Next Article in Journal
Optimal Selection of Switch Model Parameters for ADC-Based Power Converters
Previous Article in Journal
Effective Design Methodology of CLLC Resonant Converter Based on the Minimal Area Product of High-Frequency Transformer
Previous Article in Special Issue
Techno-Economic Feasibility of Biomass Gasification for the Decarbonisation of Energy-Intensive Industries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Insights into Properties of Biomass Energy Pellets Made from Mixtures of Woody and Non-Woody Biomass: A Meta-Analysis

by
Rajitha Lakshan Rupasinghe
1,
Priyan Perera
1,*,
Rangika Bandara
2,
Hiran Amarasekera
1 and
Richard Vlosky
3
1
Department of Forestry and Environmental Science, University of Sri Jayewardenepura, Colombo 10250, Sri Lanka
2
Department of Zoology and Environmental Management, University of Kelaniya, Colombo 11300, Sri Lanka
3
Louisiana Forest Products Development Center, Louisiana State University, Baton Rouge, LA 70803, USA
*
Author to whom correspondence should be addressed.
Energies 2024, 17(1), 54; https://doi.org/10.3390/en17010054
Submission received: 9 November 2023 / Revised: 1 December 2023 / Accepted: 6 December 2023 / Published: 21 December 2023
(This article belongs to the Special Issue Sustainable Energy from Biomass and Waste)

Abstract

:
There is a widespread global shift toward renewable energy sources, where the emphasis is on enhancing the utilization of renewable energy due to the rising costs associated with fossil fuels. In this light, biomass pellets made from woody and non-woody biomass and blends have gained increased attention. Extensive research has been conducted globally to enhance the quality of biomass pellets and to explore the potential to combine woody biomass with other non-woody forms of biomass in biomass pellet production. The heterogeneity of the raw materials used and resulting properties of the biomass pellets have led to the establishment of internationally recognized benchmarks such as the International Organization for Standardization (ISO) 17225 standard to regulate pellet quality. In this article, the key mechanical, physical, chemical, and energy properties of pellets made of different non-woody herbaceous biomass are investigated, and the available test values for such properties of the pellets were meta-analyzed. A comparison of the properties of these pellets with the relevant standards was also performed. A meta-analysis of studies on biomass pellet production was conducted via a comprehensive Systematic Literature Review (SLR). The SLR focuses on determining and analyzing the average values for the key physical properties of biomass pellets using woody biomass as a component in concert with other biomass materials. In addition, the optimal range of mixtures of woody and non-woody biomass was reviewed to produce biomass pellets with potential acceptance in the marketplace. The majority of studies included in the SLR concentrate on pellets made from a mixture of biomass materials. The results show that the average values for wood/non-wood mixtures such as pellet diameter, pellet length, moisture content, ash content, fine particle content, gross calorific value, and bulk density were found to adhere to the ISO standards. However, the average mechanical durability fell short of meeting the requirements of the standards. Additional comparisons were nitrogen, sulfur, volatile matter, and fixed carbon content. The findings in this meta-analysis could be useful in directing future research focused on producing high-quality and efficient biomass pellets derived from biomass blends and mixtures.

1. Introduction

1.1. Biomass as a Renewable Energy Source

The adoption of renewable energy sources is rapidly gaining momentum worldwide due to the growing global demand for energy [1,2,3]. There is a significant decrease in global concern about and dependency on fossil fuel energy sources, attributed to various factors such as fluctuations in energy demand, oil price shocks, disruptions in energy supply chains, hampered energy investments, energy price hikes, and energy security challenges [4,5]. Moreover, the urgency to address climate change and pursue low-carbon energy transitions has become a top priority in the energy sector [6]. Consequently, numerous countries have implemented policies to integrate environmentally friendly energy sources into their energy portfolios [7]. Notably, in October 2023, the European Union officially approved an updated Renewable Energy Directive aimed at increasing the share of renewable energy in Europe from 32% to 42.5% by 2030, with the ultimate goal of achieving a 45% share of renewables [8]. While specific targets for individual countries have not been established, each Member State will contribute to this collective objective. Concurrently, renewable resources such as solar, wind, geothermal, biogas, and biomass are gaining substantial recognition as viable options for sustainable and eco-friendly energy [9].
Within this realm of renewable energy, biomass has emerged as a pivotal contender over the last few decades. Its ascendancy is attributed to its renewable nature, environmental cleanliness, robust technical viability, economic feasibility, and widespread availability [10,11,12,13,14]. Moreover, biomass holds a distinct appeal as a renewable reservoir readily transformable into three distinct fuel states—gas, liquid, and solid [15,16,17,18,19,20,21].
Wood possesses the distinct advantage of negligible sulfur content, distinct from coal and liquid fuels, thus mitigating the emission of sulfur dioxide into the atmosphere [22]. Recent scientific inquiries have substantiated biomass as a key energy source with the potential to supplant fossil fuels [23]. Within this context, biomass emerges as a promising remedy to the challenges posed by fossil fuels, including coal and liquid fossil fuels, which are implicated in critical environmental concerns such as climate change, global warming, and their deleterious impact on human well-being [24]. Biomass is important in addressing such predicaments associated with fossil fuels [25].
According to the Statistical Report of Bioenergy Landscape 2020 [26], biomass-derived energy holds the second position in global bioenergy consumption, following nuclear energy, with a substantial market share of 63.11% (123,592 kilotons of oil equivalent (ktoe)), followed by hydro energy at 16.46% (32,242 ktoe) and wind energy at 11.11% (21,768 ktoe) [27,28].
The importance of bioenergy reaches far beyond developed nations and plays a pivotal role in developing nations. Recent studies have shed light on its impressive ability to deliver energy in various forms that cater to people’s needs, encompassing liquid and gaseous fuels, heat, and electricity. Therefore, bioenergy plays a significant role in reducing poverty in developing countries while simultaneously tackling the restoration of unproductive and degraded lands [29,30]. This restoration process yields multiple benefits, such as increased biodiversity, enhanced soil fertility, and improved water retention [31,32,33]. Bioenergy remains the primary source of energy in several countries and regions, including Bhutan (86%), Nepal (97%), Asia (16%), East Sahelian Africa (81%), and Africa (39%). In these areas, bioenergy is predominantly utilized for cooking and heating purposes, wherein firewood serves as the main source [31,34]. Particularly, Southeast Asia is rapidly emerging as a vibrant market for the development of biomass as an energy source [35]. Notably, countries such as Malaysia, Thailand, and Indonesia, known for their significant agricultural residues comprising rice, sugarcane, palm oil, coconut, and rubber, are among the foremost producers. Noteworthy crop residues include rice husk, sugarcane bagasse, oil palm residues, and wood residues [36]. The trajectory of bioenergy is witnessing novel trends and growing markets across the globe, with projections indicating that bioenergy will meet 30% of the world’s energy demand by 2050 [37,38].
While various forms of biomass, including wood, energy crops, agricultural residues, industrial wastes, and municipal solid waste, are available [39], the utilization of raw biomass is accompanied by certain inefficiencies. Factors such as irregular shapes, low bulk density, and elevated moisture content contribute to challenges in handling, transportation, and storage [40,41,42,43,44,45,46,47]. To tackle these issues, intensive research and implementation of biomass conversion technologies have transpired over the past decade [48,49,50,51,52].
Densification of biomass has emerged as a prominent conversion technology, achievable using distinct processes: pelletization, briquetting, extrusion, and tumbling [53]. This introduction of densification technologies has paved the way for the energy market entry of densified biomass products such as chipped wood, wooden pellets, and biomass briquettes. Moreover, the research underscores the consistent global consumption of firewood and charcoal, along with a twofold increase in the use of wood chips and wood pellets for power generation and residential heating over the past decade. This upward trajectory is projected to persist in years to come [12,13,37,54,55,56].

1.2. Biomass Pellet Market Dynamics

Biomass pellets, whether with or without additives, are compacted milled biomass typically cylindrical in shape, spanning 5 to 40 mm in standard market length [57]. The surging popularity of wood pellets in heating markets has triggered novel market dynamics and supply chains. Building and industrial heating and cooling in the European Union constitute 50% of its annual energy consumption [58], with 80% of central heating systems in Germany adopting biomass combustion technologies [59]. Similarly, growing demand for wood pellets as a heat source are observed in both the European Union and Asian countries [60].
In the Asia Pacific region, boasting 76% of the global coal generation capacity and 94% of the new coal plant pipeline [61], wood pellets are positioned as potential coal replacements in power generation. Via processes like torrefaction, hydrothermal carbonization, and steam explosion, wood pellets have gained thermal enhancements to mimic coal properties, advancing their suitability as a fuel [12,62,63,64]. Given the high concentration of coal power plants in the Asia Pacific region, their adoption of biomass pellets has risen, leading to exponential growth in wood pellet imports to South Korea, Japan, and China in recent years. Notably, South Korea’s imports surged to 2.4 million tons in 2017, a 20-fold increase from 2012 [65]. Similarly, Japan’s 2017 imports exceeded 0.5 million tons, marking a sevenfold rise since 2012 [65,66]. China, with its large population and energy source constraints, has established a substantial potential market. Though ample literature is lacking to substantiate the attainment of the 15-million-kilowatt goal set in its 2016 five-year plan, China stands as the primary producer of bioelectricity, witnessing a 4.5-fold rise in production since 2011 [67].
Approximately half of global pellet consumption serves power generation plants that have transitioned from coal to pellets or engage in co-firing with coal. The other half is predominantly allocated to household heat generation via pellet stoves, boilers, and for industrial steam demand [68,69,70,71,72]. Amidst this landscape, firewood, paraffin, electricity, liquid gas, and natural gas stand as principal competitors to wood pellets in energy generation. However, only firewood surpasses pellets economically; other energy sources falter in terms of toxic emissions, expensive handling, storage, and transportation when compared to biomass pellets [73].
Numerous sustainable indicators and multi-criteria decision analysis research conducted in Germany underscore wood pellets’ superior quality and efficiency for private households compared to alternative biomass-to-energy pathways [59,74,75]. The low density of unprocessed biomass such as wood chips (180–220 kg/m3) poses significant handling and transport challenges, unlike pellets, which offer higher density (around 600 kg/m3) and energy content per unit volume, thereby reducing costs in transportation, storage, handling, and use [68,76]. Unlike raw biomass, biomass pellets align more closely with liquid fuels in terms of their properties [73,76].

1.3. Quality Assurance of Biomass Pellets

Biomass pellets must adhere to standardized properties to optimize their utility. Designing boilers, stoves, or pellet burners aligned with these properties ensures effective deployment, catering to diverse scales of demand, from domestic appliances to large-scale power plants [68]. The primary parameters within pellet standards encompass physical attributes such as dimensions, mechanical durability, fine particle content, bulk and unit densities, additives, chemical composition, including sulfur, nitrogen, chlorine, and heavy metals, and energy properties such as moisture and ash content, net calorific value, and energy density [41,77]. These parameters are tied to raw materials, quality management, and manufacturing processes [78].
During 2000–2006, the European Committee for Standardization (CEN) under committee TC 335 established general technical specifications (TS) and testing methods for solid biofuels, culminating in the prEN14961 series by 2014 [79,80]. To align standards globally due to escalating biomass energy production and trade, these specifications transitioned to the International Organization for Standardization (ISO) via the Technical Committee: ISO TC 238 of Solid Biofuels [81]. The ISO released the EN ISO 17225 series in 2014 (ISO 17225-2:2014 [82], ISO 17225-6:2014 [83]), encompassing standards for wood pellets, chips, firewood, and non-woody briquettes, replacing EN 14961 [81,84].
EN ISO 17225-2 [82] for graded wood pellets sets limits for various applications, while EN ISO 17225-6 [83] focuses on non-woody pellets, including blends and mixtures (Table 1). Both standards underwent minor updates in 2020 republished in 2021 [85,86]. Graded wood pellets encompass property classes A1, A2, A3, I1, I2, and I3, with distinct quality characteristics for different applications [85]. Non-woody pellets, derived from diverse biomasses, bear higher ash, chlorine, nitrogen, and sulfur contents, warranting tailored combustion systems and corrosion mitigation due to their unique characteristics [84].
In this article, the key mechanical, physical, chemical, and energy properties of pellets made of different non-woody herbaceous biomass are investigated, and the available test values for such properties of the pellets are meta-analyzed. A comparison of the properties of these pellets with the relevant standards is also performed.
This study aims to conduct a comprehensive analysis of recent studies that have been conducted on pellets produced from various non-woody herbaceous biomass sources. The research will include a thorough comparison of the finest pellets from the reviewed studies against globally recognized standards. Furthermore, this investigation will seek to identify the most suitable types of biomass for efficient pellet production. By examining the quality differentiation associated with different biomass types and blending ratios, this study will provide valuable insights into the optimal combination of biomass materials for pellet production. In addition to that, this review aims to discuss the permissible limits or property ranges of the different standards that align with the desired pellet characteristics. This will enable the identification of biomass-bended pellet compositions that are compatible or incompatible with the specified quality standards. Following these objectives, this study seeks to contribute to the advancement of knowledge in the field of non-woody herbaceous biomass pellet production and guide industry professionals in making informed decisions regarding biomass selection and blending ratios to achieve high-quality pellet products.

2. Materials and Methods

2.1. Systematic Literature Review

The bibliographical search employed Boolean operators and a symbolic logic system developed around conceptual relationships and core keywords relevant to the study. This systematic approach allows for a comprehensive analysis of pertinent studies within a subject area. This methodology has been widely used in various scientific domains for presenting research data and conducting reviews [87,88,89]. Searches were conducted across the Scopus, Google Scholar, and DOAJ databases, utilizing a Boolean equation developed from keywords and phrases identified during preliminary literature surveys prior to the SLR. Publications spanning 2000 to 2021 within the “Biomass and Bioenergy” domain were considered and only English-language publications were included for analysis.

2.2. Data Extraction

Data extraction was performed with adherence to a standardized protocol within the framework of the SLR. After the selection of the studies, they were listed according to the title, author(s), and the published year. To manage, annotate, and categorize the findings, Zotero version 5.0.96.3 was employed, with duplicates excluded based on title and author congruence. Since the test results of the pellet quality parameters needed to be supplemented with standard deviation for use in the meta-analysis, articles without full-paper access were excluded from the final compilation. Subsequently, the refinement of article selection was achieved via meticulous scrutiny of each title and abstract. The exclusion criteria encompassed:
  • Studies published in languages other than English.
  • Investigations centered on biofuel liquid products (e.g., biodiesel, bio-oil).
  • Research focused on solid materials such as soil manure and compost, excluding pellets and briquettes.
  • Research inquiries into activated carbon, biochar, and similar substances.
  • Exploration of biogas production from solid waste materials originating from digestion processes.
  • Analysis of waste incineration practices.
After the exclusions, the final selection of suitable studies for meta-analysis was carried out by reading each study. Important and useful data were listed in a Microsoft Excel spreadsheet for easy reference.

