understanding of phosphorus nonpoint source pollution from soil to water and the application of amendment materials: Research trends. Water 2023, 15, 1531. [Google Scholar] [CrossRef]
  103. Wang, G.; Zeng, W.; Li, S. Adsorption characteristics of phosphate on cerium modified water hyacinth biochar. Environ. Sci. 2021, 42, 4815–4825. [Google Scholar]
  104. Netto, M.S.; Georgin, J.; Franco, D.; Mallmann, E.S.; Foletto, E.L.; Godinho, M.; Pinto, D.; Dotto, G.L. Effective adsorptive removal of atrazine herbicide in river waters by a novel hydrochar derived from Prunus serrulata bark. Environ. Sci. Pollut. Res. 2022, 29, 3672–3685. [Google Scholar] [CrossRef] [PubMed]
  105. Gao, Y.; Wang, X.; Wang, J.; Li, X.; Cheng, J.; Yang, H.; Chen, H. Effect of residence time on chemical and structural properties of hydrochar obtained by hydrothermal carbonization of water hyacinth. Energy 2013, 58, 376–383. [Google Scholar] [CrossRef]
  106. Li, M.; Xiang, P.; Jiang, W.; Zhang, Z.; Mo, J. Performance and mechanism of MnFe2O4@HTC activated persulfate system for algae removal. J. Environ. Sci. 2021, 41, 3535–3544. [Google Scholar]
  107. Waly, S.M.; El-Wakil, A.M.; Abou El-Maaty, W.M.; Awad, F.S. Hydrothermal synthesis of Mg/Al-layered double hydroxide modified water hyacinth hydrochar for remediation of wastewater containing mordant brown dye. RSC Adv. 2024, 14, 15281–15292. [Google Scholar] [CrossRef]
  108. Takaya, C.A.; Fletcher, L.A.; Singh, S.; Anyikude, K.U.; Ross, A.B. Phosphate and ammonium sorption capacity of biochar and hydrochar from different wastes. Chemosphere 2016, 145, 518–527. [Google Scholar] [CrossRef]
  109. Yu, Y.; Han, L.; Jiang, X. Production, properties and environmental application of hydrochar. Environ. Chem. 2018, 37, 1232–1244. [Google Scholar]
  110. Marco, K.; Peter, S.N.; Mark, G.J.; Markus, K. Dynamic molecular structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44, 1247–1253. [Google Scholar]
  111. Vozhdayev, G.V.; Spokas, K.A.; Molde, J.S.; Heilmann, S.M.; Wood, B.M.; Valentas, K.J. Response of maize germination and growth to hydrothermal carbonization filtrate type and amount. Plant Soil 2015, 396, 127–136. [Google Scholar] [CrossRef]
  112. Ni, P.-Y.; Zhang, X.; Ye, M.; He, R. Biochar enhanced the stability of toluene removal in extracted groundwater amended with nitrate under microaerobic conditions. Chemosphere 2024, 353, 141551. [Google Scholar] [CrossRef]
  113. Kumar, S.; Loganathan, V.A.; Gupta, R.B.; Barnett, M.O. An assessment of U(VI) removal from groundwater using biochar produced from hydrothermal carbonization. J. Environ. Manag. 2011, 92, 2504–2512. [Google Scholar] [CrossRef]
  114. Chung, J.W.; Breulmann, M.; Clemens, A.; Fühner, C.; Foppen, J.W.; Lens, P. Simultaneous removal of rotavirus and adenovirus from artificial ground water using hydrochar derived from swine feces. J. Water Health 2016, 14, 754–767. [Google Scholar] [CrossRef] [PubMed]
  115. Wang, C.; Gui, B.; Wu, C.; He, C.; Li, L.; Ling, X.; Zuo, X. Hydrothermal carbonization of sewage sludge coupled with fenton oxidation pretreatment: Moderate oxidation to enhance hydrochar yield and properties. J. Environ. Chem. Eng. 2023, 11, 110788. [Google Scholar] [CrossRef]
  116. Shi, Y.; Li, C.; Chai, R.; Wu, J.; Wang, Y. Effect of different hydrothermal parameters on calorific value and pyrolysis characteristics of hydrochar of kitchen waste. Energies 2023, 16, 3561. [Google Scholar] [CrossRef]
  117. Marx, S.; van der Merwe, K. Utilization of hydrochar derived from waste paper sludge through hydrothermal liquefaction for the remediation of phenol contaminated industrial wastewater. Water Pract. Technol. 2021, 16, 756–771. [Google Scholar] [CrossRef]