2.3. Statistical Analysis

The meta-analysis was conducted for studies on pellets made from non-woody biomass blends and non-woody/woody biomass blends. In cases where test results for various raw materials and combinations were presented within a single study, the following procedure was adhered to:
  • For biomass blended pellets produced using the same raw materials but different combinations, the quality parameters of the most optimal pellet (chosen based on the author’s recommendations) were selected for the analysis.
  • For blended biomass pellets manufactured using diverse raw materials and combinations, the quality parameters of all pellet types were chosen for analysis.
  • For pellets produced using different raw materials but the same combinations, quality parameters of all pellet types were selected for analysis.
Following this, meta-analysis was executed using the statistical software R Studio (Version 2021.09.1+372 “Ghost Orchid” Release for Windows, Mozilla/5.0).
A flowchart of the review process followed in this study is depicted in Figure 1. The initial search yielded 1002 potentially relevant hits. Out of these, 719 studies were excluded during the first round, and an additional 97 articles were removed based on the criteria outlined at the data extraction stage. From the remaining 186 articles concerning biomass pellet production, the final data extraction was performed.
From the 186 selected studies, a set of 19 research studies containing test results for pellet quality parameters was extracted for meta-analysis. Although the remaining 167 studies were excluded from the meta-analysis, they were utilized as literature sources to gather relevant information and factual data for composing the SLR. To conduct the meta-analysis, parameters such as the number of test cycles (n), the mean of the test results (mean), and the standard deviation of the results (SD) were required. Studies lacking n, mean, and SD were excluded from the meta-analysis. The meta-analyzed results were presented in forest plots, and a comparison of each parameter with pellet standards was performed.
Both commercial pelleting machines and single-unit pelletizers [90] have been used to produce the tested pellets and obtain the test results that were meta-analyzed in this article. Although different test methods have been used, the final units of the tested results were consistent with the ISO 17225 standard series. The popular standard methods that have been employed include the International Organization for Standardization (ISO) standards, American Society for Testing and Materials (ASTM) standards, German Institute for Standardization (DIN) standards, European Nations (EN) standards, Brazilian National standards (NBR), the National Renewable Energy Laboratory Standard Scenarios (NREL), and technical specifications like those of the Environmental Protection Agency (EPA) (Table 2).
The properties of the pellets manufactured using 100% non-woody biomass materials and woody and non-woody biomass blends selected were listed to identify the special features and to compare them with ISO 17225-6 standard [83] categories A and B. The structural properties of the pellets such as diameter (D) and length (L); energy properties such as moisture content (MC), ash content (A), gross calorific value (GCV), volatile matter content (VM), and fixed carbon content (FC); mechanical properties such as bulk density (BD), mechanical durability (Du), fine particle content (F), and hardness; and chemical properties like nitrogen (N) and sulfur content (S) were analyzed. The main purpose of this was to identify the unique properties of pellets produced using single biomass and to develop a comparison of these property values with pellets that are manufactured using biomass blends. Some test values without sufficient data on the test method or the process were excluded from the study (chlorine content).

3. Results and Discussion

3.1. Pellets Analyzed

A summary of the 19 studies from different geographic regions of the world is presented in Table 3. Most of the selected studies were conducted in the European region, where there is a higher demand for pellets. The summary provides insights into pellets produced from biomass materials and the properties of the biomass types employed for production, as detailed in Table 3. However, among the selected studies, some report the properties of pellets produced and tested from both 100% single biomass and biomass blends. Mechanical pellet testing processes were employed across all studies. Some studies [13,90,92,93,94,95,98,100,101,104,106] also examined and reported the pellet die pressure and temperature during the pelletizing process, with these values also documented in Table 3.
When considering blended biomass pellets (non-woody and woody biomass blended pellets), the properties of 100% non-woody pellets produced in the aforementioned 19 studies were collected. Out of these 19 studies, 8 contained test results for both 100% non-woody biomass pellets and blended biomass pellets, another 8 studies contained test results for 100% non-woody biomass pellets, and an additional 3 studies contained test results exclusively for blended biomass pellets. The selection of blended biomass pellets for meta-analysis followed the method outlined in the statistical analysis section, and the chosen blended biomass pellet types are listed in Table 3 to facilitate a fair comparison with pellets produced from 100% non-woody biomass materials. In certain studies, pre-heating of the pellet die prior to the pellet production process was noted [98,103]. The rationale behind this pellet die heating is discussed in the literature, where an increase in biomass feedstock temperature reduces the friction in the press channel of the pellet mill [53] and decreases the energy required for the pelleting process [108], leading to decreased friction with a higher die temperature [53]. Nonetheless, the pelletizing temperature varied within the analyzed studies, ranging from 70 °C to 105 °C. The pressure levels to which biomass is subjected during pelletization influence product density and durability [53]. Thus, the available pressure levels applied during pellet production are also documented in Table 3. Most studies applied the same pressure levels [92,93], although higher pressure values have also been observed [90].
Table 3 presents the properties of biomass pellets produced using single biomass materials as well as biomass blends. Repetition of the same biomass materials across different studies has been observed. The properties not compliant with ISO 17225-6 standard category A are highlighted in bold in Table 3, and those failing to meet both categories (A and B) are highlighted in gray. While the other parameters fall within the ISO 17225-6 standard range, ash content, fine particle content, net calorific value, bulk density, and mechanical durability exhibit values slightly above the standards. Notably, the nitrogen content in pellets derived from coffee industry residues surpasses the permitted levels (Table 3).
Another intriguing observation pertains to pellets produced using the same raw materials across different studies, demonstrating varying pellet properties. For instance, the literature [93,98,100,102] has discussed the properties of pellets produced from sugarcane bagasse (SB). Despite utilizing the same biomass material, properties such as the ash content, fine particle content, mechanical durability, and bulk density exhibit considerable variability between studies. Similar patterns can be discerned in other studies, wherein authors attribute the variations in biomass properties to differences in sourcing regions. Some pellet production studies involve raw material heating processing or steam explosion before pellet production [98,109,110].
The literature [106] also describes the properties of pellets derived from torrefied wheat straw (Table 2). It is clear that the torrefaction process has resulted in an improvement in as well as a decrease in certain pellet parameters. Torrefaction serves as a pretreatment for upgrading woody biomass primarily for energy production [64,111]. This thermochemical treatment subjects biomass to heat within a reaction temperature range of 200 to 300 °C, employing an inert medium like nitrogen for a specified period, often ranging from seconds to an hour, depending on the particle size (PS). The research indicates that torrefaction techniques can enhance the material properties of raw straw, including a higher heating value (HHV) or gross calorific value (GCV), hydrophobicity, grinding ability, and an improved mass yield and energy density ratio [63,106,112].
Conventional torrefaction conditions and elevated temperatures result in torrefied biomass with an elevated ash content and substantial mass loss [63,113]. This phenomenon is also evident in the torrefied pellets discussed by [106], where the ash content of the pellets increased post-torrefaction (Table 3). The inherent binding properties of lignocellulosic biomass can be fortified via structural modifications of the cellulose–hemicellulose–lignin matrix using torrefaction methods [63]. However, excessive torrefaction and over-modification can amplify the compression and compaction characteristics of the raw materials, leading to increased energy consumption during pelleting due to the heightened friction in the press channel. In certain scenarios, this can also result in reduced pellet density [62,112].
In summary (Table 4), the literature indicates a significant correlation between the pelletization pressure, temperature, bulk density, and mechanical durability values of the resultant products [41,53,114]. The quality of blended biomass pellets improves as the percentage of woody materials in the blend increases. This suggests that incorporating a higher percentage of woody materials enhances the quality of the pellets [90,92,99].

3.2. Pellet Quality Parameters

This section presents forest plot diagrams illustrating the meta-analyzed pellet properties, each compared with the standard parameters. Additionally, the findings from the reviewed literature are incorporated.

3.2.1. Pellet Dimensions

Dimensions are crucial parameters for any solid fuel as they influence feeding and furnace technologies, impacting the fuel conveyance and combustion behavior. The pellet diameter is determined by selecting a die with appropriately sized die holes. The meta-analysis revealed a diameter of approximately 6 mm, with minimal variations. Notably, the EN ISO 17225 standard series specifies a maximum pellet length of 40 mm, a significant consideration for pneumatic feeding systems. Overly lengthy pellets can obstruct the feeding system, potentially causing system standstills [80].
In the meta-analysis, the mean pellet diameter was found to be 6.37 mm, with a corresponding mean length of 15.65 mm (Figure 2 and Figure 3). These mean values comply with the stipulated diameter and length requirements outlined in the pellet standards for both domestic and industrial purposes. An intriguing observation is that the pellet diameters manifest higher values than the die holes from which they are produced. The existing literature posits several possible reasons for this phenomenon. Scatolino et al., 2017 [102] proposes that the change in pellet size may be attributed to water filling the voids within the pellets, disrupting the bonds formed during the pelletization process. Additionally, this could be the result of particle rearrangement upon the release of compression force, compounded by the varying moisture content of the raw materials [102,115]. Furthermore, the literature suggests that pellets with smaller diameters exhibit a more uniform combustion rate compared to larger diameter pellets, attributed to the greater exposed surface area, facilitating efficient combustion [101].
European standards mandate that pellet length should not exceed four times the die diameter, promoting uniformity to facilitate smooth material flow during combustion or gasification processes [115]. The length-to-diameter ratio, termed the aspect ratio, is pivotal for pellet durability, which is called “green strength,” especially in pneumatic feeding processes [41,116,117] and the production of blockage-free pellet mills [101,118]. Additionally, the aspect ratio serves as a metric for compression during palletization [111]. An increased pelletizing pressure lengthens the pellet, while a larger pellet diameter reduces the pelletizing pressure. The aspect ratio directly influences pellet durability, with a higher ratio enhancing the durability, possibly due to enhanced particle bonding and lower hygroscopicity [76,95].
Notably, pellets comprising multiple biomass materials tend to have smaller diameters [93,102], a trend evident in this analysis. This reduction in diametric variation in mixed pellets is beneficial for the efficient design of pellet burners and feeders [102,119].

3.2.2. Moisture Content (MC)

Moisture influences the binding properties, density, storage conditions, and combustion characteristics of the pellets [93,120,121,122]. The MC is expressed as a percentage of the total weight of wood pellets for energy, and the ISO 17225-6 standard delineates two distinct MC limits: 12% for category A (Table 3) and 15% for category B (Table 3) for industrial non-woody biomass pellets. Moisture acts as a binding agent, strengthening biomass pellets via inter-particle van der Waals’ forces and/or hydrogen bonding [41,120]. However, excessive moisture can impede particle compaction and thereby reduce the pellet quality [99,108]. Moreover, a higher moisture content necessitates elevated energy consumption during the drying process, a significant cost factor in pellet production [98].
Insufficient moisture content in raw materials poses challenges during pelletization by elevating the friction forces in compression zones, resulting in increased electricity consumption and repair costs for pelleting machines [102]. Determining an exact moisture content (MC) range for ideal biomass pellets is complex, as it varies based on the granulometry, raw material properties, and process conditions [95,123,124]. Post-pelletization, the pellet moisture decreases by approximately 1 to 6% compared to the fed biomass’s MC [117,125]. In addition, torrefaction can also lead to a reduction in the moisture content of the produced pellets [106]. Maintaining lower MC values is essential to prevent incomplete biomass burning and fly ash formation [97,99,103]. Optimal MC ranges are 8–12% for 100% woody pellets and up to 15% for non-woody pellets to achieve an efficient burner performance [82,126].
The meta-analysis reveals lower moisture content (MC) values in blended biomass pellets, averaging 7.48% (Figure 4). The final MC depends on the raw material types and their blending ratios in the pellet production. For instance, a blend of 65% soybean waste and 35% cotton waste yields pellets with a 9.88% MC, whereas substituting cotton waste with sorghum waste reduces the MC to 4.46% [95]. MC exhibits a positive correlation with sawdust content but a negative relationship with herbaceous biomass materials [90,100,102], showcasing its intricate dynamics in blended biomass pellets. Notably, the MC significantly influences calorific value, combustion behavior, and efficiency, and fosters mold growth [116,126].

3.2.3. Ash Content

The ash content in wood pellets serves as a vital parameter, especially for heating applications, providing manufacturers and quality checkers with a rapid initial quality assessment [107,127]. This metric is representative of the non-combustible residue left at the end of the combustion cycle, originating from the inorganic elements present in the biomass. Elevated ash levels directly impact the combustion efficiency [93,125,128], resulting in increased maintenance costs for burner systems like boilers, furnaces, stoves, and gasifiers [129,130,131,132]. Notably, biomass with a high ash content can erode the pellet die, affecting pellet-binding mechanisms [122].
While the standards permit biomass pellet ash content up to 10%, an extensive proportion of the literature emphasizes that the ash levels should ideally stay within 4% for agro-residues in biomass briquetting and pelleting [133,134]. This range ensures enhanced combustion efficiency [97,135]. Notably, pellets derived from biomass blends exhibit a lower ash content and higher calorific values compared to single biomass pellets [93].
Another important fact is that the ash content of the torrefied wheat straw pellet is higher than the non-torrefied wheat straw pellet. Although torrefied pellets remained within the standard range, these kinds of increments in ash content are quite common scenarios with other torrefied biomass pellets as well [136,137,138,139]. It has been identified that the ash content gradually increases with the increasing severity of torrefaction [137]. The literature suggests that the high alkali metal contents (Na, K, Ca, Mg, Si, Cl, and S) [139,140] and the loss of organic matter during torrefaction are the reasons for the increased ash content of the torrefied biomass pellets [137].
The majority of the analyzed pellets fall within the specified ash content limits for industrial-scale non-woody pellets (Figure 5). However, pellets containing soybean waste demonstrate significantly higher ash levels, attributed by the authors to potential soil contamination during storage [102].
With the exception of soybean blended pellets (ash content: 14.03%, 15.02%) and oil palm residue blended pellets (ash content: 11.90%), all analyzed studies remained within the prescribed maximum limits for industrial-scale non-woody biomass standards (Figure 5). Notably, pellets derived from a mix of sawdust and waste from the coffee industry surpassed the upper limit of ISO 17225-6 category A, potentially due to soil contamination. However, a significant portion of the analyzed blended pellets adhered to both the upper and lower standards compared to pellets produced from single biomass materials.