  118. He, X.; Zhang, T.; Niu, Y.; Xue, Q.; Ali, E.F.; Shaheen, S.M.; Tsang, D.C.W.; Rinklebe, J. Impact of catalytic hydrothermal treatment and Ca/Al-modified hydrochar on lability, sorption, and speciation of phosphorus in swine manure: Microscopic and spectroscopic investigations. Environ. Pollut. 2022, 299, 118877. [Google Scholar] [CrossRef] [PubMed]
  119. Zhang, T.; Wu, X.; Fan, X.; Tsang, D.C.; Li, G.; Shen, Y. Corn waste valorization to generate activated hydrochar to recover ammonium nitrogen from compost leachate by hydrothermal assisted pretreatment. J. Environ. Manag. 2019, 236, 108–117. [Google Scholar] [CrossRef] [PubMed]
  120. Luo, J.; Zhou, W.; Xu, B. A deep neural network-based assistive decision method for financial risk prediction in carbon trading market. J. Circuits Syst. Comput. 2023, 33, 2450153. [Google Scholar] [CrossRef]
  121. Luo, J.; Zhou, W.; Liu, S.; Xu, B. The Optimization of carbon emission prediction in low carbon energy economy under big data. IEEE Access 2024, 12, 14690–14702. [Google Scholar] [CrossRef]
  122. Zheng, C.; Chen, H. Revisiting the linkage between financial inclusion and energy productivity: Technology implications for climate change. Sustain. Energy Technol. Assess. 2023, 57, 103275. [Google Scholar] [CrossRef]
  123. Jia, B.; Zhou, G. Estimation of global karst carbon sink from 1950s to 2050s using response surface methodology. Geo-Spat. Inf. Sci. 2023, 2165974. [Google Scholar] [CrossRef]
  124. Zheng, S.; Hai, Q.; Zhou, X.; Stanford, R.J. A novel multi-generation system for sustainable power, heating, cooling, freshwater, and methane production: Thermodynamic, economic, and environmental analysis. Energy 2024, 290, 130084. [Google Scholar] [CrossRef]
  125. Wang, Q.; Zhang, T.; He, X.; Jiang, R. Assessment of phosphorus recovery from swine wastewater in Beijing, China. Sustainability 2017, 9, 1845. [Google Scholar] [CrossRef]
  126. Yang, C.; Nutakki, T.U.K.; Alghassab, M.A.; Alkhalaf, S.; Alturise, F.; Alharbi, F.S.; Elmasry, Y.; Abdullaev, S. Optimized integration of solar energy and liquefied natural gas regasification for sustainable urban development: Dynamic modeling, data-driven optimization, and case study. J. Clean Prod. 2024, 447, 141405. [Google Scholar] [CrossRef]
Figure 1. The distribution of products and phosphate in different phases under different thermal treatments [23].
Figure 1. The distribution of products and phosphate in different phases under different thermal treatments [23].
Water 16 01749 g001
Figure 2. The number of papers on hydrochars applied in wastewater treatment.
Figure 2. The number of papers on hydrochars applied in wastewater treatment.
Water 16 01749 g002
Figure 3. Hop map of keywords for articles on the use of hydrochar using VOSviewer1.6.19 software.
Figure 3. Hop map of keywords for articles on the use of hydrochar using VOSviewer1.6.19 software.
Water 16 01749 g003
Figure 4. HTC reaction pathways from lignocellulose biomass [45].
Figure 4. HTC reaction pathways from lignocellulose biomass [45].
Water 16 01749 g004
Figure 5. History of HTC technology development.
Figure 5. History of HTC technology development.
Water 16 01749 g005
Figure 6. The general steps of HTC technology.
Figure 6. The general steps of HTC technology.
Water 16 01749 g006
Figure 7. Schematic diagram of the HTC process for treating wastewater—liquid and solid flows.
Figure 7. Schematic diagram of the HTC process for treating wastewater—liquid and solid flows.
Water 16 01749 g007
Figure 8. Main Ce(III) adsorption mechanism onto hydrochar surface [79].
Figure 8. Main Ce(III) adsorption mechanism onto hydrochar surface [79].
Water 16 01749 g008
Figure 9. Proposed mechanism for TC adsorption by hydrochar, modified-hydrochar, and hydrochar-derived ACs [92].
Figure 9. Proposed mechanism for TC adsorption by hydrochar, modified-hydrochar, and hydrochar-derived ACs [92].
Water 16 01749 g009
Table 1. Research on HTC technology using catalysts.
Table 1. Research on HTC technology using catalysts.
Raw MaterialsHydrothermal ConditionsCatalyst TypeModification TreatmentProductReferences
Grape pomace220 °C, 1 hhydroxidesKOHThe KOH treatment increased the sorption capacity for Pb2+ from 27.8 mg/g up to 137 mg/g at pH 5.[49]
Sedum Alfredii Hance180–270 °C, 5 hhydroxidesKOHThe adsorption capacity of hydrochar for Cd(II) was greatly enhanced after KOH modification, and the maximum Cd(II) adsorption capacity was 25.69 mg/g.[50]
Bamboo sawdust473 K, 24 hhydroxidesNaOHThe adsorption capacity of hydrochar to methylene blue after modified treatment is 655.76 mg/g.[51]
Sludge180 °C, 2 hacidscitric acidFeatured a maximum surface area of 59.95 m2/g, and exhibited the highest equilibrium adsorption capacity for Pb(II).[52]
Sawdust200–250 °CacidsCH3COOHHad a higher calorific value and energy density.[53]
Bamboo powder180 °C, 6 hacidscitric acidCitric acid could improve the carbonization degree of hydrochar, and optimize the pore structure distribution.[54]
Raw sewage sludge160–250 °C, 30 minsaltsCaAc2, NaAcThe addition of calcium acetate can reduce the retention of nitrogen in the hydrate, and the addition of sodium acetate can slightly enhance the hydrolysis of the protein.[55]
Cow dung and corn digestate190–270 °Cmolecular sievesnatural zeoliteIncreases the surface area and pore volume of the hydrochar, and significantly boosts the recovery rates of nitrogen, sulfur and energy.[56]
Pig manure180, 200, and 220 °C, 10 hmetal oxidesCaOFacilitates the transformation of phosphorus from non-apatite inorganic phosphorus (NAIP) to apatite phosphorus (AP).[57]
Table 2. Research on co-HTC.
Table 2. Research on co-HTC.
Raw MaterialsHydrothermal ConditionsProductReferences
Food waste digestate (FWD), wood waste (WW)220 °C, 4 hThe comprehensive combustion and combustion stability indices were improved with the maximum values of 3.98 × 10−9 %2°C−5 and 4.22 × 102 %°C−3, respectively.[64]
Food waste (FW), woody sawdust (WS)180–260 °C, 1 h, and a stirrer speed of 100 r/minHydrochar yield consistently decreased with the increase in both the FW ratio and HTC temperature, and the C retention from 260 °C hydrochar was low (approximately 65%).[65]
Livestock manure (SM), corn cob (CC)180–260 °C, 1–4 hThe addition of CC could effectively reduce the ash content, and enhance the N recovery to 38.95–47.61% (SM:CC = 1:1).[66]
Food waste (FW), yard waste (YW)220–260 °C, 1–4 hThe sulfur in SS and CS gradually converted into thiophenic sulfur and sulfates.[67]
Corn straw, chlorella240 °C, 1 hModerate reaction conditions favored nitrogen enrichment and increased porosity in the hydrochar.[68]
Table 3. Research on HTC at different temperatures.
Table 3. Research on HTC at different temperatures.
Raw MaterialsHydrothermal ConditionsProductReferences
Spartina alterniflora biomass240 °C, 4 hThe contents of carboxyl group and carbonyl group on the surface of hydrochar were increased by low temperature oxidation.[69]
Livestock manure180–240 °C, 1 hBoth hydrochar yield and carbon retention decreased with increasing reaction temperatures.[70]
Grape marc180–260 °C, 30 minThe 260 °C solid hydrochar exhibited the shortest ignition delay time (0.2 s) and the lowest ignition temperature (179 °C).[71]
Eucalyptus bark220–300 °C, 2–10 hThe oxygen-containing functional group decreased with the increase in temperature.[72]
Glucose160–220 °C, 1–12 hThe increase in the reaction temperature (160–220 °C) enhanced the aromatization degree of hydrochar.[73]
Chicken, dairy, swine manures200–350 °C, 2 hWith the increase in temperature, the carbon retention rate decreased, and the aromatics and thermal oxidation resistance is increased.[74]
Table 4. Methods for different types of HTC.
Table 4. Methods for different types of HTC.
TypeMethodsAdvantage
HTCBiomass is mixed with water and directly carbonized under specific temperature and pressure conditions.Makes full use of biomass resources, less process, and low difficulty to operate.
HTC with catalystA certain amount of catalyst was added to modify the hydrochar.Complex structure and excellent performance, which improves the HTC yield and energy recovery efficiency.
Multi-step HTCThe reaction conditions are manipulated through multiple stages of varying temperatures and durations.Has unique pore structure and surface characteristics, which optimizes the structure, improves the carbonization efficiency and reduces the energy loss.