3.2.4. Mechanical Durability

Durability, a measure of a pellet’s resistance to abrasion, is an important parameter in ensuring smooth fuel-feeding systems and reducing dust emissions during handling. Pellets with low durability can lead to storage and transportation challenges, along with health and environmental concerns, due to the susceptibility to disintegration caused by moisture adsorption, falls, or friction [105]. In light of the meta-analysis results and existing literature, maintaining the mechanical durability of both single and blended pellets is of paramount importance. For instance, single biomass pellets produced from soybean waste exhibited a low mechanical durability value of 18% (Table 3). However, when combined with sawdust, this value significantly improved to 67.42%. A similar trend is observed in pellets made from other non-woody biomass materials [102,141].
Most of the analyzed pellets exhibit mechanical durability values well within the ISO 17225-6 standards, with an extracted mean value of 89.03% (Figure 6). The literature has studied what causes reduced mechanical durability values. Recent research shows that pellet durability depends on several important factors. These factors include the material used (such as starch, protein, fiber, fat, lignin, extractives, moisture content, and particle size and distribution) [76,93,95,99,142], the processes used before pellet formation (such as steam conditioning, preheating, and adding binders) [41,114,143,144,145], and the equipment used for pellet formation (such as forming pressure, pellet mill, and roll press variables) [76,111,146]. These factors can affect the strength and durability of the pellets. Additionally, how the pellets are treated after production, such as via cooling, drying, or storage in high humidity conditions, can also affect their strength and durability [41,147].
Woody biomass, distinguished by its higher cellulose content than non-woody (agricultural) biomass, accounts for the diminished mechanical durability of pellets produced from agricultural residues, challenging the commercial viability of these solid biofuels [148]. To bolster mechanical stability, contemporary research suggests increasing the proportion of sawdust over non-woody biomass in biomass blends [76,90,115,149].
According to Lavergne et al. (2021) [150], moisture plays a dual role in enhancing particle bonding while also reducing the friction within the die due to its lubricating properties. However, excessive moisture levels can result in the formation of a layer that is too large to effectively hold the particles together [41]. Once an optimal moisture content is achieved, the moisture content enhances not only the durability but also the physical characteristics and overall product quality. Studies have indicated that a higher moisture content diminishes pellet durability [151]. It is recommended that the moisture content falls within the range of 10–15% for optimal results [152].
Furthermore, the size of the particles has a direct effect on the mechanical strength of the biomass pellets. Studies have shown that using smaller particles improves the density, leading to higher yield stress and ultimately resulting in stronger pellets compared to those produced with larger particles [153,154,155]. It is generally recommended to use particles with a diameter below 5 mm for optimal pellet quality [53,102]. While it is beneficial to have a wide range of particle sizes for improved pellet quality, an excessive amount of fine particles (less than 0.5 mm in diameter) in the raw material negatively influences friction and the overall quality of the pellets [53,156].
Researchers have revealed how the equipment parameters affect the strength and durability of the pellets. Increasing the length of the press channel creates more friction between the biomass particles and the channel walls. This also increases the amount of time for which the material is exposed to heat and pressure [157]. The temperature of the die during pelletizing is mainly influenced by friction [153,157,158,159,160,161]. The die temperature, along with the composition of the feedstock and moisture content, affects the softening of the natural binding agents in the biomass (such as lignin and mono sugars). This leads to the enhancement of the pellet strength and durability and is therefore a major factor in pellet quality [162]. It has been observed that using small die holes with a high die width (a higher L/D ratio) can improve the pellet durability [111,114,146]. However, excessively long die lengths increase friction without providing any additional improvement in pellet durability.
Torrefaction is a common pre-production technique that directly affects the quality of the pellets. Despite its benefits in improving pellet density by reducing the moisture content, enhancing the grinding properties, and altering the chemical compositions [144], it has been observed that torrefied materials have diminished durability properties [136,138,163]. This finding is supported by the analyzed data, as demonstrated by the significantly lower mechanical durability values in torrefied wheat straw pellets compared to non-torrefied pellets (Table 3).
The research and analysis conducted indicated that freezing pellets during storage or transport only has a slight impact on their mechanical durability. If the pellets have a high initial mechanical durability in normal conditions, freezing and subsequent defrosting do not significantly affect their mechanical durability. However, if the pellets have low quality and low initial mechanical durability, their mechanical durability may further decrease if they were previously frozen. Therefore, companies and consumers involved in pellet storage need to consider their mechanical durability index. To minimize the risk of deterioration, it is recommended to prevent freezing pellets by storing them at temperatures above 0 degrees Celsius. However, proper storage practices play a crucial role in achieving a high mechanical durability index (DU > 97.5%) [147].

3.2.5. Fine Particle Content

Excessive fines in pellets can pose challenges during combustion due to the risk of elevated temperatures caused by the rapid burn rate of fine particles compared to pellets [164]. Notably, bagged pellets generally contain fewer fines compared to those delivered in bulk, while pellets stored in silos may exhibit a higher fine particle content upon delivery [165]. The amount of fine particles in pellets is determined by both the ISO 17225-2 and ISO 17225-6 standards. Interestingly, ISO 17225-2 allows for a fine particle content of up to 6% in woody pellets, while ISO 17225-6 requires the fine particle content to be below 3% for biomass pellets. It is important to note that the reasons for this difference in the maximum limits are not documented. However, one possible explanation is that many agricultural biomass materials are more explosive and pose a higher risk of ignition compared to woody biomass. This is due to their lower ignition temperature and lower ignition energy. Another reason could be the mechanical durability (Du) limit, where very good-quality pellets must have a Du higher than 97.5%. This limit helps restrict the presence of loose particles to 3%.
In the meta-analysis of the studied pellets, the fine particle content values consistently fell below the ISO 17225-6 standards, yielding an average value of 1.03%. While soybean waste pellets and sugarcane bagasse pellets individually exhibited higher fine particle content values (3.32% and 44.00%, respectively), the blending of these materials resulted in a lower fine particle content of 2.50%. This observation suggests that the incorporation of multiple biomass materials into pellet production tends to reduce the fine particle content in comparison to pellets derived from a single biomass source (Figure 7).

3.2.6. Calorific Value

The calorific value of wood pellets is intrinsically tied to the type of raw materials utilized, influenced by the presence of organic components like lignin, cellulose, starch, protein, and fat, which vary across plant species and plant parts. The gross calorific value (GCV) measures the heat generated by burning a unit volume, whereas the net calorific value (NCV) takes into consideration the gross calorific value minus the latent heat in the water caused by hydrogen combustion above an atmospheric temperature. For domestic use, pellets require higher standards of calorific value compared to their industrial counterparts (Table 3).
Recent literature findings underscore the pivotal role of carbon and hydrogen content in boosting pellets’ GCV, while ash (A), oxygen, moisture content (MC), and nitrogen content tend to have adverse effects. Blended pellets, incorporating a mix of biomass materials, consistently exhibit an enhanced GCV compared to single biomass pellets. This observation aligns with the previous literature, advocating the blending of agricultural residues with wood as a promising approach to energy production.
The thermal pretreatment process of biomass, such as torrefaction, has been proven to enhance the combustibility of biomass [166,167,168]. Consequently, biomass retains the majority of its energy components after undergoing torrefaction, resulting in a torrefied product that has a higher calorific value and increased energy density [122,139]. This process is also consistent with the study conducted by Azócar et al. in 2019 [106], which examined pellets made from torrefied wheat straw (Table 3).
Even though the standards provide a permissible limit for the NCV, some studies have examined both the NCV and GCV [92,93,95,102], while others have only examined the GCV/HHV values of the developed pellets [97,99,100,107]. Therefore, the GCV values were used for the meta-analysis, as stated in all the studies.
To address the misinterpretation that may arise when comparing the meta-analyzed values with the permissible limits of the standards, an assumption was made based on studies that present both the GCV and NCV values.
The difference between the GCV and NCV values was calculated for each study [92,93,95,102], and an average value of 2.55 MJ/kg was obtained. It was assumed that this difference represents the calorific value of the water vapor resulting from the combustion of hydrogen and the vaporization of the original water, which is then condensed into a liquid state.
The GCV of the pellets that were analyzed ranged from 17.03 MJ/kg to 18.54 MJ/kg, with an average of 17.79 MJ/kg. Based on the above assumption, the NCV can be calculated as 15.24 MJ/kg, which is higher than the lowest NCV of 14.5 MJ/kg mentioned in the ISO standard (Table 3). Therefore, it can be concluded that the meta-analysis results of the NCV suggest that biomass blends are a favorable source for industrial usage in terms of the NCV (Figure 8).
When the sawdust percentage increases in the biomass blends, an increment in the GCV can be seen [102,161]. In other words, the heating value of a material is linearly related to its lignin content [169]. Further, Serrano et al., 2011 [125] described that pellets produced from mixtures of herbaceous and woody materials with a low ash content and high lignin content will help to improve the heating value of the pellets produced [93].

3.2.7. Bulk Density

The bulk density of pellets is determined by the particle density and overall bulk porosity. Elevating the pellet bulk density not only heightens the energy density within the compressed pellets but also streamlines transport and storage, simplifying logistics. The ISO standards for pellets mandate a minimum bulk density of 600 kg/m3. Concurrently, blended biomass pellets, typically comprising various lignocellulosic materials, consistently exhibit elevated bulk density values. This phenomenon is notably associated with the chemical composition of biomass mixtures, specifically the presence of cellulose, hemicellulose, lignin, and extracts. An increase in cellulose and lignin content augments the bulk density, as both these components serve as natural binders. The analyzed blended biomass pellets, with a few exceptions, demonstrate higher bulk density values than their single biomass counterparts. Most of the studied bulk density values fall within the range of 600–700 kg/m3, with a mean bulk density of 666.52 kg/m3 extracted from the meta-analysis (Figure 9).

3.2.8. Elemental Analysis of Pellets

The ISO 17225 standard series focuses on various elements present in biomass pellets, including nitrogen (N), sulfur (S), chlorine (Cl), and heavy metals such as arsenic (As), cadmium (Cd), and chromium (Cr), among others. Nevertheless, the ISO 17225-6 standards establish notably elevated permissible thresholds for nitrogen, sulfur, and chlorine, surpassing the ISO 17225-2 guidelines. According to the literature, it is possible to strategically control these elemental contents by carefully selecting biomass materials and their mixing compositions [92,100].

Nitrogen

Nitrogen is permitted at 1.5% in ISO 17225-6 category A, while category B allows for 2.0%. Although the meta-analysis reveals a mean nitrogen content of 1.49%, most analyzed studies report nitrogen levels exceeding these standards (Figure 10). Notably, agricultural residues like sugarcane bagasse, sorghum, coffee residues, soybean residues, and elephant grass naturally contain an elevated nitrogen content. Such disparities in nitrogen levels can be attributed to the use of fertilizers during crop growth. The nitrogen content of residual biomasses from coffee production, specifically stem bark and leaves, was reported to be 2.13% and 3.54%, respectively. In conclusion, the intrinsic properties of the biomass significantly influence the pellet quality, with blended biomass pellets generally exhibiting a lower nitrogen content than their single biomass counterparts.

Sulfur

In the ISO 17225-6 standards, the allowable sulfur content for non-woody biomass pellets stands at 0.20% for category A and 0.30% for category B. While these limits may appear relatively low, they are essential for mitigating sulfur emissions during biomass combustion, in accordance with stringent emission policies. All the examined blended biomass pellets conform to the prescribed standards (Figure 11). Notably, only pellets derived from soybean waste and cotton waste (65SyB and 35CW) exhibited sulfur content levels surpassing those of ISO 17225-6 category A, with a mean sulfur content of 0.08%. This pattern of adhering to the standard sulfur limits holds for both blended biomass and single biomass pellets.

Chlorine

In compliance with the ISO 17225-6 standards, the allowable chlorine content in biomass pellets is significantly restricted (0.1% in category A and 0.3% in category B). It is noteworthy that although biomass generally exhibits lower nitrogen, sulfur, and chlorine contents, these elements can lead to combustion chamber corrosion and the emission of greenhouse gases, akin to the effects of ash and fines [42,170,171]. In specific agricultural biomasses, reed canary grass and switchgrass demonstrated elevated silicon levels, while timothy hay displayed a heightened potassium content. However, when reed canary grass was blended with woody biomass, the silicon content decreased by nearly half compared to the individual reed canary grass [96]. Alkaline minerals such as calcium, phosphorous, chloride, and potassium, although not significantly impacting pelletization, can result in severe issues like corrosion, slagging, and fouling within thermal systems during combustion formation [100,161,172,173,174].
However, the test parameters for chloride could not be found in the analyzed literature. This could be due to the calculation of ash and fine particle contents separately. Nevertheless, these findings indicate that biomass residues can be harnessed in pellet production with further research and development [96].

3.2.9. Volatile Matter Content

The volatile matter content in biomass significantly impacts its reactivity, contributing to increased frictional heat. Biomass solid fuel, possessing a higher volatile matter content, exhibits a swifter combustion rate during devolatilization, rendering it easier to ignite and burn effectively [97]. Research by Kataki and Konwer, 2002 [175] reinforces that pellets with over 30% volatile matter yield greater heat and facilitate quicker and more efficient combustion. It is noteworthy that agro-pellets display a variable volatile matter content, ranging from 9.40% to 89.31% [99] and 79.58% to 85.11% [176]. The analysis of these studies reveals a mean volatile matter content of 80.55%, aligning closely with the reported ranges of 76.99% to 84.11% (Figure 12). Similar to other properties, blended biomass pellets exhibit a higher volatile matter content compared to the raw biomass before compaction [93].

3.2.10. Fixed Carbon

The inherent low fixed carbon content of biomass renders it an exceptional and highly reactive fuel. This attribute facilitates rapid combustion, as biomass with a low carbon content tends to burn more swiftly [97,135]. One study [93] substantiates this observation by noting a reduction in fixed carbon content following the pelletizing process. The meta-analysis conducted on various samples yielded a mean fixed carbon content value of 12.98% (Figure 13).

4. Conclusions

The findings of this study suggest that combining non-woody biomass (such as agricultural residues) with sawdust can provide a viable and environmentally friendly method for generating energy. Additionally, the study indicates that using a blend of non-woody and woody biomass in pellet production results in higher-quality pellets.
Furthermore, various agricultural residues such as palm oil industry waste, sugarcane industry waste, rice industry waste, soybean industry waste, and other environmentally generated materials like water hyacinth can be explored as raw materials for biomass pellet production. This approach has the potential to provide a lasting solution for waste management and promote the development of rural livelihoods, notably in developing nations.
Challenges persist in processing lignocellulosic biomass waste materials, specifically in optimizing the compositions of various biomass types. If these issues can be addressed via continued research and development, biomass pellets derived from unconventional materials have the potential to compete with wood pellets as a sustainable heat generation source. Notably, studies have identified nitrogen and ash content as the most restrictive parameters for achieving high-quality pellets.
After examining the researched studies and evaluating the parameters, it becomes clear that pellets made from a mixture of woody and non-woody materials are considered to be of intermediate quality between 100% woody and non-woody biomass pellets. They boast higher heating values, mechanical durability, hardness, and reduced fine particle and ash contents in comparison to pellets derived from single non-woody biomass materials. Blended biomass pellets are also known to possess a heightened ability to release more energy per unit volume during the combustion process, resulting in reduced particle emissions, CO emissions, and minimized slag formation. These superior quality parameters are guaranteed since the mixed pellets contain a portion of woody biomass that possesses higher-quality parameters.
However, certain gaps still exist within this field. The temperature and pressure applied during the pelletization process are essential factors that require further exploration. Although the present analysis does not provide substantial evidence for the relationship between pelletizing temperature and pressure and their involvement in mechanical durability and bulk density, it is clear that elevated pelletizing temperatures lead to lower pressure requirements.
In summary, the meta-analysis results provide mean values for various parameters including diameter (6.37 mm), length (15.65 mm), moisture content (7.48%), ash content (4.19%), fine particle content (1.03%), gross calorific value (17.79 MJ/kg), bulk density (666.52 kg/m3), nitrogen content (1.49%), sulfur content (0.08%), volatile matter content (80.55%), and fixed carbon content (12.98%). All these mean values align with the requirements of both EN ISO 17225-6 standard categories A and B. However, mechanical durability is the exception, showing a lower value than the standards dictate (89.03%). This deviation is attributed to the lower mechanical durability values of pellets produced from non-woody biomass mixtures.
Based on the findings of the examined research, it is advisable to consider the utilization of a composition range of 0–50% non-woody biomass (varying according to the specific material type) in conjunction with woody biomass to produce mixed pellets. This process should be conducted at a pelleting temperature exceeding 80 °C, alongside a pelleting pressure of approximately 29.42, as it yields superior quality parameters.
Moreover, the study highlights that the bulk density and mechanical durability values are generally lower in blended biomass pellets, which can vary by the raw material types and blending compositions.
Consequently, conducting thorough test runs and adhering to quality standards is crucial when introducing any new biomass or agricultural residues as raw ingredients into pellet production. Potential methods for enhancing product quality and efficiency include machinery modification, steam explosion, torrefaction, and other processes. The successful development of these waste-to-energy products stands to positively impact global energy needs and industrial sustainability.
In light of the current analysis, it is recommended to:
  • Optimize biomass compositions: Selecting the right blend of woody and non-woody materials and their proportions ensures high-quality biomass pellets.
  • Machinery and process modifications: Innovative processes such as steam explosion and torrefaction shall be considered to improve the pellet quality.
  • Encourage agricultural residue utilization: Exploring the use of agricultural residue as raw materials can contribute to effective waste management and the development of sustainable energy solutions.
  • Prioritize research into compositional effects: Further investigation into the influence of pelletizing temperature and pressure on pellet properties is warranted.
  • Manage elemental content: Careful selection of biomass materials and blending compositions can effectively manage the nitrogen and ash content, resulting in higher-quality pellets.
  • Consider woody materials: To enhance the pellet quality, incorporating a higher percentage of woody materials into biomass blends is recommended.
This meta-analysis analysis provides valuable insights into the improvement of biomass pellet quality and sustainability, which holds great promise in addressing global energy challenges.

Author Contributions

Conceptualization, P.P.; methodology, P.P., R.B. and R.L.R.; software, R.L.R.; validation, R.L.R.; formal analysis, P.P. and R.L.R.; investigation, P.P. and R.L.R.; resources, P.P. and R.L.R.; data curation, R.L.R.; writing—original draft preparation, R.L.R.; writing—review and editing, P.P., R.B. and R.V.; visualization, R.L.R.; supervision, P.P., R.V., R.B. and H.A.; project administration, P.P.; funding acquisition, P.P. and R.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the University of Sri Jayewardenepura: Research Grant ASP/01/RE/SCI/2021/27. The APC was funded by the Louisiana State University, Baton Rouge, LA, USA.

Data Availability Statement

Data are contained within the article.

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. Ng, A.W.; Nathwani, J.; Fu, J.; Zhou, H. Green Financing for Global Energy Sustainability: Prospecting Transformational Adaptation beyond Industry 4.0. Sustain. Sci. Pract. Policy 2021, 17, 377–390. [Google Scholar] [CrossRef]
  2. Omer, A.M. Green Energies and the Environment. Renew. Sustain. Energy Rev. 2008, 12, 1789–1821. [Google Scholar] [CrossRef]
  3. Shakeel, S.R.; Takala, J.; Shakeel, W. Renewable Energy Sources in Power Generation in Pakistan. Renew. Sustain. Energy Rev. 2016, 64, 421–434. [Google Scholar] [CrossRef]
  4. Bilan, Y.; Kozmenko, S.; Makarenko, I. Recent Advances in the Energy Market Development: Current Challenges and Perspectives of Energy Crises in Academia. Energies 2023, 16, 2332. [Google Scholar] [CrossRef]
  5. Zakeri, B.; Paulavets, K.; Barreto-Gomez, L.; Echeverri, L.G.; Pachauri, S.; Boza-Kiss, B.; Zimm, C.; Rogelj, J.; Creutzig, F.; Ürge-Vorsatz, D.; et al. Pandemic, War, and Global Energy Transitions. Energies 2022, 15, 6114. [Google Scholar] [CrossRef]
  6. Ozili, P.K.; Ozen, E. Global Energy Crisis: Impact on the Global Economy. In The Impact of Climate Change and Sustainability Standards on the Insurance Market; Sood, K., Grima, S., Young, P., Ozen, E., Balusamy, B., Eds.; Wiley: Hoboken, NJ, USA, 2023; pp. 439–454. ISBN 978-1-394-16651-0. [Google Scholar]
  7. Perera, P.; Perera, R.; Vlosky, R.P.; Darby, P. Potential of Using Poultry Litter as a Feedstock for Energy Production; Louisiana Forest Products Development Center, Louisiana State University Agricultural Center Baton Rouge: Baton Rouge, LA, USA, 2010. [Google Scholar]
  8. European Environment Agency. Share of Energy Consumption from Renewable Sources in Europe. Available online: https://www.eea.europa.eu/en/analysis/indicators/share-of-energy-consumption-from#:~:text=According%20to%20European%20Enviroment%20Agency,strong%20gro-wth%20in%20solar%20power.com (accessed on 23 November 2023).
  9. Mao, G.; Huang, N.; Chen, L.; Wang, H. Research on Biomass Energy and Environment from the Past to the Future: A Bibliometric Analysis. Sci. Total Environ. 2018, 635, 1081–1090. [Google Scholar] [CrossRef]
  10. Al-Kayiem, H.; Mohammad, S. Potential of Renewable Energy Resources with an Emphasis on Solar Power in Iraq: An Outlook. Resources 2019, 8, 42. [Google Scholar] [CrossRef]
  11. Bandara, W.A.R.T.W.; Ranasinghe, O.; Perera, P.; Vlosky, R.; Kizha, A.R. Potential to Use Invasive Plants in Biomass Energy Production: A Case Study Prosopis Juliflora in Coastal Wetlands of Sri Lanka. Trees For. People 2022, 10, 100330. [Google Scholar] [CrossRef]
  12. Duranay, N.D.; Akkuş, G. Solid Fuel Production with Torrefaction from Vineyard Pruning Waste. Biomass Convers. Biorefinery 2019, 11, 2335–2346. [Google Scholar] [CrossRef]
  13. Szyszlak-Bargłowicz, J.; Słowik, T.; Zając, G.; Blicharz-Kania, A.; Zdybel, B.; Andrejko, D.; Obidziński, S. Energy Parameters of Miscanthus Biomass Pellets Supplemented with Copra Meal in Terms of Energy Consumption during the Pressure Agglomeration Process. Energies 2021, 14, 4167. [Google Scholar] [CrossRef]
  14. Wattana, W.; Phetklung, S.; Jakaew, W.; Chumuthai, S.; Sriam, P.; Chanurai, N. Characterization of Mixed Biomass Pellet Made from Oil Palm and Para-Rubber Tree Residues. Energy Procedia 2017, 138, 1128–1133. [Google Scholar] [CrossRef]
  15. Bimbela, F.; Abrego, J.; Gonzalo, A.; Sanchez, J.L.; Arauzo, J. Biomass Pyrolysis Liquids. Fundamentals, Technologies and New Strategies. Boletín Grupo Español Carbón 2014, 33, 11–14. [Google Scholar]
  16. Hammar, T.; Stendahl, J.; Sundberg, C.; Holmström, H.; Hansson, P.-A. Climate Impact and Energy Efficiency of Woody Bioenergy Systems from a Landscape Perspective. Biomass Bioenergy 2019, 120, 189–199. [Google Scholar] [CrossRef]
  17. Ioelovich, M. Recent Findings and the Energetic Potential of Plant Biomass as a Renewable Source of Biofuels–a Review. Bioresources 2015, 10, 1879–1914. [Google Scholar] [CrossRef]
  18. Lauri, P.; Havlík, P.; Kindermann, G.; Forsell, N.; Böttcher, H.; Obersteiner, M. Woody Biomass Energy Potential in 2050. Energy Policy 2014, 66, 19–31. [Google Scholar] [CrossRef]
  19. Mao, G.; Zou, H.; Chen, G.; Du, H.; Zuo, J. Past, Current and Future of Biomass Energy Research: A Bibliometric Analysis. Renew. Sustain. Energy Rev. 2015, 52, 1823–1833. [Google Scholar] [CrossRef]
  20. Mola-Yudego, B.; Arevalo, J.; Díaz-Yáñez, O.; Dimitriou, I.; Haapala, A.; Ferraz Filho, A.C.; Selkimäki, M.; Valbuena, R. Wood Biomass Potentials for Energy in Northern Europe: Forest or Plantations? Biomass Bioenergy 2017, 106, 95–103. [Google Scholar] [CrossRef]
  21. Welfle, A.; Gilbert, P.; Thornley, P.; Stephenson, A. Generating Low-Carbon Heat from Biomass: Life Cycle Assessment of Bioenergy Scenarios. J. Clean. Prod. 2017, 149, 448–460. [Google Scholar] [CrossRef]
  22. Shepel, O. Wood Pellets-As a Key Resource for Thermal Industry. In Proceedings of the VI International (XIX Regional) Scientific Conference on Technogenic Systems and Environmental Risk, Obninsk, Russia, 20–21 April 2017; p. 23. [Google Scholar]
  23. Situmorang, Y.A.; Zhao, Z.; Yoshida, A.; Abudula, A.; Guan, G. Small-Scale Biomass Gasification Systems for Power Generation (< 200 kW Class): A Review. Renew. Sustain. Energy Rev. 2020, 117, 109486. [Google Scholar]
  24. Cardoen, D.; Joshi, P.; Diels, L.; Sarma, P.M.; Pant, D. Agriculture Biomass in India: Part 1. Estimation and Characterization. Resour. Conserv. Recycl. 2015, 102, 39–48. [Google Scholar] [CrossRef]
  25. Tursi, A. A Review on Biomass: Importance, Chemistry, Classification, and Conversion. Biofuel Res. J. 2019, 6, 962–979. [Google Scholar] [CrossRef]
  26. Bioenergy Europe. Statistical Report—Bioenergy Landscape 2020. Available online: https://biom.cz/upload/9982d8381d3da848a8072e06cf96ec87/sr20_bioenergy-landscape_non_members.pdf (accessed on 20 November 2023).
  27. Bilandzija, N.; Voca, N.; Jelcic, B.; Jurisic, V.; Matin, A.; Grubor, M.; Kricka, T. Evaluation of Croatian Agricultural Solid Biomass Energy Potential. Renew. Sustain. Energy Rev. 2018, 93, 225–230. [Google Scholar] [CrossRef]
  28. Lamers, P.; Mai-Moulin, T.; Junginger, M. Challenges and Opportunities for International Trade in Forest Biomass. In Mobilisation of Forest Bioenergy in the Boreal and Temperate Biomes; Elsevier: Amsterdam, The Netherlands, 2016; pp. 127–164. ISBN 978-0-12-804514-5. [Google Scholar]
  29. Himandi, S.; Perera, P.; Amarasekera, H.; Rupasinghe, R.; Vlosky, R.P. Wood Residues in the Moratuwa Woodworking Industry Cluster of Sri Lanka: Potential for Sector Synergies and Value-Added Products. For. Prod. J. 2021, 71, 379–390. [Google Scholar] [CrossRef]
  30. Perera, P.; Rupasinghe, R.L.; Weerasekera, D.; Vlosky, R.; Bandara, R. Revisiting Forest Certification in Sri Lanka: The Forest Management and Export Wood-Based Manufacturing Sector Perspectives. Forests 2022, 13, 179. [Google Scholar] [CrossRef]
  31. Demirbas, M.F.; Balat, M.; Balat, H. Potential Contribution of Biomass to the Sustainable Energy Development. Energy Convers. Manag. 2009, 50, 1746–1760. [Google Scholar] [CrossRef]
  32. Irfan, M.; Elavarasan, R.M.; Ahmad, M.; Mohsin, M.; Dagar, V.; Hao, Y. Prioritizing and Overcoming Biomass Energy Barriers: Application of AHP and G-TOPSIS Approaches. Technol. Forecast. Soc. Change 2022, 177, 121524. [Google Scholar] [CrossRef]
  33. Karekezi, S.; Lata, K.; Coelho, S.T. Traditional Biomass Energy: Improving Its Use and Moving to Modern Energy Use. In Renewable Energy; Routledge: London, UK, 2012; pp. 258–289. [Google Scholar]
  34. Hoogwijk, M.; Faaij, A.; Eickhout, B.; Devries, B.; Turkenburg, W. Potential of Biomass Energy out to 2100, for Four IPCC SRES Land-Use Scenarios. Biomass Bioenergy 2005, 29, 225–257. [Google Scholar] [CrossRef]
  35. Klimowicz, G. Southeast Asia Set for Biomass Boom. Eco-Business. 2013. Available online: http://www.eco-business.com/news/southeast-asia-set-biomass-boom/ (accessed on 18 July 2023).
  36. Carlos, R.M.; Ba Khang, D. Characterization of Biomass Energy Projects in Southeast Asia. Biomass Bioenergy 2008, 32, 525–532. [Google Scholar] [CrossRef]
  37. Guo, M.; Song, W.; Buhain, J. Bioenergy and Biofuels: History, Status, and Perspective. Renew. Sustain. Energy Rev. 2015, 42, 712–725. [Google Scholar] [CrossRef]
  38. Haberl, H.; Erb, K.-H.; Krausmann, F.; Running, S.; Searchinger, T.D.; Kolby Smith, W. Bioenergy: How Much Can We Expect for 2050? Environ. Res. Lett. 2013, 8, 031004. [Google Scholar] [CrossRef]
  39. Strezov, V.; Evans, T.J. Biomass Processing Technologies; CRC Press: Boca Raton, FL, USA, 2014; ISBN 1-4665-6616-7. [Google Scholar]
  40. Gilvari, H.; Van Battum, C.H.H.; Farnish, R.; Pang, Y.; De Jong, W.; Schott, D.L. Fragmentation of Fuel Pellets during Transport via a Belt Conveyor: A Design of Experiment Study. Particuology 2022, 66, 29–37. [Google Scholar] [CrossRef]
  41. Kaliyan, N.; Vance Morey, R. Factors Affecting Strength and Durability of Densified Biomass Products. Biomass Bioenergy 2009, 33, 337–359. [Google Scholar] [CrossRef]
  42. Nunes, L.J.R.; Matias, J.C.O.; Catalão, J.P.S. Mixed Biomass Pellets for Thermal Energy Production: A Review of Combustion Models. Appl. Energy 2014, 127, 135–140. [Google Scholar] [CrossRef]
  43. Puig-Arnavat, M.; Shang, L.; Sárossy, Z.; Ahrenfeldt, J.; Henriksen, U.B. From a Single Pellet Press to a Bench Scale Pellet Mill —Pelletizing Six Different Biomass Feedstocks. Fuel Process. Technol. 2016, 142, 27–33. [Google Scholar] [CrossRef]
  44. Rentizelas, A.A.; Tolis, A.J.; Tatsiopoulos, I.P. Logistics Issues of Biomass: The Storage Problem and the Multi-Biomass Supply Chain. Renew. Sustain. Energy Rev. 2009, 13, 887–894. [Google Scholar] [CrossRef]
  45. Shojaeiarani, J.; Bajwa, D.S.; Bajwa, S.G. Properties of Densified Solid Biofuels in Relation to Chemical Composition, Moisture Content, and Bulk Density of the Biomass. BioResources 2019, 14, 4996–5015. [Google Scholar] [CrossRef]
  46. Jayawardhane, J.; Perera, P.K.P.; Lokupitiya, R.S.; Amarasekara, H.S.; Ruwanpathirana, N. The Effect of Quality Attributes in Determination of Price for Plantation-Grown Teak (Tectona Grandis) Logs in Sri Lanka. Ann. For. Res. 2015, 59, 1–12. [Google Scholar] [CrossRef]
  47. Perera, P.; Amarasekera, H.; Weerawardena, N.D.R. Effect of Growth Rate on Wood Specific Gravity of Three Alternative Timber Species in Sri Lanka; Swietenia Macrophylla, Khaya Senegalensis and Paulownia Fortunei. J. Trop. For. Environ. 2012, 2, 567. [Google Scholar] [CrossRef]
  48. Esen, M.; Yuksel, T. Experimental Evaluation of Using Various Renewable Energy Sources for Heating a Greenhouse. Energy Build. 2013, 65, 340–351. [Google Scholar] [CrossRef]
  49. Jha, S.K.; Puppala, H. Prospects of Renewable Energy Sources in India: Prioritization of Alternative Sources in Terms of Energy Index. Energy 2017, 127, 116–127. [Google Scholar] [CrossRef]
  50. Martinez, C.L.M.; Rocha, E.P.A.; Carneiro, A.d.C.O.; Gomes, F.J.B.; Batalha, L.A.R.; Vakkilainen, E.; Cardoso, M. Characterization of Residual Biomasses from the Coffee Production Chain and Assessment the Potential for Energy Purposes. Biomass Bioenergy 2019, 120, 68–76. [Google Scholar] [CrossRef]
  51. Oh, T.H.; Pang, S.Y.; Chua, S.C. Energy Policy and Alternative Energy in Malaysia: Issues and Challenges for Sustainable Growth. Renew. Sustain. Energy Rev. 2010, 14, 1241–1252. [Google Scholar] [CrossRef]
  52. Shuba, E.S.; Kifle, D. Microalgae to Biofuels: ‘Promising’Alternative and Renewable Energy, Review. Renew. Sustain. Energy Rev. 2018, 81, 743–755. [Google Scholar] [CrossRef]
  53. Stelte, W.; Sanadi, A.R.; Shang, L.; Holm, J.K.; Ahrenfeldt, J.; Henriksen, U.B. Recent Developments in Biomass Pelletization–A Review. BioResources 2012, 7, 4451–4490. [Google Scholar] [CrossRef]
  54. Baidya, R.; Bhattacharya, T.; Kumar, G.; Ghosh, S.K. Energy Analysis of Water Hyacinth–Cow Dung–Sawdust Mixture Briquettes—An Indian Perspective. In Waste Management and Resource Efficiency; Springer: Berlin/Heidelberg, Germany, 2018; pp. 719–725. [Google Scholar]
  55. Johnston, C.M.T.; Guo, J.; Prestemon, J.P. U.S. and Global Wood Energy Outlook under Alternative Shared Socioeconomic Pathways. Forests 2022, 13, 786. [Google Scholar] [CrossRef]
  56. Junginger, H.M.; Mai-Moulin, T.; Daioglou, V.; Fritsche, U.; Guisson, R.; Hennig, C.; Thrän, D.; Heinimö, J.; Hess, J.R.; Lamers, P.; et al. The Future of Biomass and Bioenergy Deployment and Trade: A Synthesis of 15 Years IEA Bioenergy Task 40 on Sustainable Bioenergy Trade: The Future of Biomass and Bioenergy Deployment and Trade: A Synthesis of 15 Years IEA Bioenergy Task 40 on Sustainable Bioenergy Trade. Biofuels Bioprod. Bioref. 2019, 13, 247–266. [Google Scholar] [CrossRef]
  57. Jakob, M.; Steckel, J.C. How Climate Change Mitigation Could Harm Development in Poor Countries. Wiley Interdiscip. Rev. Clim. Chang. 2014, 5, 161–168. [Google Scholar] [CrossRef]
  58. Patronen, J.; Kaura, E.; Torvestad, C. Nordic Heating and Cooling: Nordic Approach to EU’s Heating and Cooling Strategy; Nordic Council of Ministers: Copenhagen, Denmark, 2017; ISBN 92-893-4992-1.
  59. Beyer, B.; Geldermann, J.; Lauven, L.-P. Agent-Based Model of the German Heating Market: Simulations Concerning the Use of Wood Pellets and the Sustainability of the Market. In Proceedings of the 2017 14th International Conference on the European Energy Market (EEM), Dresden, Germany, 6–9 June 2017; pp. 1–6. [Google Scholar]
  60. Ladanai, S.; Vinterbäck, J. Global Potential of Sustainable Biomass for Energy; Department of Energy and Technology, Swedish University of Agricultural Sciences: Uppsala, Sweden, 2009; Volume 13. [Google Scholar]
  61. ESCAP. ESCAP Newsletter. 2021. Available online: https://repository.unescap.org/bitstream/handle/20.500.12870/5959/ESCAP-2021-JN-ESCAP-Newsletter-Jan.pdf?sequence=1 (accessed on 24 October 2023).
  62. Proskurina, S.; Heinimö, J.; Schipfer, F.; Vakkilainen, E. Biomass for Industrial Applications: The Role of Torrefaction. Renew. Energy 2017, 111, 265–274. [Google Scholar] [CrossRef]
  63. Stelte, W.; Nielsen, N.P.K.; Hansen, H.O.; Dahl, J.; Shang, L.; Sanadi, A.R. Pelletizing Properties of Torrefied Wheat Straw. Biomass Bioenergy 2013, 49, 214–221. [Google Scholar] [CrossRef]
  64. Uslu, A.; Faaij, A.P.C.; Bergman, P.C.A. Pre-Treatment Technologies, and Their Effect on International Bioenergy Supply Chain Logistics. Techno-Economic Evaluation of Torrefaction, Fast Pyrolysis and Pelletisation. Energy 2008, 33, 1206–1223. [Google Scholar] [CrossRef]
  65. Levinson, R. Introducing Socio-Scientific Inquiry-Based Learning (SSIBL). Sch. Sci. Rev. 2018, 100, 31–35. [Google Scholar]
  66. Strauss, W. A Short Update on the Japanese Industrial Wood Pellet Markets: Policies, and How They Will Drive Current and Future Demand. Future Metrics. 2017. Available online: http://www.futuremetrics.info/wp-content/uploads/2017/08/Japanese_Industrial_Wood_Pellet_Markets_August_2017_by_FutureMetrics.pdf (accessed on 16 October 2023).
  67. Winter, N. Renewables 2022 Global Status Report China Factsheet-Key Headlines; United Nations Environment Programme: Nairobi, Kenya, 2022. [Google Scholar]
  68. World Bioenergy Association. Pellets—A Fast Growing Energy. Available online: http://www.worldbioenergy.org/uploads/Factsheet%20-%20Pellets.pdf (accessed on 26 August 2023).
  69. World Bioenergy Association. Global Bioenergy Statistics 2014. Available online: https://www.worldbioenergy.org/uploads/WBA%20Global%20Bioenergy%20Statistics%202014.pdf (accessed on 26 August 2023).
  70. Bung, B. Forecasting the European Wood Biomass Market: An Investigation of Drivers That Impact Supply and Demand. 2016. Available online: https://open.library.ubc.ca/media/stream/pdf/52966/1.0314331/5 (accessed on 11 November 2023).
  71. Bioenergy Europe. Statistical Pellet Report 2019. Available online: https://epc.bioenergyeurope.org/wp-content/uploads/2020/02/SR19_Pellet_final-web-1.pdf (accessed on 6 October 2023).
  72. Kummamuru, B. WBA Global Bioenergy Statistics 2017; World Bioenergy Association: Stockholm, Sweden, 2016. [Google Scholar]
  73. Hernández, D.; Fernández-Puratich, H.; Rebolledo-Leiva, R.; Tenreiro, C.; Gabriel, D. Evaluation of Sustainable Manufacturing of Pellets Combining Wastes from Olive Oil and Forestry Industries. Ind. Crops Prod. 2019, 134, 338–346. [Google Scholar] [CrossRef]
  74. Beyer, B.; Lauven, L.; Schröder, T. Bioenergy Pathway Sustainability Assessment in Germany; Georg-August University Goettingen: Göttingen, Germany, 2014. [Google Scholar]
  75. Schröder, T.; Lauven, L.-P.; Beyer, B.; Lerche, N.; Geldermann, J. Using PROMETHEE to Assess Bioenergy Pathways. Cent. Eur. J. Oper. Res. 2019, 27, 287–309. [Google Scholar] [CrossRef]
  76. Tumuluru, J.S. Effect of Pellet Die Diameter on Density and Durability of Pellets Made from High Moisture Woody and Herbaceous Biomass. Carbon Resour. Convers. 2018, 1, 44–54. [Google Scholar] [CrossRef]
  77. Prvulovic, S.; Gluvakov, Z.; Tolmac, J.; Tolmac, D.; Matic, M.; Brkic, M. Methods for Determination of Biomass Energy Pellet Quality. Energy Fuels 2014, 28, 2013–2018. [Google Scholar] [CrossRef]
  78. Arshadi, M.; Gref, R.; Geladi, P.; Dahlqvist, S.-A.; Lestander, T. The Influence of Raw Material Characteristics on the Industrial Pelletizing Process and Pellet Quality. Fuel Process. Technol. 2008, 89, 1442–1447. [Google Scholar] [CrossRef]
  79. Abdoli, M.A.; Golzary, A.; Hosseini, A.; Sadeghi, P. Wood Pellet Production Standards. In Wood Pellet as a Renewable Source of Energy; University of Tehran Science and Humanities Series; Springer International Publishing: Cham, Switzerland, 2018; pp. 101–110. ISBN 978-3-319-74481-0. [Google Scholar]
  80. Obernberger, I.; Thek, G. The Pellet Handbook: The Production and Thermal Utilisation of Pellets; Routledge: London, UK, 2010; ISBN 1-84407-631-8. [Google Scholar]
  81. Kofman, P.D. Review of Worldwide Standards for Solid Biofuels. COFORD Process. Prod. 2016, 39, 1–12. [Google Scholar]
  82. ISO 17225-2:2014; Solid Biofuels—Fuel Specifications and Classes Part 2: Graded Wood Pellets. The British Standards Institution: London, UK, 2014.
  83. ISO 17225-6:2014; Solid Biofuels—Fuel Specifications and Classes Part 6: Graded Non-Woody Pellets. The British Standards Institution: London, UK, 2014.
  84. Alakangas, E. European Standards for Fuel Specification and Classes of Solid Biofuels. In Solid Biofuels for Energy; Springer: Berlin/Heidelberg, Germany, 2011; pp. 21–41. [Google Scholar]
  85. ISO 17225-2:2021; Solid Biofuels—Fuel Specifications and Classes Part 2: Graded Wood Pellets. The British Standards Institution: London, UK, 2021.
  86. ISO 17225-6:2021; Solid Biofuels—Fuel Specifications and Classes Part 6: Graded Non-Woody Pellets. The British Standards Institution: London, UK, 2021.
  87. Manzoor, B.; Othman, I.; Durdyev, S.; Ismail, S.; Wahab, M.H. Influence of Artificial Intelligence in Civil Engineering toward Sustainable Development—A Systematic Literature Review. Appl. Syst. Innov. 2021, 4, 52. [Google Scholar] [CrossRef]
  88. Picchio, R.; Proto, A.R.; Civitarese, V.; Di Marzio, N.; Latterini, F. Recent Contributions of Some Fields of the Electronics in Development of Forest Operations Technologies. Electronics 2019, 8, 1465. [Google Scholar] [CrossRef]
  89. Tobi, R.C.A.; Harris, F.; Rana, R.; Brown, K.A.; Quaife, M.; Green, R. Sustainable Diet Dimensions. Comparing Consumer Preference for Nutrition, Environmental and Social Responsibility Food Labelling: A Systematic Review. Sustainability 2019, 11, 6575. [Google Scholar] [CrossRef]
  90. Harun, N.Y.; Afzal, M.T. Chemical and Mechanical Properties of Pellets Made from Agricultural and Woody Biomass Blends. Trans. ASABE 2015, 58, 921–930. [Google Scholar] [CrossRef]
  91. Acampora, A.; Civitarese, V.; Sperandio, G.; Rezaei, N. Qualitative Characterization of the Pellet Obtained from Hazelnut and Olive Tree Pruning. Energies 2021, 14, 4083. [Google Scholar] [CrossRef]
  92. De Souza, H.J.P.L.; Arantes, M.D.C.; Vidaurre, G.B.; Andrade, C.R.; Carneiro, A.d.C.O.; de Souza, D.P.L.; Protásio, T.d.P. Pelletization of Eucalyptus Wood and Coffee Growing Wastes: Strategies for Biomass Valorization and Sustainable Bioenergy Production. Renew. Energy 2020, 149, 128–140. [Google Scholar] [CrossRef]
  93. Da Silva, S.B.; Arantes, M.D.C.; de Andrade, J.K.B.; Andrade, C.R.; Carneiro, A.D.C.O.; Protásio, T.D.P. Influence of Physical and Chemical Compositions on the Properties and Energy Use of Lignocellulosic Biomass Pellets in Brazil. Renew. Energy 2020, 147, 1870–1879. [Google Scholar] [CrossRef]
  94. Pegoretti Leite de Souza, H.J.; Muñoz, F.; Mendonça, R.T.; Sáez, K.; Olave, R.; Segura, C.; de Souza, D.P.L.; de Paula Protásio, T.; Rodríguez-Soalleiro, R. Influence of Lignin Distribution, Physicochemical Characteristics and Microstructure on the Quality of Biofuel Pellets Made from Four Different Types of Biomass. Renew. Energy 2021, 163, 1802–1816. [Google Scholar] [CrossRef]
  95. Santana, D.A.R.; Scatolino, M.V.; Lima, M.D.R.; de Oliveira Barros Junior, U.; Garcia, D.P.; Andrade, C.R.; de Cássia Oliveira Carneiro, A.; Trugilho, P.F.; de Paula Protásio, T. Pelletizing of Lignocellulosic Wastes as an Environmentally Friendly Solution for the Energy Supply: Insights on the Properties of Pellets from Brazilian Biomasses. Environ. Sci. Pollut. Res. 2021, 28, 11598–11617. [Google Scholar] [CrossRef]
  96. Senila, L.; Tenu, I.; Carlescu, P.; Corduneanu, O.R.; Dumitrachi, E.P.; Kovacs, E.; Scurtu, D.A.; Cadar, O.; Becze, A.; Senila, M.; et al. Sustainable Biomass Pellets Production Using Vineyard Wastes. Agriculture 2020, 10, 501. [Google Scholar] [CrossRef]
  97. Rezania, S.; Md Din, M.F.; Kamaruddin, S.F.; Taib, S.M.; Singh, L.; Yong, E.L.; Dahalan, F.A. Evaluation of Water Hyacinth (Eichhornia Crassipes) as a Potential Raw Material Source for Briquette Production. Energy 2016, 111, 768–773. [Google Scholar] [CrossRef]
  98. Almeida, L.F.P.d.; Sola, A.V.H.; Behainne, J.J.R. Sugarcane Bagasse Pellets: Characterization and Comparative Analysis. Acta Sci. Technol. 2017, 39, 461. [Google Scholar] [CrossRef]
  99. Carrillo-Parra, A.; Contreras-Trejo, J.C.; Pompa-García, M.; Pulgarín-Gámiz, M.Á.; Rutiaga-Quiñones, J.G.; Pámanes-Carrasco, G.; Ngangyo-Heya, M. Agro-Pellets from Oil Palm Residues/Pine Sawdust Mixtures: Relationships of Their Physical, Mechanical and Energetic Properties, with the Raw Material Chemical Structure. Appl. Sci. 2020, 10, 6383. [Google Scholar] [CrossRef]
  100. Garcia, D.P.; Caraschi, J.C.; Ventorim, G.; Vieira, F.H.A.; de Paula Protásio, T. Assessment of Plant Biomass for Pellet Production Using Multivariate Statistics (PCA and HCA). Renew. Energy 2019, 139, 796–805. [Google Scholar] [CrossRef]
  101. Pradhan, P.; Arora, A.; Mahajani, S.M. Pilot Scale Evaluation of Fuel Pellets Production from Garden Waste Biomass. Energy Sustain. Dev. 2018, 43, 1–14. [Google Scholar] [CrossRef]
  102. Scatolino, M.V.; Neto, L.F.C.; Protásio, T.P.; Carneiro, A.C.O.; Andrade, C.R.; Guimarães Júnior, J.B.; Mendes, L.M. Options for Generation of Sustainable Energy: Production of Pellets Based on Combinations Between Lignocellulosic Biomasses. Waste Biomass Valorization 2018, 9, 479–489. [Google Scholar] [CrossRef]
  103. Tenorio, C.; Roque, R.; Valaert, J. Characterisation of Pellets Made from Oil Palm Residues in Costa Rica. J. Palm Oil Res. 2016, 28, 198–210. [Google Scholar] [CrossRef]
  104. Trejo-Zamudio, D.; Gutiérrez-Antonio, C.; García-Trejo, J.F.; Feregrino-Pérez, A.A.; Toledano-Ayala, M. Production of Fuel Pellets from Bean Crop Residues (Phaseolus Vulgaris). IET Renew. Power Gener. 2021, 16, 2978–2987. [Google Scholar] [CrossRef]
  105. Chavalparit, O.; Ongwandee, M.; Trangkaprasith, K. Production of Pelletized Fuel from Biodiesel-Production Wastes: Oil Palm Fronds and Crude Glycerin. Eng. J. 2013, 17, 61–70. [Google Scholar] [CrossRef]
  106. Azócar, L.; Hermosilla, N.; Gay, A.; Rocha, S.; Díaz, J.; Jara, P. Brown Pellet Production Using Wheat Straw from Southern Cities in Chile. Fuel 2019, 237, 823–832. [Google Scholar] [CrossRef]
  107. Amirta, R.; Anwar, T.; Sudrajat; Yuliansyah; Suwinarti, W. Trial Production of Fuel Pellet from Acacia Mangium Bark Waste Biomass. IOP Conf. Ser. Earth Environ. Sci. 2018, 144, 012040. [Google Scholar] [CrossRef]
  108. Nielsen, N.P.K.; Holm, J.K.; Felby, C. Effect of Fiber Orientation on Compression and Frictional Properties of Sawdust Particles in Fuel Pellet Production. Energy Fuels 2009, 23, 3211–3216. [Google Scholar] [CrossRef]
  109. Erlich, C.; Bjornbom, E.; Bolado, D.; Giner, M.; Fransson, T. Pyrolysis and Gasification of Pellets from Sugar Cane Bagasse and Wood. Fuel 2006, 85, 1535–1540. [Google Scholar] [CrossRef]
  110. Erlich, C.; Fransson, T.H. Downdraft Gasification of Pellets Made of Wood, Palm-Oil Residues Respective Bagasse: Experimental Study. Appl. Energy 2011, 88, 899–908. [Google Scholar] [CrossRef]
  111. Lam, P.S.; Lam, P.Y.; Sokhansanj, S.; Bi, X.T.; Lim, C.J. Mechanical and Compositional Characteristics of Steam-Treated Douglas Fir (Pseudotsuga Menziesii L.) during Pelletization. Biomass Bioenergy 2013, 56, 116–126. [Google Scholar] [CrossRef]
  112. Chen, W.-H.; Peng, J.; Bi, X.T. A State-of-the-Art Review of Biomass Torrefaction, Densification and Applications. Renew. Sustain. Energy Rev. 2015, 44, 847–866. [Google Scholar] [CrossRef]
  113. Bridgeman, T.G.; Jones, J.M.; Shield, I.; Williams, P.T. Torrefaction of Reed Canary Grass, Wheat Straw and Willow to Enhance Solid Fuel Qualities and Combustion Properties. Fuel 2008, 87, 844–856. [Google Scholar] [CrossRef]
  114. Adapa, P.; Tabil, L.; Schoenau, G.; Opoku, A. Pelleting Characteristics of Selected Biomass with and without Steam Explosion Pretreatment. Int. J. Agric. Biol. Eng. 2010, 3, 61–79. [Google Scholar] [CrossRef]
  115. Said, N.; Abdel daiem, M.M.; García-Maraver, A.; Zamorano, M. Influence of Densification Parameters on Quality Properties of Rice Straw Pellets. Fuel Process. Technol. 2015, 138, 56–64. [Google Scholar] [CrossRef]
  116. Obernberger, I.; Thek, G. Physical Characterisation and Chemical Composition of Densified Biomass Fuels with Regard to Their Combustion Behaviour. Biomass Bioenergy 2004, 27, 653–669. [Google Scholar] [CrossRef]
  117. Pradhan, P.; Mahajani, S.M.; Arora, A. Production and Utilization of Fuel Pellets from Biomass: A Review. Fuel Process. Technol. 2018, 181, 215–232. [Google Scholar] [CrossRef]
  118. Holm, J.K.; Henriksen, U.B.; Hustad, J.E.; Sørensen, L.H. Toward an Understanding of Controlling Parameters in Softwood and Hardwood Pellets Production. Energy Fuels 2006, 20, 2686–2694. [Google Scholar] [CrossRef]
  119. Liu, Z.; Quek, A.; Balasubramanian, R. Preparation and Characterization of Fuel Pellets from Woody Biomass, Agro-Residues and Their Corresponding Hydrochars. Appl. Energy 2014, 113, 1315–1322. [Google Scholar] [CrossRef]
  120. McKeown, M.S.; Trabelsi, S.; Nelson, S.O.; Tollner, E.W. Microwave Sensing of Moisture in Flowing Biomass Pellets. Biosyst. Eng. 2017, 155, 152–160. [Google Scholar] [CrossRef]
  121. Nyström, J.; Dahlquist, E. Methods for Determination of Moisture Content in Woodchips for Power Plants—A Review. Fuel 2004, 83, 773–779. [Google Scholar] [CrossRef]
  122. Tumuluru, J.S.; Wright, C.T.; Hess, J.R.; Kenney, K.L. A Review of Biomass Densification Systems to Develop Uniform Feedstock Commodities for Bioenergy Application. Biofuels Bioprod. Bioref. 2011, 5, 683–707. [Google Scholar] [CrossRef]
  123. González, W.A.; López, D.; Pérez, J.F. Biofuel Quality Analysis of Fallen Leaf Pellets: Effect of Moisture and Glycerol Contents as Binders. Renew. Energy 2020, 147, 1139–1150. [Google Scholar] [CrossRef]
  124. Nguyen, Q.N.; Cloutier, A.; Achim, A.; Stevanovic, T. Effect of Process Parameters and Raw Material Characteristics on Physical and Mechanical Properties of Wood Pellets Made from Sugar Maple Particles. Biomass Bioenergy 2015, 80, 338–349. [Google Scholar] [CrossRef]
  125. Serrano, C.; Monedero, E.; Lapuerta, M.; Portero, H. Effect of Moisture Content, Particle Size and Pine Addition on Quality Parameters of Barley Straw Pellets. Fuel Process. Technol. 2011, 92, 699–706. [Google Scholar] [CrossRef]
  126. Lehtikangas, P. Quality Properties of Pelletised Sawdust, Logging Residues and Bark. Biomass Bioenergy 2001, 20, 351–360. [Google Scholar] [CrossRef]
  127. Duca, D.; Riva, G.; Foppa Pedretti, E.; Toscano, G. Wood Pellet Quality with Respect to EN 14961-2 Standard and Certifications. Fuel 2014, 135, 9–14. [Google Scholar] [CrossRef]
  128. Toscano, G.; Duca, D.; Foppa Pedretti, E.; Pizzi, A.; Rossini, G.; Mengarelli, C.; Mancini, M. Investigation of Woodchip Quality: Relationship between the Most Important Chemical and Physical Parameters. Energy 2016, 106, 38–44. [Google Scholar] [CrossRef]
  129. Bustamante-García, V.; Carrillo-Parra, A.; González-Rodríguez, H.; Ramírez-Lozano, R.G.; Corral-Rivas, J.J.; Garza-Ocañas, F. Evaluation of a Charcoal Production Process from Forest Residues of Quercus Sideroxyla Humb., & Bonpl. in a Brazilian Beehive Kiln. Ind. Crops Prod. 2013, 42, 169–174. [Google Scholar] [CrossRef]
  130. Carroll, J.P.; Finnan, J. Physical and Chemical Properties of Pellets from Energy Crops and Cereal Straws. Biosyst. Eng. 2012, 112, 151–159. [Google Scholar] [CrossRef]
  131. Garcia, D.P.; Caraschi, J.C.; Ventorim, G. Emissões de Gases Do Efeito Estufa da Queima de Pellets de Madeira. Floresta 2017, 47, 297–306. [Google Scholar] [CrossRef]
  132. Poddar, S.; Kamruzzaman, M.; Sujan, S.M.A.; Hossain, M.; Jamal, M.S.; Gafur, M.A.; Khanam, M. Effect of Compression Pressure on Lignocellulosic Biomass Pellet to Improve Fuel Properties: Higher Heating Value. Fuel 2014, 131, 43–48. [Google Scholar] [CrossRef]
  133. Kishore, V.V.N. Renewable Energy Engineering and Technology: Principles and Practice; The Energy and Resources Institute (TERI): Mithapur, India, 2010; ISBN 81-7993-221-4. [Google Scholar]
  134. Supatata, N.; Buates, J.; Hariyanont, P. Characterization of Fuel Briquettes Made from Sewage Sludge Mixed with Water Hyacinth and Sewage Sludge Mixed with Sedge. Int. J. Environ. Sci. Dev. 2013, 1, 1793–8201. [Google Scholar] [CrossRef]
  135. Akowuah, J.O.; Kemausuor, F.; Mitchual, S.J. Physico-Chemical Characteristics and Market Potential of Sawdust Charcoal Briquette. Int. J. Energy Environ. Eng. 2012, 3, 20. [Google Scholar] [CrossRef]
  136. Mościcki, K.J.; Niedźwiecki, Ł.; Owczarek, P.; Wnukowski, M. Commoditization of Wet and High Ash Biomass: Wet Torrefaction-a Review. J. Power Technol. 2017, 97, 354–369. [Google Scholar]
  137. Prasongthum, N.; Duangwongsa, N.; Khowattana, P.; Suemanotham, A.; Wongharn, P.; Thanmongkhon, Y.; Reubroycharoen, P.; Attanatho, L. Influence of Torrefaction on Yields and Characteristics of Densified Solid Biofuel. J. Phys. Conf. Ser. 2022, 2175, 012027. [Google Scholar] [CrossRef]
  138. Shao, J.; Cheng, W.; Zhu, Y.; Yang, W.; Fan, J.; Liu, H.; Yang, H.; Chen, H. Effects of Combined Torrefaction and Pelletization on Particulate Matter Emission from Biomass Pellet Combustion. Energy Fuels 2019, 33, 8777–8785. [Google Scholar] [CrossRef]
  139. Acharya, B.; Dutta, A. Fuel Property Enhancement of Lignocellulosic and Nonlignocellulosic Biomass through Torrefaction. Biomass Conv. Bioref. 2016, 6, 139–149. [Google Scholar] [CrossRef]
  140. Sadaka, S.; Negi, S. Improvements of Biomass Physical and Thermochemical Characteristics via Torrefaction Process. Environ. Prog. Sustain. Energy 2009, 28, 427–434. [Google Scholar] [CrossRef]
  141. Niedziółka, I.; Szpryngiel, M.; Kachel-Jakubowska, M.; Kraszkiewicz, A.; Zawiślak, K.; Sobczak, P.; Nadulski, R. Assessment of the Energetic and Mechanical Properties of Pellets Produced from Agricultural Biomass. Renew. Energy 2015, 76, 312–317. [Google Scholar] [CrossRef]
  142. Picchio, R.; Latterini, F.; Venanzi, R.; Stefanoni, W.; Suardi, A. Pellet Production from Woody and Non-Woody Feedstocks: A Review on Biomass Quality Evaluation. Energies 2020, 13, 2937. [Google Scholar] [CrossRef]
  143. Kaddour, O.; Alavi, S.; Behnke, K. Factors Influencing the Quality of Extruded Sinking Aquatic Feed Pellets. Misr J. Agric. Eng. 2010, 27, 1953–1982. [Google Scholar] [CrossRef]
  144. Olugbade, T.O.; Ojo, O.T. Biomass Torrefaction for the Production of High-Grade Solid Biofuels: A Review. Bioenerg. Res. 2020, 13, 999–1015. [Google Scholar] [CrossRef]
  145. Thomas, M.; van Zuilichem, D.J.; van der Poel, A.F.B. Physical Quality of Pelleted Animal Feed. 2. Contribution of Processes and Its Conditions. Anim. Feed Sci. Technol. 1997, 64, 173–192. [Google Scholar] [CrossRef]
  146. Tabil, L., Jr.; Sokhansanj, S. Process Conditions Affecting the Physical Quality of Alfalfa Pellets. Appl. Eng. Agric. 1996, 12, 345–350. [Google Scholar] [CrossRef]
  147. Dyjakon, A.; Noszczyk, T. The Influence of Freezing Temperature Storage on the Mechanical Durability of Commercial Pellets from Biomass. Energies 2019, 12, 2627. [Google Scholar] [CrossRef]
  148. Faria, W.S.; Protásio, T.d.P.; Trugilho, P.F.; Pereira, B.L.C.; Carneiro, A.; Andrade, C.R.; Guimarães Junior, J.B. Transformation of Lignocellulosic Waste of Coffee into Pellets for Thermal Power Generation. Coffee Sci. 2016, 11, 137–147. [Google Scholar]
  149. Castellano, J.M.; Gómez, M.; Fernández, M.; Esteban, L.S.; Carrasco, J.E. Study on the Effects of Raw Materials Composition and Pelletization Conditions on the Quality and Properties of Pellets Obtained from Different Woody and Non Woody Biomasses. Fuel 2015, 139, 629–636. [Google Scholar] [CrossRef]
  150. Lavergne, S.; Larsson, S.H.; Da Silva Perez, D.; Marchand, M.; Campargue, M.; Dupont, C. Effect of Process Parameters and Biomass Composition on Flat-Die Pellet Production from Underexploited Forest and Agricultural Biomass. Fuel 2021, 302, 121076. [Google Scholar] [CrossRef]
  151. Whittaker, C.; Shield, I. Factors Affecting Wood, Energy Grass and Straw Pellet Durability—A Review. Renew. Sustain. Energy Rev. 2017, 71, 1–11. [Google Scholar] [CrossRef]
  152. Ungureanu, N.; Vladut, V.; Voicu, G.; Dinca, M.-N.; Zabava, B.-S. Influence of Biomass Moisture Content on Pellet Properties—Review. In Proceedings of the 17th International Scientific Conference Engineering for Rural Development, Jelgava, Latvia, 23–25 May 2018. [Google Scholar]
  153. Harun, N.Y.; Afzal, M.T. Effect of Particle Size on Mechanical Properties of Pellets Made from Biomass Blends. Procedia Eng. 2016, 148, 93–99. [Google Scholar] [CrossRef]
  154. Mani, S.; Tabil, L.G.; Sokhansanj, S. Compaction of Biomass Grinds—An Overview of Compaction of Biomass Grinds. Powder Handl. Process. 2003, 15, 160–168. [Google Scholar]
  155. Stelte, W.; Holm, J.K.; Sanadi, A.R.; Barsberg, S.; Ahrenfeldt, J.; Henriksen, U.B. A Study of Bonding and Failure Mechanisms in Fuel Pellets from Different Biomass Resources. Biomass Bioenergy 2011, 35, 910–918. [Google Scholar] [CrossRef]
  156. Lisowski, A.; Matkowski, P.; Dąbrowska, M.; Piątek, M.; Świętochowski, A.; Klonowski, J.; Mieszkalski, L.; Reshetiuk, V. Particle Size Distribution and Physicochemical Properties of Pellets Made of Straw, Hay, and Their Blends. Waste Biomass Valor. 2020, 11, 63–75. [Google Scholar] [CrossRef]
  157. Obidziński, S. Pelletization Process of Postproduction Plant Waste. Int. Agrophys. 2012, 26, 279–284. [Google Scholar] [CrossRef]
  158. Carone, M.T.; Pantaleo, A.; Pellerano, A. Influence of Process Parameters and Biomass Characteristics on the Durability of Pellets from the Pruning Residues of Olea europaea L. Biomass Bioenergy 2011, 35, 402–410. [Google Scholar] [CrossRef]
  159. Gil, M.V.; Oulego, P.; Casal, M.D.; Pevida, C.; Pis, J.J.; Rubiera, F. Mechanical Durability and Combustion Characteristics of Pellets from Biomass Blends. Bioresour. Technol. 2010, 101, 8859–8867. [Google Scholar] [CrossRef]
  160. Gilbert, P.; Ryu, C.; Sharifi, V.; Swithenbank, J. Effect of Process Parameters on Pelletisation of Herbaceous Crops. Fuel 2009, 88, 1491–1497. [Google Scholar] [CrossRef]
  161. Stasiak, M.; Molenda, M.; Bańda, M.; Wiącek, J.; Parafiniuk, P.; Gondek, E. Mechanical and Combustion Properties of Sawdust—Straw Pellets Blended in Different Proportions. Fuel Process. Technol. 2017, 156, 366–375. [Google Scholar] [CrossRef]
  162. Kaliyan, N.; Morey, R.V. Natural Binders and Solid Bridge Type Binding Mechanisms in Briquettes and Pellets Made from Corn Stover and Switchgrass. Bioresour. Technol. 2010, 101, 1082–1090. [Google Scholar] [CrossRef] [PubMed]
  163. Dyjakon, A.; Noszczyk, T.; Mostek, A. Mechanical Durability and Grindability of Pellets after Torrefaction Process. Energies 2021, 14, 6772. [Google Scholar] [CrossRef]
  164. Dafnomilis, I.; Lodewijks, G.; Junginger, M.; Schott, D.L. Evaluation of Wood Pellet Handling in Import Terminals. Biomass Bioenergy 2018, 117, 10–23. [Google Scholar] [CrossRef]
  165. Kofman, P. Simple Ways to Check Wood Pellet Quality. Bioenergy News 2007, 2006, 8–9. [Google Scholar]
  166. Acharya, B.; Sule, I.; Dutta, A. A Review on Advances of Torrefaction Technologies for Biomass Processing. Biomass Conv. Bioref. 2012, 2, 349–369. [Google Scholar] [CrossRef]
  167. Bergman, P.C.; Boersma, A.R.; Zwart, R.W.R.; Kiel, J.H.A. Torrefaction for Biomass Co-Firing in Existing Coal-Fired Power Stations; ECN-C-05-013; Energy Research Centre of the Netherlands ECN: Petten, The Netherlands, 2005; Volume 31. [Google Scholar]
  168. Pimchuai, A.; Dutta, A.; Basu, P. Torrefaction of Agriculture Residue to Enhance Combustible Properties. Energy Fuels 2010, 24, 4638–4645. [Google Scholar] [CrossRef]
  169. Demirbas, A. Combustion Characteristics of Different Biomass Fuels. Prog. Energy Combust. Sci. 2004, 30, 219–230. [Google Scholar] [CrossRef]
  170. Biswas, A.K.; Rudolfsson, M.; Broström, M.; Umeki, K. Effect of Pelletizing Conditions on Combustion Behaviour of Single Wood Pellet. Appl. Energy 2014, 119, 79–84. [Google Scholar] [CrossRef]
  171. Wang, W.; Lemaire, R.; Bensakhria, A.; Luart, D. Review on the Catalytic Effects of Alkali and Alkaline Earth Metals (AAEMs) Including Sodium, Potassium, Calcium and Magnesium on the Pyrolysis of Lignocellulosic Biomass and on the Co-Pyrolysis of Coal with Biomass. J. Anal. Appl. Pyrolysis 2022, 163, 105479. [Google Scholar] [CrossRef]
  172. De Paula Protásio, T.; Bufalino, L.; Denzin Tonoli, G.H.; Guimarães, M., Jr.; Trugilho, P.F.; Mendes, L.M. Brazilian Lignocellulosic Wastes for Bioenergy Production: Characterization and Comparison with Fossil Fuels. BioResources 2013, 8, 1166–1185. [Google Scholar]
  173. Ríos-Badrán, I.M.; Luzardo-Ocampo, I.; García-Trejo, J.F.; Santos-Cruz, J.; Gutiérrez-Antonio, C. Production and Characterization of Fuel Pellets from Rice Husk and Wheat Straw. Renew. Energy 2020, 145, 500–507. [Google Scholar] [CrossRef]
  174. Zeng, T.; Weller, N.; Pollex, A.; Lenz, V. Blended Biomass Pellets as Fuel for Small Scale Combustion Appliances: Influence on Gaseous and Total Particulate Matter Emissions and Applicability of Fuel Indices. Fuel 2016, 184, 689–700. [Google Scholar] [CrossRef]
  175. Kataki, R.; Konwer, D. Fuelwood Characteristics of Indigenous Tree Species of North-East India. Biomass Bioenergy 2002, 22, 433–437. [Google Scholar] [CrossRef]
  176. Onochie, U.; Obanor, A.; Aliu, S.; Igbodaro, O. Proximate and Ultimate Analysis of Fuel Pellets from Oil Palm Residues. Niger. J. Technol. 2017, 36, 987–990. [Google Scholar] [CrossRef]
Figure 1. Selection process of the analyzed studies.
Figure 1. Selection process of the analyzed studies.
Energies 17 00054 g001
Figure 2. Average pellet diameter.
Figure 2. Average pellet diameter.
Energies 17 00054 g002
Figure 3. Average pellet length.
Figure 3. Average pellet length.
Energies 17 00054 g003
Figure 4. Mean moisture content of the pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 4. Mean moisture content of the pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g004
Figure 5. Mean and permissible range of ash content of the pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 5. Mean and permissible range of ash content of the pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g005
Figure 6. Mean and permissible range of mechanical durability of pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 6. Mean and permissible range of mechanical durability of pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g006
Figure 7. Mean and permissible range of fine particle content of pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 7. Mean and permissible range of fine particle content of pellets. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g007
Figure 8. Mean and permissible range of gross calorific value of pellets.
Figure 8. Mean and permissible range of gross calorific value of pellets.
Energies 17 00054 g008
Figure 9. Mean and permissible range of bulk density of pellets.
Figure 9. Mean and permissible range of bulk density of pellets.
Energies 17 00054 g009
Figure 10. Mean and permissible range of nitrogen. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 10. Mean and permissible range of nitrogen. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g010
Figure 11. Mean and permissible range of sulfur.
Figure 11. Mean and permissible range of sulfur.
Energies 17 00054 g011
Figure 12. Mean and permissible range of volatile matter. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 12. Mean and permissible range of volatile matter. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g012
Figure 13. Mean and permissible range of fixed carbon content. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Figure 13. Mean and permissible range of fixed carbon content. The arrow can be seen when the range of the final results goes beyond the range of the graph.
Energies 17 00054 g013
Table 1. Specification of graded woody and non-woody pellets.
Table 1. Specification of graded woody and non-woody pellets.
ParameterUnitEN ISO 17225-2EN ISO 17225-6
Utility-Commercial and Residential
Applications
Industrial UseIndustrial Use
A1A2A3I1I2I3AB
Diameter (D)Mm6 ± 16 ± 16 ± 16 ± 16 ± 16 ± 16–256–25
Length (L)Mm3.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 403.15 ≤ L ≤ 40
Moisture content (MC)%≤10≤10≤10≤10≤10≤10≤12≤15
Ash content (A)%≤0.7≤1.2≤2≤1≤1.5≤3≤6≤10
Mechanical durability (Du)%≥98≥97.5≥96.597.5 ≤ Du ≤ 99.097.0 ≤ Du ≤ 99.096.5 ≤ Du ≤ 99.0≥97.5≥96
Fines (F)%≤1≤1≤1≤4≤5≤6≤2≤3
Net calorific value (NCV)MJ/kg≥16.5≥16.5≥16.5≥16.5≥16.5≥16.5≥14.5≥14.5
Bulk density (BD)kg/m3600 ≤ BD ≤ 750600 ≤ BD ≤ 750600 ≤ BD ≤ 750600≤600≤600≤600≤600≤
N%≤0.3≤0.5≤1≤0.3≤0.3≤0.6≤1.5≤2.0
S%≤0.04≤0.04≤0.05≤0.05≤0.05≤0.05≤0.2≤0.3
Cl%≤0.02≤0.02≤0.03≤0.03≤0.05≤0.1≤0.1≤0.3
Asmg/kg≤1≤1≤1≤2≤2≤2≤1-
Cdmg/kg≤0.5≤0.5≤0.5≤1≤1≤1≤0.5-
Crmg/kg≤10≤10≤10≤15≤15≤15≤50-
Cumg/kg≤10≤10≤10≤20≤20≤20≤20-
Pbmg/kg≤10≤10≤10≤10≤10≤10≤10-
Hgmg/kg≤0.1≤0.1≤0.1≤0.1≤0.1≤0.1≤0.1-
Nimg/kg≤10≤10≤10≤10≤10≤10≤10-
Table 2. Different standards used in different studies to test the quality parameters of the pellets.
Table 2. Different standards used in different studies to test the quality parameters of the pellets.
Test ParameterStandard Followed for TestingReferences
Pellet dimensionsDIN EN 16127[91,92]
EN 16127[92,93,94,95]
EN 17829 (2015)[96]
ASTM standard E711-87[97]
Directly from vernier caliper[13,90,98,99,100,101,102,103,104]
Moisture contentPROY-NOM-211-SSA1-2002 standard[104]
DIN EN 14774-1[92,102]
EN ISO 18134[13,96]
EN 14774-1[93,94]
ASTM D1762[103]
UNE-EN 14774-3[99]
D1762, ISO 1822[100]
NBR 7993[98]
Oven-drying method (105 °C for 24 h)[90,91,101]
Ash contentASTM D3174-04, 2004[103,105]
ASTM D 1762-84[95,100,102,106]
DIN EN 14775[91,92,93,94]
EN ISO 18122[13,96]
NBR 8112[98]
NREL/TP-510-42622[90]
ONORM M7135[107]
UNE-EN 14775[99]
Mechanical durabilityDIN EN 15210-1[93,94,95,98,102,103,106]
EN 17831-1:2015[13,96]
ASABE standards (ASABE, 2007)[101]
ASTM E1641-04[90]
Bulk densityDIN EN 15103[92,93,94,95,98,100,102,104,106]
BS EN ISO 17828 (2015)[13,91,96,99]
SS 187120[107]
ASTM E1641-04[90]
X-ray densitometry method[103]
N and S contentDIN EN 15104[92,94]
ISO 16948 (2015)[13,96]
ASTM D5291[90]
ASTM E778/08[100]
Using CHNS (O) analyzers[91,101]
Using elemental analyzers[95,98,103,105,106]
Volatile matter contentASTM D1762-84[92,95,100,102,103,106]
ASTM D3175-07, 2007[90,97]
EN 15148[93,99]
DIN EN 51720[13]
EN ISO 18123[13]
NBR 8112[98]
Gross calorific valueBomb calorimeter methodAll
Table 3. Summary and the properties of pellets in selected studies. NA—not available.
Table 3. Summary and the properties of pellets in selected studies. NA—not available.
Property of the PelletPhysical
Properties
Mechanical PropertiesEnergy PropertiesChemical
Properties
DLDuFinesBDHardnessMCAshCVVMFCNS
Unitmmmm%%kg/m3%%%MJ/kg%%%%
1. Pellet Standards
EN ISO 17225-6 category A6 to 253.15≤ L ≤ 40≥97.5≤2600≤NA≤12≤6≥14.5NANA≤1.5≤0.20
EN ISO 17225-6 category B6 to 253.15 ≤ L ≤40≥96≤3600≤NA≤15≤10≥14.5NANA≤2.0≤0.30
2. Pellets produced from both 100% biomass and biomass blends
Reference: Rezania et al., 2016 [97]
Focal region: MALAYSIA
Production parameters: RM M: 10–15%, RM PS: 0.85 mm, Die P: NA, Die T: NA
Biomass pellets100% Water HyacinthNANANANANANA9.90
(0.00)
6.90
(0.11)
14.58 (0.05)66.27 (0.10)26.83 (0.00)NANA
Blended biomass pellets (selected for meta-analysis)25% Water Hyacinth + 75% Empty Fruit BunchesNANANANANANA9.30
(0.03)
3.73
(0.54)
NA80.3 (1.02)15.97
(3)
NANA
Reference: Scatolino et al., 2018b [102]
Focal region: BRAZIL
Production parameters: RM M: 9–10%, RM PS: 3–4 mm, Die P: NA, Die T: NA
Biomass pellets100% Soybean Waste6.47
(0.13)
17.13
(0.65)
47.49 (1.09)3.32 (0.81)686.00 (0.00)3.87
(2.00)
9.03
(0.50)
26.72 (0.74)16.70 (0.13)62.47 (0.71)10.81 (0.03)NANA
100% Sugarcane Bagasse6.14
(0.06)
18.46
(1.35)
96.64 (0.27)0.18 (0.03)698.00 (0.00)39.46
(7.33)
5.57 (0.16)5.58
(0.44)
17.40 (0.08)80.56 (1.98)13.87 (1.77)NANA
Blended biomass pellets (selected for meta-analysis)50% Sawdust +
50% Soybean Waste
6.17
(0.13)
15.60
(1.70)
67.42 (3.39)1.53 (0.52)6106.13
(2.58)
6.87 (0.12)14.03
(1.4)
17.92 (0.09)71.13 (0.26)14.84 (1.15)NANA
50% Sugarcane Bagasse + 50% Soybean Wastes6.13
(0.11)
15.93 (1.72)67.01 (3.6)2.50 (0.64)6347.75
(3.83)
6.48 (0.17)15.02 (0.71)17.25 (0.12)69.47 (0.60)15.48 (0.26)NANA
Reference: Garcia et al., 2019 [100]
Focal region: BRAZIL
Production parameters: RM M: 15%, RM PS: 4 mm, Die P: NA, Die T: 90 °C
Biomass pellets100% Elephant GrassNANA89.82 (1.40)NA509.80 (8.20)NANA6.80
(0.30)
18.51 (0.26)79.09 (3.48)14.11 (1.10)1.51 (0.08)0.07 (0.01)
100% Sugarcane BagasseNANA87.54 (3.10)NA579.90 (30.60)NANA4.78
(0.06)
18.52 (0.18)79.61 (1.60)15.62 (1.34)1.21 (0.07)0.08 (0.01)
100% Sorghum
NANA93.59 (1.10)NA607.70 (34.50)NANA3.42
(0.12)
19.34 (0.22)78.47 (2.34)18.10 (1.20)1.68 (0.52)0.08 (0.01)
Blended biomass pellets (selected for meta-analysis)95% Sawdust + 5% Charcoal of Eucalyptus spp.NANA92.63 (2.7)NA667.6 (30.1)NANA0.48
(0.07)
20.42 (0.14)78.35 (2.25)21.18 (0.92)0.88 (0.12)0.07 (0.06)
Reference: da Silva et al., 2020 [93]
Focal region: BRAZIL
Production parameters: RM M: 12–16%, RM PS: 3–5 mm, Die P: 29.42 MPa, Die T: 80–95 °C
Biomass pellets100% Elephant Grass
6.20 (0.05)17.07 (1.35)96.58 (0.19)2.36 (0.42)654.10 (3.69)NA8.83 (0.36)6.48
(0.11)
14.84 (0.10)81.20 (0.26)11.75 (0.26)NANA
100% Sugarcane Bagasse6.18 (0.50)12.87 (1.48)92.22 (0.48)44.00 (1.69)574.74 (4.82)NA9.85 (0.14)2.40
(0.01)
15.20 (0.23)84.27 (0.56)13.33 (0.50)NANA
Biomass blended pellets (selected for meta-analysis)50% Elephant Grass + 50% Sawdust6.14 (0.04)17.85 (1.16)98.66 (0.77)2.69 (0.23)690.09 (3.2)NA8.12 (0.12)2.89
(0.01)
15.74 (0.06)86.24 (0.85)10.87 (0.85)NANA
50% Elephant Grass + 50% Sugarcane Bagasse6.13 (0.04)15.01 (1.45)96.18 (0.38)2.55 (0.21)653.51 (1.38)NA8.96 (0.05)4.8
(0.01)
15.09 (0.10)83.31 (1.28)11.89 (1.42)NANA
Reference: Carrillo-Parra et al., 2020 [99]
Focal region: MEXICO
Production parameters: RM M: 15%, RM PS: 3 mm, Die P: NA, Die T: NA
Biomass pellets100% Oil Palm ResidueNANANANA540.00 (170.00)NA6.45 (0.26)0.54
(0.07)
21.89 (0.16)79.40 (1.90)20.05 (1.83)NANA
Blended biomass pellets (selected for meta-analysis)40% Oil Palm Residue + 60% SawdustNANANANA660 (10)NA6.76 (0.89)0.44
(0.11)
20.75 (0.06)87.32 (1.92)12.23 (1.81)NANA
20% Oil Palm Residue + 80% SawdustNANANANA680 (10)NA5.84 (0.19)0.53
(0.07)
19.70 (0.13)85.68 (1.63)13.79 (1.57)NANA
Reference: de Souza et al., 2020 [92]
Focal region: BRAZIL
Production parameters: RM M: 12%, RM PS: 3 mm, Die P: 29.42 MPa, Die T: 80–95 °C
Biomass pellets100% Coffee Husk6.12 (0.02)18.43 (1.17)97.09 (0.30)0.17 (0.07)687.48 (4.16)29.76
(3.11)
9.50 (0.26)9.69
(0.03)
15.76 (0.19)73.15 (0.19)NA3.18 (0.03)0.19 (0.04)
100% Coffee Parchment6.15 (0.02)14.01 (0.73)93.77 (2.96)0.46 (0.16)632.56 (0.73)25.84
(4.21)
8.64 (0.18)2.55
(0.03)
16.96 (0.09)85.24 (0.52)NA1.85 (0.05)0.05 (0.01)
100% Coffee Silver Skin6.11 (0.03)21.02 (2.85)97.10 (2.16)0.10 (0.21)644.36 (6.69)18.32
(2.67)
8.84 (0.21)9.90
(0.03)
16.26 (0.04)75.95 (0.61)NA3.29 (0.04)0.18 (0.02)
Blended biomass pellets (selected for meta-analysis)40% Sawdust + 30% Parchment + 30% Silver Skin6.17 (0.02)13.85 (0.83)93.28 (1.57)0.22 (0.07)634.26 (4.97)24.28
(4.20)
9.79 (0.30)4.00
(0.11)
17.08 (0.14)84.38 (0.20)
NA2.45 (0.05)0.13 (0.02)
40% Sawdust + 30% Parchment + 30% Coffee Husk6.13 (0.08)14.43 (0.99)95.33 (1.19)0.19 (0.05)690.79 (5.47)29.56
(8.01)
8.88 (0.07)6.28
(0.17)
16.51 (0.21)82.37 (0.30)NA1.99 (0.05)0.08 (0.02)
Reference: Szyszlak-Bargłowicz et al., 2021 [13]
Focal region: POLAND
Production parameters: RM M: 12%, RM PS: 0.5–1.0 mm, Die P: NA, Die T: 85 °C
Biomass pellets100% MiscanthusNANA91.40 (1.00)NA567.30 (6.70)NA7.20 (0.05)2.36
(0.14)
16.31 (0.02)73.61 (0.29)16.40 (0.23)0.24 (0.00)0.000 (0.00)
100% Copra MealNANA87.2 (0.00)NA255.10 (0.50)NA5.43 (0.02)5.46
(0.04)
18.38 (0.05)75.62 (0.15)13.49 (0.18)3.15 (0.02)0.13 (0.00)
Blended biomass pellets (selected for meta-analysis)90% Miscanthus + 10% CopraNANA97.20 (0.60)NA514.90 (3.4)NA6.76 (0.03)2.70
(0.04)
17.80 (0.05)74.93 (0.43)15.94 (0.37)0.37 (0.01)0.01 (0.00)
70% Miscanthus + 30% CopraNANA95.10 (0.00)NA417.40 (3.2)NA6.41 (0.02)3.09
(0.54)
17.92 (0.03)74.56 (0.43)15.94 (0.31)0.55 (0.02)0.02 (0.00)
Reference: Harun and Afzal, 2015 [90]
Focal region: CANADA
Production parameters: RM M: 10%, RM PS: 0.1–0.6 mm, Die P: 159 MPa (5 s holding time), Die T: 80 °C
Biomass pellets100% Reed Canary GrassNANA18.61 (0.02)NANANA6.40 (0.00)5.34
(0.45)
NANANA0.17 (0.00)0.04 (0.01)
100% SwitchgrassNANA18.20 (0.17)NANANA7.0 (0.00)3.61
(0.44)
NANANA0.12 (0.00)0.03 (0.00)
100% Timothy HayNANA17.58 (0.06)NANANA6.90 (0.00)4.06
(0.37)
NANANA0.18 (0.00)0.04 (0.01)
Biomass blended pellet (selected for Meta-analysis)50% Spruce Sawdust and 50% RCGNANANANANANA7.22.07
(0.21)
18.56 (0.04)NANA0.04 (0.00)0.02 (0.00)
50% Spruce Sawdust and 50% Timothy HayNANANANANANA7.31.51
(0.18)
18.37 (0.06)NANA0.04 (0.00)0.02 (0.00)
50% Spruce Sawdust and 50% SwitchgrassNANANANANANA7.71.53
(0.5)
18.46 (0.14)NANA0.03
(0.00)
0.01 (0.00)
50% Pine Sawdust + 50% RCGNANANANANANA7.51.54
(0.51)
19.00 (0.16)NANA0.04 (0.00)0.02 (0.00)
50% Pine Sawdust + 50% Timothy HayNANANANANANA7.61.62
(0.05)
18.68 (0.1)NANA0.04 (0.00)0.02 (0.00)
50% Pine Sawdust + 50% SwitchgrassNANANANANANA7.51.55
(0.4)
18.44 (0.08)NANA0.03 (0.00)0.01 (0.00)
3. Pellets produced only from 100% biomass materials
Reference: Tenorio et al., 2016 [103]
Focal region: COSTA RICA
Production parameters: RM M: 5%, RM PS: ≤0.5 mm, Die P: NA, Die T: NA
Biomass pellets100% Oil Palm Empty Fruit Bunches6.09 (2.01)22.94 (9.59)92.76 (0.72)NA575.00 (1.53)NA9.05 (6.87)5.75
(2.69)
14.18 (0.097)71.70 (0.13)NANANA
100% Oil Palm Fruit Mesocarp6.12 (4.08)17.34
(26.68)
92.82 (1.38)NA595.80 (1.62)NA9.2 (3.06)6.24
(0.88)
15.83 (0.06)72.41 (0.78)NANANA
Reference: Almeida et al., 2017 [98]
Focal region: BRAZIL
Production parameters: RM M: Initially, 50% and after drying MC was not mentioned, RM PS: 5 mm, Die P: NA, Die T: 80 °C
Biomass pellets100% Sugarcane Bagasse9.70 (0.10)22.70 (4.94)NANA726.32 (0.62)NA5.49 (0.04)8.70
(0.34)
NA77.27 (2.24)14.03 (0.84)0.28 (0.05)0.02 (0.03)
Reference: Pradhan et al., 2018a [101]
Focal region: INDIA
Production parameters: RM M: 10%, RM PS: 6 mm, Die P: NA, Die T: 80–90 °C
Biomass pellets100% Garden Waste14.70 (0.20)39.20 (5.0)97.70 (0.00)5.90 (0.00)NA24.5
(0.00)
4.50 (1.00)NANANANANANA
Reference: Azócar et al., 2019 [106]
Focal region: CHILE
Production parameters: RM M: 15 ± 2%, RM PS: 0.1–1.2 mm, Die P: NA, Die T: 70 °C
Biomass pelletsWheat Straw6.48 (0.16)22.07
(9.23)
97.23 (0.39)0.33 (0.02)469.00 (8.00)NA9.57 (0.60)2.64
(0.60)
15.43 (0.17)NANA0.33 (0.02)0.00 (0.00)
Pretreated biomass pelletsTorrefied Wheat Straw
(Brown Pellets)
6.28 (0.11)25.38
(9.98)
96.23 (0.39)0.26 (0.06)568.00 (4.00)NA7.12 (0.00)3.19
(0.00)
16.01 (0.09)NANA0.45 (0.05)0.01 (0.00)
Reference: Trejo-Zamudio et al., 2021 [104]
Focal region: MEXICO
Production parameters: RM M: 20%, RM PS: 8 mm, Die P: NA, Die T: 95–105 °C
Biomass pellets100% Bean Crop Residues8.13 (0.01)18.50 (0.16)NANA607.38 (7.69)NA11.67 (0.72)5.32
(0.00)
16.09 (1.98)NANANANA
Reference: (Acampora et al., 2021) [91]
Focal region: ITALY
Production parameters: RM M: 10–20%, RM PS: 6 mm, Die P: NA, Die T: NA
Biomass pellets100% Hazelnut6.20 (0.12)10.47 (2.67)98.00 (0.5)NA581.00 (3.00)NANA3.10
(0.60)
NANANA0.77 (0.21)0.00
(0.00)
100% Olive Tree Pruning Waste6.20 (0.10)16.66 (1.82)98.30 (0.60)NA562.00 (6.00)NANA2.50
(0.10)
NANANA1.24 (0.36)0.00 (0.00)
Reference: Pegoretti Leite de Souza et al., 2021 [94]
Focal region: CHILE
Production parameters: RM M: 5–7%, RM PS: 4 mm, Die P: NA, Die T: 80–100 °C
Biomass pellets100% MiscanthusNANA96.86 (0.07)0.19 (0.01)615.00 (1.5)NA7.42 (0.10)2.94
(0.07)
16.22 (0.02)NANANANA
Biomass pellets100% MiscanthusNANA96.86 (0.07)0.19 (0.01)615.00 (1.5)NA7.42 (0.10)2.94
(0.07)
16.22 (0.02)NANANANA
Reference: Senila et al., 2020 [96]
Focal region: ROMANIA
Production parameters: RM M: 12%, RM PS: 5 mm, Die P: NA, Die T: NA
Biomass pelletsVineyard Waste (VW)20.90 (4.50)10.10 (0.04)97.80 (2.20)1.25 (0.07)657.65 (4.30)NA10.30 (0.40)NA17.35 (1.20)NANA1.23 (0.07)0.02 (0.10)
4. Pellets produced from biomass blends
Reference: Chavalparit et al., 2013 [105]
Focal region: THAILAND
Production parameters: RM M: 20%, RM PS: 2 mm, Die P: NA, Die T: NA
Blended biomass pellets (selected for meta-analysis)55% Oil Palm Frond + 45% Crude GlycerinNANANANA994
(8.4)
NA4.35 (0.07)11.9
(0.1)
20.42.38
(2.6)
2.38
(2.6)
NA81.3 (2.7)
Reference: Amirta et al., 2018 [107]
Focal region: INDONESIA
Production 17.2: RM M: 12%, RM PS: Dust, Die P: NA, Die T: NA
Blended biomass pellet (selected for meta-analysis)70% Sawdust + 20% Tapioca + 20% Glycerol7.7 (0.03)3.03 (0.26)NANA730
(20)
NA7.42 (3.37)4.3
(0.81)
NA89.5 (3.03)6.2 (3.84)NANA
Reference: (Santana et al., 2021) [95]
Focal region: BRAZIL
Production parameters: RM M: 7–11%, RM PS: 5 mm, Die P: NA, Die T: 80–95 °C
Blended biomass pellet (selected for meta-analysis)65% Soybean Waste + 35% Cotton Waste (65SyB + 35CW)6.34 (0.09)18.75 (1.10)92.59 (0.46)0.36 (0.03)NANA9.88 (0.29)NA15.59 (0.04)NANA4.29 (0.22)0.25 (0.01)
65% Soybean Waste + 35% Sorghum Waste (65SyW + 35SoW)6.39 (0.09)18.83 (1.34)87.20 (2.28)0.30 (0.19)NANA4.46 (0.11)NA14.79 (0.49)NANA3.47 (0.15)0.16 (0.03)
65% Soybean Waste + 35% Pine Needles (65SyW + 35PN)6.42 (0.10)17.36 (0.86)76.78 (3.37)0.76 (0.22)NANA6.89 (0.26)NA15.90 (0.14)NANA3.54 (0.10)0.16 (0.01)
65% Rice Powder + 35% Sawdust (65RP + 35SD)6.40 (0.21)18.14 (1.25)94.28 (0.30)0.16 (0.10)NANA7.50 (0.08)NA17.15 (0.11)NANA3.16 (0.16)0.17 (0.00)
65% Rice Powder + 35% Charcoal Fines (65RP + 35CF)6.28 (0.08)18.95 (1.38)97.75 (0.14)0.17 (0.07)NANA10.34 (1.40)NA20.14 (0.22)NANA2.88 (0.03)0.12 (0.01)
NOTE: Standard deviations are shown in brackets, D: diameter of the pellets, L: length of the pellets, MC: moisture content, Ash: ash content, Du: mechanical durability, Fines: fine particle content, BD: bulk density, CV: calorific value (ISO standards NCV and analyzed studies GCV) and, N: nitrogen content, S: sulfur content, VM: volatile matter content, FC: fixed carbon content, RM M: moisture content of the raw materials, RM PS: particle size of the raw materials, Die P: pellet die pressure, Die T: pellet die temperature.
Table 4. Property variation of the Du and BD of the pellets produced in the same conditions in different studies.
Table 4. Property variation of the Du and BD of the pellets produced in the same conditions in different studies.
Main Biomass MaterialAuthor and the
Published Year
MC of the
Biomass
(%)
PS of Biomass (mm)Pelletizing Pressure (MPa)Pelletizing Temperature (°C)Du of the Pellets
(%)
BD of the Pellets (kg/m3)
Elephant Grass
(EG)
da Silva et al., 2020 [93]12–163–529.4280–9596.58654.10
Garcia et al., 2019 [100]9–103–429.4280–9589.82509.80
Miscanthus
(M)
Pegoretti Leite de Souza et al., 2021 [94]5–74NA80–10096.86615.00
Szyszlak-Bargłowicz et al., 2021 [13]120.5–1.0NA8591.40567.30
Sugarcane Bagasse (SB)da Silva et al., 2020 [93]12–163–529.4280–9592.22574.74
Scatolino et al., 2018a [102]154NA9096.64698.00
Garcia et al., 2019 [100]9–103–429.4280–9587.54579.90
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

Rupasinghe, R.L.; Perera, P.; Bandara, R.; Amarasekera, H.; Vlosky, R. Insights into Properties of Biomass Energy Pellets Made from Mixtures of Woody and Non-Woody Biomass: A Meta-Analysis. Energies 2024, 17, 54. https://doi.org/10.3390/en17010054

AMA Style

Rupasinghe RL, Perera P, Bandara R, Amarasekera H, Vlosky R. Insights into Properties of Biomass Energy Pellets Made from Mixtures of Woody and Non-Woody Biomass: A Meta-Analysis. Energies. 2024; 17(1):54. https://doi.org/10.3390/en17010054

Chicago/Turabian Style

Rupasinghe, Rajitha Lakshan, Priyan Perera, Rangika Bandara, Hiran Amarasekera, and Richard Vlosky. 2024. "Insights into Properties of Biomass Energy Pellets Made from Mixtures of Woody and Non-Woody Biomass: A Meta-Analysis" Energies 17, no. 1: 54. https://doi.org/10.3390/en17010054

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