Liquid phase circulating HTCRepeatedly recycle by-products of hydrothermal fluid.Improves the conversion rate of biomass and utilization efficiency of water resources and reduces the cost of carbonization process and wastewater treatment.
Co-HTCMix a variety of raw materials together and carry out HTC at the same time.Reduces the carbonization temperature, and improves the material performance, combustion performance and resource utilization rate.
High temperature HTCThe hydrothermal temperature is higher in the carbonization process.High calorific value, and excellent product stability, with long-term carbon sequestration potential.
Low temperature HTCThe hydrothermal temperature is low in the carbonization process.Has better hydrochar yield, carbon recovery and energy recovery, and has an obvious short-term carbon sequestration effect.
Table 5. Research on adsorption of heavy metal ions by hydrochar.
Table 5. Research on adsorption of heavy metal ions by hydrochar.
Raw MaterialsHydrothermal ConditionsModification TreatmentAdsorbatesProductReferences
Black liquor150, 180, 200, 220 °C, and 0.5 hhexamethylenediamine (HMDA)50 mL of Cr(VI) solutionsThe maximum adsorption capacity of Cr(VI) was 741.74 mg/g at 45 °C and pH 2.[80]
Glucose180 °C, 48 h——20.00 mL of pure uranyl ion (UO22+) solutionsThe highly selective separation of U(VI) was achieved, and the maximum adsorption capacity reached 408.36 mg at 333.15 K and pH 4.5.[81]
Sucrose190 °C, 11 hKOH activated25.0 mL of methylene blue solutions (500 mg/L)The maximum adsorption capacity of methylene blue was 704.2 mg/g.[82]
Banana peels230 °C, 2 hphosphoric acid40 mL of Pb2+ stock solutions (2000 mg/L)The adsorption capacities of Pb2+ for dehydrated and fresh banana peel hydrochar were 359 mg/g and 193 mg/g, respectively.[83]
Willow twigs300 °C, 30 min——50 mL aqueous copper or cadmium solutions (40 mg/L)The hydrochar exhibited adsorption capacities of 34 mg/g (0.313 mmol/g) for Cd2+ and 31 mg/g (0.503 mmol/g) for Cu2+.[84]
Pine sawdust 260 °C, 2 hH2O225 mL of Pb2+ solutions (1–200 mg/L)The maximum adsorption capacity of Pb2+ reached 92.80 mg/g at pH 5.0 and 298 K.[85]
Pseudomonas aeruginosa shells200 °C, 20 min——50 mL of Pb2+ or Cd2+ aqueous solution (10–100 mg/L)It was found to be more effective at adsorbing Pb2+ and Cd2+ compared to biochar.[86]
Arecanut husk180–220 °C, 9 hcitric acidZn2+, Cr6+, Ni2+, Pb2+ solutions (0–100 mg/L)The maximum removal efficiency of Pb2+ in wastewater was 95.08% at 25 mg/L.[87]
Pinecone200 °C, 5 hFeCl3·6H2O, CH3COONa, Na3C6H5O7100 mL of Cd(II)/Pb(II) solutions (10–300 mg/L)The maximum single-layer adsorption capacity of MHC-S4 for Cd (II) and Pb (II) was 62.49 and 149.33 mg/g, respectively.[88]
Table 6. Research on adsorption of organic pollutants by hydrochar.
Table 6. Research on adsorption of organic pollutants by hydrochar.
Raw MaterialsHydrothermal ConditionsModification TreatmentAdsorbatesProductReferences
Municipal sludge600 °C, 2 hZn/Fe0.10–0.30 g/L TC, 0.20–0.50 g/L CIPThe product demonstrated maximum adsorption capacities of 145.0 mg/g for TC and 74.2 mg/g for CIP.[93]
Bamboo shavings200 °C, 3 hacid assisted30 mL of Congo red or 2-naphthol solutions (0.5 mg/mL)The resultant material exhibited maximum adsorption capacities of 90.51 mg/g for Congo Red and 72.93 mg/g for 2-naphthol.[94]
Montmorillonite and rice husks180 °C, 16 hKOH100 mL of estrogen stock solutions (2.5 mg/mL)KOH improved the adsorption capacity of estrogen.[91]
Rice straw160–200 °C, 40–70 min——30 mL of Congo red, berberine hydrochloride, 2-naphthol, Zn2+ and Cu2+ solutions (0.5 mg/mL)The maximum adsorption capacities for Congo Red, berberine hydrochloride, 2-naphthol, Zn2+, and Cu2+ were 222.1 mg/g, 174.0 mg/g, 48.7 mg/g, 112.8 mg/g, and 144.9 mg/g, respectively.[95]
Coffee husks210 °C, 243 h——25.0 mL of the MB solutions (300 mg/L)The adsorption capacity of methylene blue was 34.85 mg/g, mainly by physical adsorption.[96]
Moroccan nut shells (ANS)180, 200 °C, and 6 h.——200 mL of BPA (60 mg/L) and diuron (40 mg/L) solutionsThe hydrochar generated at 200 °C removed about 92% and 95% of BPA and diuron, respectively, with high adsorption efficiency.[97]
Fructose180 °C, 2 hphloroglucinol25 mL of tetracycline solutionsThe adsorption capacity of the product to tetracycline in water was 274.7 mg/g.[98]
Table 7. Research on adsorption of inorganic anions by hydrochar.
Table 7. Research on adsorption of inorganic anions by hydrochar.
Raw MaterialsHydrothermal ConditionsModification TreatmentAdsorbatesProductReferences
Anaerobically digested cattails200, 250 and 300 °C, 1 hCH3COOH and NaOH1470 mg/L COD, 50 mg/L NH4+-N and 151 mg/L TPThe adsorption capacities of NH4+-N and PO43−-P were 92.6–122.4 mg/g and 1.6–15.8 mg/g, respectively.[99]
Chicken feathers150–170 °C, 1–3 h——0–150 mg/L KH2PO4The maximum adsorption capacity (qm) of hydrochar was 21.70 mg/g at 30 °C.[100]
Microalgae348 K, 40 minMg20 mL of P solutions (20–15,000 mg/L)The modified hydrochar exhibited strong affinity for phosphate, with a maximum adsorption capacity of 89.61 mg/g.[25]
Sodium alginate210 °C, 5 hFe3O4As (V) solutions (1–100 mg/L), F solutions (1–50 mg/L)At a dosage of 2 g/L, the maximum adsorption capacity of arsenic and fluoride was 26.06 mg/g and 15.64 mg/g, respectively.[101]
Table 9. Comparison of different water environment treatment technologies.
Table 9. Comparison of different water environment treatment technologies.
Type of TechnologyEnvironmental Management EffectivenessAdvantageShortcomingEnergy Consumption
AdsorptionEffective removal of organic matter and certain inorganic substancesEasy to operate, high removal efficiency can be achieved. Adaptable, adsorbents can be reused.The cost of adsorbents can be high, and energy is required to regenerate or replace adsorbents.Medium to high
Membrane technologySeparation and purification of particulates and dissolved substances in water bodiesModular design, easy to expand, high separation efficiency, and high degree of automation.Membrane materials can be expensive, with membrane fouling and membrane replacement costs, and may require chemical cleaning.Medium to high
PrecipitationRemoval of suspended solids and certain dissolved solids from waterThe processing capacity is large, the operating cost is relatively low, and the application range is wide.A large amount of sedimentation tank space is required and may produce a large amount of sludge that is difficult to control.Low to moderate
Enzyme technologyDecomposes and transforms organic matterEfficient for specific pollutants, pollutants can be converted into harmless substances, reducing sludge production.Enzyme stability may be low, requiring appropriate Ph and temperature conditions, and initial enzyme cost.Medium to low
HTCEfficient treatment of organic waste into high-value productsEnvironmental protection, reduce secondary pollution, good resource effect, high treatment efficiency under high-temperature and high-pressure conditions.Equipment costs are higher, dependence on biomass or carbon-containing waste, and energy consumption can be higher.high
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, G.; Xu, Q.; Abou-Elwafa, S.F.; Alshehri, M.A.; Zhang, T. Hydrothermal Carbonization Technology for Wastewater Treatment under the “Dual Carbon” Goals: Current Status, Trends, and Challenges. Water 2024, 16, 1749. https://doi.org/10.3390/w16121749

AMA Style

Liu G, Xu Q, Abou-Elwafa SF, Alshehri MA, Zhang T. Hydrothermal Carbonization Technology for Wastewater Treatment under the “Dual Carbon” Goals: Current Status, Trends, and Challenges. Water. 2024; 16(12):1749. https://doi.org/10.3390/w16121749

Chicago/Turabian Style

Liu, Guoqing, Qing Xu, Salah F. Abou-Elwafa, Mohammed Ali Alshehri, and Tao Zhang. 2024. "Hydrothermal Carbonization Technology for Wastewater Treatment under the “Dual Carbon” Goals: Current Status, Trends, and Challenges" Water 16, no. 12: 1749. https://doi.org/10.3390/w16121749

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop