Next Article in Journal
Evaluating the Emissions of the Heat Supplied by District Heating Networks through A Life Cycle Perspective
Next Article in Special Issue
Development of a Low-Cost Experimental Procedure for the Production of Laboratory Samples of Torrefied Biomass
Previous Article in Journal
Benchmarking Wind Farm Projects by Means of Series Two-Stage DEA
Previous Article in Special Issue
Torrefied Biomass as an Alternative in Coal-Fueled Power Plants: A Case Study on Grindability of Agroforestry Waste Forms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Waste Recovery through Thermochemical Conversion Technologies: A Case Study with Several Portuguese Agroforestry By-Products

by
Leonel J. R. Nunes
1,2,3,*,
Liliana M. E. F. Loureiro
4,
Letícia C. R. Sá
4 and
Hugo F. C. Silva
5
1
PROMETHEUS—Unidade de Investigação em Materiais, Energia e Ambiente para a Sustentabilidade, Escola Superior Agrária, Instituto Politécnico de Viana do Castelo, Rua da Escola Industrial e Comercial de Nun’Alvares, 4900-347 Viana do Castelo, Portugal
2
GOVCOPP—Unidade de Investigação em Governança, Competitividade e Políticas Públicas, Universidade de Aveiro, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal
3
DEGEIT—Departamento de Economia, Gestão, Engenharia Industrial e Turismo, Universidade de Aveiro, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal
4
YGE—Yser Green Energy SA, Área de Acolhimento Empresarial de Úl/Loureiro, Lote 17, 3720-075 Loureiro OAZ, Portugal
5
AFS—Advanced Fuel Solutions SA, Área de Acolhimento Empresarial de Úl/Loureiro, Lote 17, 3720-075 Loureiro OAZ, Portugal
*
Author to whom correspondence should be addressed.
Clean Technol. 2020, 2(3), 377-391; https://doi.org/10.3390/cleantechnol2030023
Submission received: 11 June 2020 / Revised: 20 August 2020 / Accepted: 4 September 2020 / Published: 10 September 2020
(This article belongs to the Special Issue Green Process Engineering)

Abstract

:
Agroforestry waste stores a considerable amount of energy that can be used. Portugal has great potential to produce bioenergy. The waste generated during agricultural production and forestry operation processes can be used for energy generation, and it can be used either in the form in which it is collected, or it can be processed using thermochemical conversion technologies, such as torrefaction. This work aimed to characterize the properties of a set of residues from agroforestry activities, namely rice husk, almond husk, kiwi pruning, vine pruning, olive pomace, and pine woodchips. To characterize the different materials, both as-collected and after being subjected to a torrefaction process at 300 °C, thermogravimetric analyses were carried out to determine the moisture content, ash content, fixed carbon content, and the content of volatile substances; elementary analyses were performed to determine the levels of carbon, nitrogen, hydrogen, and oxygen, and the high and low heating values were determined. With these assumptions, it was observed that each form of residual biomass had different characteristics, which are important to know when adapting to conversion technology, and they also had different degrees of efficiency, that is, the amount of energy generated and potentially used when analyzing all factors.

1. Introduction

The use of biomass as an energy source currently presents itself as an alternative to fossil fuels in the generation of different forms of energy, namely in the generation of electrical and thermal energy [1,2,3]. There are several research works that have characterized the main forms of biomass commonly used in energy generation [4,5,6]. However, given the growing demand for these traditional forms of biomass, there has been an increasing pressure on resources, which are beginning to suffer a sharp decrease in availability, leading to the need to find other alternatives, namely those forms of biomass that are considered waste or by-products [7].
Among this group of residues or by-products, it is possible to identify a wide range of materials that, if properly framed and processed, can be presented as alternatives to traditional forms of biomass, both for the characteristics they present and the quantities available [8]. The fact these by-products exist in abundance is also problematic in some situations, mainly from an environmental perspective, since the volume generated from some waste is significant enough that, if an alternative is not found for its use and recovery, simple placement and landfilling does not appear to be a viable solution [9].
In this way, energy recovery is an alternative to landfilling since it can dispose large quantities of waste and by-products that, until now, have had no use (or when they do, it has been the limited use of a small amount of the total produced and, therefore, is not a definitive solution to the problem) [10]. This group of residual biomass forms includes those from the agroforestry industry, and Portugal has a number of examples that can be introduced into the supply chain of units dedicated to the production of energy from biomass [11].
However, from accumulated experience and the information made available by various previous works, it is possible to identify a diverse set of disadvantages and problems associated with the energy recovery of these materials, namely problems related to the low density, low calorific value, high moisture content, and geographic dispersion, but also issues related to extensive contamination with inert materials, high levels of chlorine, or even the presence of high levels of alkali metals [12].
The use of thermochemical conversion technologies, namely torrefaction, presents itself as a potential solution to pre-process these forms of residual biomass since it has shown good results in improving properties related to the combustibility of materials, namely those related to increasing energy density, or to the elimination of chlorine [13]. In the specific case of Portugal, there are several residues and by-products of agricultural and forestry origin. For the present study, six types of residual biomass or by-products were chosen that were readily available in terms of quantity and geographical dispersion. The selected residues or by-products were rice husk, almond husk, kiwi pruning, vine pruning, olive pomace, and pine woodchips.
This work aims to characterize these forms of waste and by-products for energy generation through laboratory tests, namely elemental, thermogravimetric, and energy analyses of materials, both in thermally untreated versions and with the materials submitted to torrefaction at 300 °C. In this way, we intend to effectively compare potential improvements before and after heat treatment and, thus, assess whether waste and by-products have potential use in energy recovery or any other application.

2. State-of-the-Art

Sustainable biomass use for the production of bioenergy has increasingly called for the use of biomass waste, to the detriment of dedicated biomass, which competes with other land uses [14,15,16,17]. However, biomass waste often has a low bulk density, high moisture or ash content, and is mechanically resistant to crushing, which limits the use of the available biomass for energy or material recovery [18,19,20]. Torrefaction is a thermal pre-treatment that can contribute to improving the physical and chemical properties of biomass, facilitating its mechanical processing and increasing its stability and energy density [21,22,23]. In the case of species attacked by pests, torrefaction is also a sterilization process that limits the risk of these pests spreading during storage [24,25,26].
In Portugal, the demand for forest-based biomass in energy production has increased significantly due to incentives for the construction and operation of thermoelectric plants with subsidized tariffs [12,25,27,28]. The recent national growth in the pellet market for heat production in domestic or industrial boilers has resulted in a further increase in demand for biomass [25,29]. However, this growing use of biomass for energy production may lead to its scarcity; therefore, it is necessary to assess the potential use of other types of biomass available in Portugal, namely residual biomass from agricultural activities and food processing (agroindustry) and other woody species and shrubs [16,30,31]. These alternative biomasses have different characteristics due to the quality of woody biomasses, which can significantly influence the supply and pre-treatment chain, combustion processes, ash behavior (slagging, fouling, and corrosion), and environmental constraints associated with energy conversion processes [32,33,34].
The use of solid biofuels for energy production plays an important role in reducing greenhouse gas emissions, as well as in the diversification of energy sources, thereby reducing dependence on external sources [35,36,37]. The increase in research in this area, as well as the development of standards and technical documents, highlight the political, social, and economic relevance of solid biofuels [38,39]. Intensive use of wood-based fuels in co-combustion processes has put enormous pressure on forests [40,41]. To alleviate this pressure and simultaneously increase the use of biomass in co-combustion processes, thereby reducing CO2 emissions, it is necessary to increase the use of alternative residual fuels, namely agricultural waste [29,42]. Co-combustion of these alternative biomasses can cause operational problems due to the presence of alkali metals, chlorine, and other elements in the ashes of these materials, and it can enhance the corrosion of metallic surfaces and the emission of small particles [43,44]. This can limit the variety of residual biomass that can actually be used in co-combustion processes [32,38,45].
From an economic perspective, the use of residues of agroforestry origin has a set of very significant advantages, since they constitute an alternative to the use of other types of biomass that already have other uses [46,47,48]. In this way, when other supply chains are created, competition between traditional industries is avoided, such as pulp and paper or the production of wood pellets, with the production of energy [49,50,51]. This non-competition allows, first, the use of resources that until now have not been used, and therefore, with low market value, its cost depending only on the complexity of its supply chain; and, second, the creation of a truly sustainable option, in a perspective of circular economy, where all resources are valued, to the detriment of the intensive exploitation of forest resources [52,53,54].

3. Materials and Methods

3.1. Sample Collection and Preparation

To verify the evolution of the properties of the residues and by-products resulting from the torrefaction process, several analyses were carried out. Thus, different techniques were used such as thermogravimetric (fixed carbon content, volatiles, ash, and moisture), elemental (C, H, O, and N), and heating value analyses.
Samples of rice husk, almond shells, kiwi pruning, vine pruning, olive pomace, and pine woodchips were collected and, when necessary, cut to obtain a granulometry that would allow the materials to be efficiently processed. These materials were selected because they are representative of types of waste and by-products from the agroforestry industry. They exist in great abundance and are, therefore, available to potentially work as alternatives if they are found to have physicochemical characteristics compatible with efficient energy recovery. After collection, the samples were dried in a laboratory oven at 90 °C for 24 h.
Samples were prepared by weighing approximately 250 g of the material. Preparations were made with the aid of conventional aluminum foil in which the objective was to wrap the samples in a cylindrical shape. Since aluminum foil has two distinct sides, it should be noted that the opaque part of the sheet was directed towards the outside so that during the torrefaction process the heat was not reflected.
In this work, the torrefaction protocol used was the one described by Ribeiro et al. (2018) [55], in which a high-temperature oven or ceramic muffle was used. The oven was composed of a metallic monobloc, covered inside by refractory bricks, and had a kaolin canvas for insulation. The muffle was heated by the heat transfer produced from the electrical resistances found inside. The samples were torrefied in a muffle, which had a built-in controller and the ability to program the residence time and four temperature thresholds. There was a cavity on top of the equipment in order to facilitate extraction of the torrefaction gases. The muffle was then programmed according to the desired temperature (°C) and residence time (minutes), with each level corresponding to different torrefaction phases, as presented in Table 1.
Table 2 defines the parameters used for each series of torrefaction tests, and the materials obtained in the torrefaction process are shown in Figure 1.

3.2. Ultimate Analysis

To determine the elemental composition of the samples, an elemental carbon, hydrogen, and nitrogen analyzer was used. Its operation consisted of incinerating the samples at 900 °C in an atmosphere rich in oxygen so that all organic compounds were burnt using CO2, H2O, N2, and SO2 as final products. Subsequently, through a gas chromatography detector, the levels of carbon, hydrogen, and nitrogen were obtained. The calibration line used for this procedure was obtained from the analysis of barley (barley sample calibration method for CHNS, LECO®) with known concentrations of carbon, hydrogen, and nitrogen. After obtaining the results, the oxygen content of the samples was then estimated based on Equation (1):
w(O) = 100 − w(C) − w(H) − w(N),
where w(O) is the oxygen content (%), w(C) is the carbon content (%), w(H) is the hydrogen content (%), and w(N) is the nitrogen content (%).

3.3. Proximate Analysis

Thermogravimetric analysis (TGA) is a method used to analyze the loss of mass of a given sample of organic, inorganic, or synthetic origin, depending on the evolution of temperature in a controlled atmosphere defined by the user [56]. For the TGA analysis, the equipment used consisted of an oven with a precision scale inside where melting pots for measuring the samples were inserted. The analyses were conducted in an atmosphere with a nitrogen flow of 150 mL/min and a heating rate of 50 °C/min, from room temperature to 900 °C, in accordance with the standards EN 14775:2009, Solid Biofuels—Determination of Ash Content; EN 15148:2009, Solid Biofuels—Determination of Volatiles Content; and EN 14774-3:2009, Solid Biofuel—Determination of Moisture Content. During the heating process, the moisture content, volatile content, and fixed carbon content were determined precisely in this sequential order. Finally, the ash content was determined from the final residue present in the sample. This procedure required the samples to be previously ground before being introduced into the melting pots. The grinding process was carried out for short periods of time of 8 to 10 s. When starting the process, the melting pots were inserted into the equipment, where 1 g of sample was introduced in each of the melting pots. An empty reference melting pot was used as a blank sample.

3.4. Determination of Heating Value

Moran and Shapiro (2002) objectively defined the heating value of a fuel as a positive number equal to its combustion enthalpy module, and this can be further defined as the high heating value (HHV) and the low heating value (LHV) [57]. The HHV is obtained when all the water formed in the combustion is liquid, and the LHV is obtained when all the water formed in the combustion is steam; the difference between these heating values is equivalent to the energy needed to vaporize the water formed in the combustion [58]. These values have great relevance in designing combustion equipment, and the measurement of HHV is important to characterize a fuel. For solid fuels with certain previously known characteristics, the HHV and LHV can be calculated with specific equations, using both elemental and thermogravimetric analyses [59]. In the literature, the HHV is commonly calculated with an equation according to the elemental analysis of biomass fuels, as variations of the main elements such as carbon, hydrogen, and oxygen are relatively small for any biomass. Observing this information, it is possible to use an equation to determine the HHV as long as the elemental analysis is known [60]. Large variations in the heating value of a biomass are caused by its moisture and ash content. Channiwala and Parikh (2002) presented a universal equation (Equation (2)) for calculating the HHV of several fuels [61]:
HHV = 0.3491 × C + 1.1783 − 0.1034 × O − 0.0151 × N − 0.0211 × Ashes + 0.1005 × S,
whose validity range is 0% ≤ C ≤ 92.25%; 0.43% ≤ H ≤ 25.15%; 0.00% ≤ O ≤ 50.00%; 0.00% ≤ n ≤ 5.60%; 0.00% ≤ S ≤ 94.08; 0.00% ≤ Ashes ≤ 71.4%; and 4.75 MJ/kg ≤ HHV ≤ 55.35 MJ/kg. Normally, for solid fuels, the value determined in the laboratory is the HHV of the dry material. The LHV of the dry material is calculated from the HHV and the elemental analysis, where the enthalpy of water vaporized during combustion is discounted [62]. The formula for calculating the LHV is shown as follows (Equation (3)) [61]:
LHV = HHV m H 2 O × Δ H H 2 O vap   ( 25   ) ,
where Δ H H 2 O vap   ( 25   ) is the enthalpy of water vaporization at 25 °C.
The mass of water formed during combustion is calculated using the following expression (Equation (4)):
m H 2 O = 9 × H
where H is the hydrogen content in the dry solid fuel, determined from the elementary analysis presented above [58,61].

4. Results

4.1. Elemental Analysis

The results obtained from the elemental composition analysis are shown in Table 3.
From the results, it can be seen that the carbon content increases in all samples analyzed, varying for dry material from 48.30%, from vine pruning, up to a maximum of 58.40%, from almond shell. The dry samples of rice husk, kiwi pruning, olive pomace, and pine woodchips showed values, respectively, of 48.60%, 49.40%, 56.20%, and 50.20%. After the torrefaction test, the carbon content rose in all samples, with gains between 8% for rice husk and 60% for pine woodchips. The samples of almond shells, kiwi pruning, vine pruning, and olive pomace, showed gains, respectively, of 37%, 53%, 52%, and 45%. Hydrogen and oxygen showed a decrease in the levels quantified in the dry samples, for the roasted samples. Thus, there are decreases of, respectively, 34%, 28%, 36%, 33%, 46%, and 33%, for hydrogen contents. For oxygen levels, there are decreases of 6%, 56%, 57%, 53%, 63%, and 64%, respectively. For nitrogen levels, there are percentage increases of 45%, 94%, 186%, 76%, 29%, and 127%, respectively.

4.2. Thermogravimetric Analysis

The results obtained from the proximate composition analysis are shown in Table 4.
Fixed carbon showed a percentage increase from the dry to the torrefied state, in all samples analyzed, being, respectively, 157%, 255%, 251%, 217%, 292%, and 310%. On the other hand, moisture shows a decrease of 62%, 85%, 73%, 75%, 36%, and 28%, respectively. Ashes have a contrary trend, clearly increasing, and are related to the loss of mass, as will be analyzed in more detail, with a percentage increase in the content of, respectively, 109%, 154%, 281%, 176%, 248%, and 248%. On the other hand, the volatile content decreases, with decreasing values of 67%, 71%, 67%, 63%, 73%, and 68%, respectively.

4.3. Heating Value

The results obtained from the high heating value calculations are shown in Table 5.
The values of HHV and LHV, with the exception of the values corresponding to the rice husk samples, which suffered a decrease from the dry to the torrefied state, showed a significant increase, which reached its maximum value in the samples of pine woodchips, with a 55% gain for HHV. Mass losses were significant in all samples, with the lowest value corresponding to the rice husk sample, with 56.90%, and the highest value corresponding to the olive pomace sample, with a mass loss of 71.10%.

5. Discussion

The fixed carbon content present in a given fuel is directly related to the amount of energy it is able to supply [63]. The higher this content, the slower the fuel will burn, and this can guarantee total combustion and more efficient energy recovery [64,65,66]. The values of fine carbon content for the six materials studied, even without thermal treatment at 300 °C, were in the range between 17.55% for rice husks and 20.73% for almond shells. For the remaining materials, the fixed carbon contents were around an average of 19.06 ± 0.83%. The values obtained were in line with those presented in previous studies, namely by Kalderis et al. (2014) for rice husk [67], Demirbaş (2002) for almond husk [68], Torreiro et al. (2020) for kiwi pruning [69], San José et al. (2013) for vine pruning [70], Nunes et al. (2020) for olive pomace [71], and Nunes et al. (2019) for pine chips [72]. For the same materials, after being subjected to the heat treatment process by torrefaction, there was an expected increase in the fixed carbon content as described in several previous works for biomass materials, namely the works of Faizal et al. (2018), Lau et al. (2018), and Conag et al. (2017) [73,74,75]. The smallest percentage increase corresponded to rice husk, which stood at 45.10%, a value 25% lower than the average value obtained for the remaining materials, which was 70.51 ± 3.96%. These values were also in line with those obtained in previous studies, namely Chen et al. (2020) for rice husk [76], Chiou et al. (2018) for almond shell [77], Margaritis et al. (2020) for vine pruning [78], Volpe et al. (2015) for olive pomace [79], and Phanphanich and Mani (2011) for pine chips [80]. This significant difference in fixed carbon content for rice husk is certainly related to the very high ash content, which was also demonstrated during product characterization. This ash content, as also demonstrated in the works by Pode (2016), Prasara and Gheewala (2017), Raisi et al. (2018), or da Silva et al. (2019), results from the amount of silica that accumulates during the plant’s growth in wet soils with a deeply clayey and silty nature [81,82,83,84]. This factor is almost entirely eliminated when using rice husk for the purposes of energy recovery. However, this can be of great interest if, in association with energy recovery, the recovery of silica is the main objective, for example, in the production of abrasive products. In the case of kiwi pruning, the only works found in the literature refer to its characterization as a fuel after drying, without any other type of thermal processing such as torrefaction, including works developed by Dyjakon and García-Galindo (2019) or by Boumancher et al. (2019) [85,86]. With reference to works carried out on the application of thermochemical conversion technologies, in this case pyrolysis, Rene et al. (2020) studied the production of biochar from pruning kiwi, employed as an amendment, aiming to evaluate its remediation potential on smelter and mining contaminated soils [87]; however, the data from this study were not used for comparison because the methodological assumptions related to sample preparation were significantly different from those used in the present work. In this way, it can be said quite safely that the characterization of material from pruning kiwi orchards for energy purposes, after pre-treatment using torrefaction, is novel in this work. In the case of Portugal, this may contribute to the creation of a value chain for waste produced annually in large quantities, since the production of kiwi is growing given the excellent edaphoclimatic conditions for this species [88].
When the biomass has a high moisture content, the combustion process is less efficient when compared to a material with a lower moisture content. This is because the higher the moisture content present in the biomass, the greater the energy needed to start the combustion process; in other terms, if more energy is needed to vaporize the water present in the fuel, then less energy will be provided for the endothermic reaction responsible for maintaining combustion [89,90,91]. Several authors reported that the presence of moisture makes combustion difficult, as the calorific value is reduced, and this increases fuel consumption [92,93]. Other studies also claimed that high humidity generates environmental pollution due to an increase in the volume of combustion products and particulate material, not to mention the corrosion process is accelerated in the final part of the steam generator, and particles accumulate on the heating surfaces [94,95]. In the case of converting biomass into fuel, or as an intermediate pre-processing step specifically for use in gasification processes, Lucena Tavares and dos Santos (2013) stated that a high moisture content does not lead to technical difficulties in gasification, but instead a reduction in the efficiency of the process, since the energy needed to evaporate water and maintain the operating temperature is obtained by increasing the fuel and oxidants [96]. In short, humidity is a limiting factor when choosing fuel, and fuels with values above 50% should not be used because there is not enough energy released to ensure the maintenance of combustion and, consequently, the production of heat [97]. In the present work, all analyzed materials were submitted to a drying process in the laboratory immediately after collection in order to reduce the initial moisture content, which varied between 25% and 60% depending on the material. After this period of forced drying, the thermogravimetric analysis showed values between 2.61% for pine chips and 11.55% for rice husk. After the torrefaction test, there was a profound reduction in the moisture content of the products, with almond shells showing the lowest value at 1.26%, while rice shells showed a value of 4.37%. These values are clearly low and indicate a high potential for the combustion of this waste. These results are in line with those presented in previous studies, such as those carried out by Bazargan et al. (2014) with rice husks, Demirbaş (2002) with almond husks, Manzone et al. (2019) with kiwi pruning, San José et al. (2013) with vine pruning, Nunes et al. (2020) with olive pomace, and Nunes et al. (2019) with pine chips, and all works indicated a good potential for energy recovery [68,70,71,72,98,99].
Several authors, such as Nunes et al. (2016) or Kalembkiewicz and Chmielarz (2012), found that a high ash content led to a decrease in efficiency due to both the increased consumption of oxygen to melt the ash as well as the loss of heat from ash leaving the reactor, which cannot be fully recovered [100,101]. Other studies also reported that, for use in pre-processing for later gasification, a lower ash content reduced the likelihood of equipment getting clogged and encrusted [102,103]. As with humidity, the ash content also interferes with the calorific value, which causes energy losses in addition to impairing heat transfer [104,105]. Several authors agreed on removing ash from the combustion site, as ash also worsens problems related to the corrosion of the metal equipment, especially if there is a high ash content or high content of alkali metals [106,107]. In the products analyzed in the present work, only the rice husk showed values much higher than acceptable for use in combustion processes: 15.86% and 33.19%, respectively, for samples before and after the torrefaction process. This increase is related to the loss of mass, which in this specific case was 56.90%. The remaining materials presented values in accordance with those determined by previous works, with values ranging from 0.27% for pine chips to 2.89% for vine pruning. After torrefaction, the values fluctuated between 0.94% for pine chips and 7.89% for wine pruning, thus maintaining compliance due to mass losses.
When biomass has a high volatile content, it is easier to start and maintain combustion. For this reason, the combustion process may be faster or become difficult to control, leading to greater fuel consumption; thus, a high content of volatile materials can also affect the combustion process in general [108]. The volatile contents in the present study showed great uniformity. Products not thermally processed presented an average value around 77.04 ± 4.93%, while products submitted to the torrefaction process presented an average value of 24.44 ± 2.66%. This is in agreement with the works of García et al. (2012) or by Sun et al. (2017), where several residual biomasses without thermal treatment are characterized [6,109], or with the work of Singh and Zondlo (2017), where samples of torrefied residual biomass are treated [110].
The results obtained in the elementary analysis differed slightly and insignificantly from other studies [111,112,113,114,115]. These differences are related to the fact that the elementary analysis could have been determined using different methodologies from the one used in the present work. Another reason may be related to the different geographic origins of residual biomass forms used in this characterization, or different soils, a fact that alters the chemical composition of the residues and, consequently, the results [116]. The present work showed, however, a common trend to what would be expected and described in the literature, that is, there was an increase in the levels of carbon and nitrogen in all analyzed materials, and this was consistent with the concentration of the components not eliminated during the process, usually oxygenated and hydrogenated compounds, which lowered the content in all materials after thermal processing by torrefaction.
Regarding the low heating value, there was an increase in line with what was expected, according to previous experiences for all materials, with the exception of rice husk, which showed a very small increase in LHV [32,117,118,119,120,121]. This situation is due to the high ash and mineral content in the rice husk, which will in no way contribute to increasing the energy density after the torrefaction process, a fact that is in line with the verified loss of mass, which was only 56.90%, while for the remaining materials it was 67.62 ± 2.66%. All low heating values indicated a good potential for energetic valorization of the analyzed materials.

6. Conclusions

The growing demand for alternative forms of energy has led to biomass becoming increasingly used an alternative to fossil fuels. However, different forms of biomass have a number of disadvantages when used as fuels, mainly due to factors such as a lower heating value, low density, high moisture content, or geographic dispersion. In this way, thermochemical conversion technologies, such as torrefaction, can be used as an alternative to improve the properties of these materials, especially if residual forms of biomass are used.
Residual forms of biomass are interesting, as they allow new flows to be incorporated in the supply chain for energy recovery. The agroforestry sector may contribute to this supply chain with various residues and by-products, of which the materials used in this study can be used as an example: rice husks, almond husks, kiwi pruning, vine pruning, olive pomace, and pine woodchips.
We showed an improvement in the properties of the materials after being subjected to the torrefaction process at 300 °C, mainly with regard to the calorific value, reduction in volatile content, and increase in the fixed carbon content. However, in all cases, there was an increase in ash content associated with a loss of mass. In the specific case of rice husks, because they are a form of biomass that, even without thermal treatment, already has a high ash content, the increase in ash after torrefaction renders it unusable as a fuel, except if the ash is extracted and the combustion systems are frequently cleaned, which would help to deter ash encrustation and vitrification.
Despite the viable uses of biomass for energy production, further tests are needed to confirm this potential, including chemical characterizations of different types of waste, namely with regard to the levels of halogens, such as chlorine, or alkali metals, such as sodium and potassium, since these wastes are mainly responsible for the corrosion of combustion equipment, fouling, or slagging.

Author Contributions

Conceptualization, L.J.R.N.; methodology, L.J.R.N. and H.F.C.S.; validation, L.J.R.N., L.C.R.S. and L.M.E.F.L.; formal analysis, L.J.R.N.; investigation, L.J.R.N., L.C.R.S., L.M.E.F.L.; resources, H.F.C.S.; data curation, L.J.R.N., L.C.R.S. and L.M.E.F.L.; writing—original draft preparation, L.J.R.N.; writing—review and editing, L.J.R.N. and H.F.C.S.; supervision, L.J.R.N.; funding acquisition, H.F.C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to acknowledge the Portuguese companies YGE—Yser Green Energy SA and AFS—Advanced Fuel Solutions SA, both in Portugal, that allowed the execution of the laboratory tests.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Madanayake, B.N.; Gan, S.; Eastwick, C.N.; Ng, H.K. Biomass as an energy source in coal co-firing and its feasibility enhancement via pre-treatment techniques. Fuel Process. Technol. 2017, 159, 287–305. [Google Scholar] [CrossRef]
  2. Bhattacharya, S. State-of-the-art of utilizing residues and other types of biomass as an energy source. Int. Energy J. 2017, 15, 1–21. [Google Scholar]
  3. Gejguš, M.; Aschbacher, C.; Sablik, J. Potential for the Use of Biomass as a Prospective Renewable Energy Source. Res. Pap. Fac. Mater. Sci. Technol. Slovak Univ. Technol. 2017, 25, 27–30. [Google Scholar] [CrossRef] [Green Version]
  4. Abdulrahman, A.O.; Huisingh, D. The role of biomass as a cleaner energy source in Egypt’s energy mix. J. Clean. Prod. 2018, 172, 3918–3930. [Google Scholar] [CrossRef]
  5. Naik, S.; Goud, V.V.; Rout, P.K.; Jacobson, K.; Dalai, A.K. Characterization of Canadian biomass for alternative renewable biofuel. Renew. Energy 2010, 35, 1624–1631. [Google Scholar] [CrossRef]
  6. Garcia, R.; Pizarro, C.; Lavín, A.G.; Bueno, J.L. Characterization of Spanish biomass wastes for energy use. Bioresour. Technol. 2012, 103, 249–258. [Google Scholar] [CrossRef] [PubMed]
  7. Gumisiriza, R.; Hawumba, J.F.; Okure, M.; Hensel, O. Biomass waste-to-energy valorisation technologies: A review case for banana processing in Uganda. Biotechnol. Biofuels 2017, 10, 11. [Google Scholar] [CrossRef] [Green Version]
  8. Tripathi, N.; Hills, C.D.; Singh, R.S.; Atkinson, C. Biomass waste utilisation in low-carbon products: Harnessing a major potential resource. NPJ Clim. Atmos. Sci. 2019, 2, 1–10. [Google Scholar] [CrossRef] [Green Version]
  9. Toklu, E. Biomass energy potential and utilization in Turkey. Renew. Energy 2017, 107, 235–244. [Google Scholar] [CrossRef]
  10. Torrisi, S.; Anastasi, E.; Longhitano, S.; Longo, I.C.; Zerbo, A.; Borzì, G. Circular economy and the benefits of biomass as a renewable energy source. Procedia Environ. Sci. Eng. Manag. 2018, 5, 175–181. [Google Scholar]
  11. Ferreira, S.; Monteiro, E.; Brito, P.S.; Vilarinho, C. Biomass resources in Portugal: Current status and prospects. Renew. Sustain. Energy Rev. 2017, 78, 1221–1235. [Google Scholar] [CrossRef]
  12. Nunes, L.J.; Matias, J.C.O.; Catalão, J. Biomass in the generation of electricity in Portugal: A review. Renew. Sustain. Energy Rev. 2017, 71, 373–378. [Google Scholar] [CrossRef]
  13. Bach, Q.-V.; Skreiberg, Ø. Upgrading biomass fuels via wet torrefaction: A review and comparison with dry torrefaction. Renew. Sustain. Energy Rev. 2016, 54, 665–677. [Google Scholar] [CrossRef]
  14. Elum, Z.A.; Modise, D.; Nhamo, G. Climate change mitigation: The potential of agriculture as a renewable energy source in Nigeria. Environ. Sci. Pollut. Res. 2016, 24, 3260–3273. [Google Scholar] [CrossRef] [PubMed]
  15. Cherubini, F. The biorefinery concept: Using biomass instead of oil for producing energy and chemicals. Energy Convers. Manag. 2010, 51, 1412–1421. [Google Scholar] [CrossRef]
  16. Pedroli, B.; Elbersen, B.; Frederiksen, P.; Grandin, U.; Heikkilä, R.; Krogh, P.H.; Izakovičová, Z.; Johansen, A.; Meiresonne, L.; Spijker, J. Is energy cropping in Europe compatible with biodiversity?—Opportunities and threats to biodiversity from land-based production of biomass for bioenergy purposes. Biomass Bioenergy 2013, 55, 73–86. [Google Scholar] [CrossRef]
  17. Mehmood, M.A.; Ibrahim, M.; Rashid, U.; Nawaz, M.; Ali, S.; Hussain, A.; Gull, M. Biomass production for bioenergy using marginal lands. Sustain. Prod. Consum. 2017, 9, 3–21. [Google Scholar] [CrossRef]
  18. Creutzig, F.; Ravindranath, N.H.; Berndes, G.; Bolwig, S.; Bright, R.; Cherubini, F.; Chum, H.; Corbera, E.; Delucchi, M.; Faaij, A.; et al. Bioenergy and climate change mitigation: An assessment. GCB Bioenergy 2014, 7, 916–944. [Google Scholar] [CrossRef] [Green Version]
  19. Khan, A.; De Jong, W.; Jansens, P.; Spliethoff, H. Biomass combustion in fluidized bed boilers: Potential problems and remedies. Fuel Process. Technol. 2009, 90, 21–50. [Google Scholar] [CrossRef]
  20. Kang, K.; Qiu, L.; Sun, G.; Zhu, M.; Yang, X.; Yao, Y.; Sun, R. Codensification technology as a critical strategy for energy recovery from biomass and other resources—A review. Renew. Sustain. Energy Rev. 2019, 116, 109414. [Google Scholar] [CrossRef]
  21. Tumuluru, J.S.; Sokhansanj, S.; Hess, J.R.; Wright, C.T.; Boardman, R.D. REVIEW: A review on biomass torrefaction process and product properties for energy applications. Ind. Biotechnol. 2011, 7, 384–401. [Google Scholar] [CrossRef] [Green Version]
  22. Nunes, L.J.; Matias, J.C.O.; Catalão, J. A review on torrefied biomass pellets as a sustainable alternative to coal in power generation. Renew. Sustain. Energy Rev. 2014, 40, 153–160. [Google Scholar] [CrossRef]
  23. Wigley, T.; Yip, A.C.; Pang, S. The use of demineralisation and torrefaction to improve the properties of biomass intended as a feedstock for fast pyrolysis. J. Anal. Appl. Pyrolysis 2015, 113, 296–306. [Google Scholar] [CrossRef]
  24. 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]
  25. Nunes, L.J. Torrefied Biomass as an Alternative in Coal-Fueled Power Plants: A Case Study on Grindability of Agroforestry Waste Forms. Clean Technol. 2020, 2, 270–290. [Google Scholar] [CrossRef]
  26. Bichot, A.; Delgenès, J.-P.; Méchin, V.; Carrere, H.; Bernet, N.; García-Bernet, D. Understanding biomass recalcitrance in grasses for their efficient utilization as biorefinery feedstock. Rev. Environ. Sci. Bio-Technol. 2018, 17, 707–748. [Google Scholar] [CrossRef]
  27. Proskurina, S.; Sikkema, R.; Heinimö, J.; Vakkilainen, E. Five years left—How are the EU member states contributing to the 20% target for EU’s renewable energy consumption; The role of woody biomass. Biomass Bioenergy 2016, 95, 64–77. [Google Scholar] [CrossRef]
  28. Sikkema, R.; Steiner, M.; Junginger, M.; Hiegl, W.; Hansen, M.T.; Faaij, A.P.C. The European wood pellet markets: Current status and prospects for 2020. Biofuels Bioprod. Biorefining 2011, 5, 250–278. [Google Scholar] [CrossRef]
  29. Nunes, L.J.; Matias, J.C.O.; Catalão, J. Wood pellets as a sustainable energy alternative in Portugal. Renew. Energy 2016, 85, 1011–1016. [Google Scholar] [CrossRef]
  30. Nunes, L.J.; Meireles, C.I.R.; Gomes, C.J.P.; Ribeiro, N.M.C.D.A. Socioeconomic Aspects of the Forests in Portugal: Recent Evolution and Perspectives of Sustainability of the Resource. Forests 2019, 10, 361. [Google Scholar] [CrossRef] [Green Version]
  31. Ericsson, K.; Nilsson, L.J. Assessment of the potential biomass supply in Europe using a resource-focused approach. Biomass Bioenergy 2006, 30, 1–15. [Google Scholar] [CrossRef] [Green Version]
  32. Rodrigues, A.; Loureiro, L.; Nunes, L.J. Torrefaction of woody biomasses from poplar SRC and Portuguese roundwood: Properties of torrefied products. Biomass Bioenergy 2018, 108, 55–65. [Google Scholar] [CrossRef]
  33. Nunes, L.J.; Raposo, M.A.M.; Meireles, C.I.R.; Gomes, C.J.P.; Ribeiro, N.M.C.A. Control of Invasive Forest Species through the Creation of a Value Chain: Acacia dealbata Biomass Recovery. Environments 2020, 7, 39. [Google Scholar] [CrossRef]
  34. Sahu, S.; Chakraborty, N.; Sarkar, P. Coal–biomass co-combustion: An overview. Renew. Sustain. Energy Rev. 2014, 39, 575–586. [Google Scholar] [CrossRef]
  35. Nunes, L.J.; Matias, J.C.O.; Catalão, J. Analysis of the use of biomass as an energy alternative for the Portuguese textile dyeing industry. Energy 2015, 84, 503–508. [Google Scholar] [CrossRef]
  36. Demirbas, A. Political, economic and environmental impacts of biofuels: A review. Appl. Energy 2009, 86, S108–S117. [Google Scholar] [CrossRef]
  37. Amponsah, N.Y.; Troldborg, M.; Kington, B.; Aalders, I.; Hough, R.L. Greenhouse gas emissions from renewable energy sources: A review of lifecycle considerations. Renew. Sustain. Energy Rev. 2014, 39, 461–475. [Google Scholar] [CrossRef]
  38. Rodrigues, A.; Nunes, L.J. Evaluation of ash composition and deposition tendencies of biomasses and torrefied products from woody and shrubby feedstocks: SRC poplar clones and common broom. Fuel 2020, 269, 117454. [Google Scholar] [CrossRef]
  39. Sepulveda, N.; Jenkins, J.; De Sisternes, F.J.; Lester, R.K. The Role of Firm Low-Carbon Electricity Resources in Deep Decarbonization of Power Generation. Joule 2018, 2, 2403–2420. [Google Scholar] [CrossRef] [Green Version]
  40. Nunes, L.J.; Matias, J.C.O.; Catalão, J. Biomass waste co-firing with coal applied to the Sines Thermal Power Plant in Portugal. Fuel 2014, 132, 153–157. [Google Scholar] [CrossRef] [Green Version]
  41. Purkus, A.; Hagemann, N.; Bedtke, N.; Gawel, E. Towards a sustainable innovation system for the German wood-based bioeconomy: Implications for policy design. J. Clean. Prod. 2018, 172, 3955–3968. [Google Scholar] [CrossRef]
  42. Monte, M.C.; Fuente, E.; Blanco, A.; Negro, C. Waste management from pulp and paper production in the European Union. Waste Manag. 2009, 29, 293–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Niu, Y.; Tan, H.; Hui, S. Ash-related issues during biomass combustion: Alkali-induced slagging, silicate melt-induced slagging (ash fusion), agglomeration, corrosion, ash utilization, and related countermeasures. Prog. Energy Combust. Sci. 2016, 52, 1–61. [Google Scholar] [CrossRef]
  44. Demirbas, A. Potential applications of renewable energy sources, biomass combustion problems in boiler power systems and combustion related environmental issues. Prog. Energy Combust. Sci. 2005, 31, 171–192. [Google Scholar] [CrossRef]
  45. Ribeiro, J.P.; Vicente, E.D.; Gomes, A.P.; Nunes, M.; Alves, C.; Tarelho, L. Effect of industrial and domestic ash from biomass combustion, and spent coffee grounds, on soil fertility and plant growth: Experiments at field conditions. Environ. Sci. Pollut. Res. 2017, 24, 15270–15277. [Google Scholar] [CrossRef]
  46. Keoleian, G.A.; Volk, T.A. Renewable Energy from Willow Biomass Crops: Life Cycle Energy, Environmental and Economic Performance. Crit. Rev. Plant. Sci. 2005, 24, 385–406. [Google Scholar] [CrossRef]
  47. Langeveld, J.W.A.; Dixon, J.; Jaworski, J.F. Development Perspectives of the Biobased Economy: A Review. Crop. Sci. 2010, 50, S142–S151. [Google Scholar] [CrossRef]
  48. Paredes-Sánchez, J.P.; López-Ochoa, L.M.; López, L.M.; Las-Heras-Casas, J.; Xiberta-Bernat, J. Evolution and perspectives of the bioenergy applications in Spain. J. Clean. Prod. 2019, 213, 553–568. [Google Scholar] [CrossRef]
  49. Pan, S.-Y.; Du, M.A.; Huang, I.-T.; Liu, I.-H.; Chang, E.; Chiang, P.-C. Strategies on implementation of waste-to-energy (WTE) supply chain for circular economy system: A review. J. Clean. Prod. 2015, 108, 409–421. [Google Scholar] [CrossRef]
  50. Cambero, C.; Sowlati, T. Assessment and optimization of forest biomass supply chains from economic, social and environmental perspectives—A review of literature. Renew. Sustain. Energy Rev. 2014, 36, 62–73. [Google Scholar] [CrossRef]
  51. Iakovou, E.; Karagiannidis, A.; Vlachos, D.; Toka, A.; Malamakis, A. Waste biomass-to-energy supply chain management: A critical synthesis. Waste Manag. 2010, 30, 1860–1870. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, B.; Rajagopal, D. Life-cycle energy and climate benefits of energy recovery from wastes and biomass residues in the United States. Nat. Energy 2019, 4, 700–708. [Google Scholar] [CrossRef] [Green Version]
  53. Pippo, W.A.; Luengo, C.A.; Alberteris, L.A.M.; Garzone, P.; Cornacchia, G. Energy Recovery from Sugarcane-Trash in the Light of 2nd Generation Biofuel. Part 2: Socio-Economic Aspects and Techno-Economic Analysis. Waste Biomass Valorizat. 2011, 2, 257–266. [Google Scholar] [CrossRef]
  54. Kliopova, I.; Makarskienė, K. Improving material and energy recovery from the sewage sludge and biomass residues. Waste Manag. 2015, 36, 269–276. [Google Scholar] [CrossRef]
  55. Godina, R.; Godina, R.; Matias, J.C.O.; Nunes, L.J. Future Perspectives of Biomass Torrefaction: Review of the Current State-Of-The-Art and Research Development. Sustainability 2018, 10, 2323. [Google Scholar] [CrossRef] [Green Version]
  56. Materazzi, S. Mass Spectrometry Coupled to Thermogravimetry (TG-MS) for Evolved Gas Characterization: A Review. Appl. Spectrosc. Rev. 1998, 33, 189–218. [Google Scholar] [CrossRef]
  57. Baratieri, M.; Baggio, P.; Fiori, L.; Grigiante, M. Biomass as an energy source: Thermodynamic constraints on the performance of the conversion process. Bioresour. Technol. 2008, 99, 7063–7073. [Google Scholar] [CrossRef]
  58. Friedl, A.; Padouvas, E.; Rotter, H.; Varmuza, K. Prediction of heating values of biomass fuel from elemental composition. Anal. Chim. Acta 2005, 544, 191–198. [Google Scholar] [CrossRef]
  59. Yin, C.-Y. Prediction of higher heating values of biomass from proximate and ultimate analyses. Fuel 2011, 90, 1128–1132. [Google Scholar] [CrossRef] [Green Version]
  60. Annamalai, K.; Sweeten, J.M.; Ramalingam, S.C. Technical Notes: Estimation of Gross Heating Values of Biomass Fuels. Trans. ASAE 1987, 30, 1205–1208. [Google Scholar] [CrossRef]
  61. Channiwala, S.; Parikh, P. A unified correlation for estimating HHV of solid, liquid and gaseous fuels. Fuel 2002, 81, 1051–1063. [Google Scholar] [CrossRef]
  62. Prins, M.J.; Ptasinski, K.J.; Janssen, F.J. Torrefaction of wood: Part 2. Analysis of products. J. Anal. Appl. Pyrolysis 2006, 77, 35–40. [Google Scholar] [CrossRef]
  63. Demirbas, A. Combustion characteristics of different biomass fuels. Prog. Energy Combust. Sci. 2004, 30, 219–230. [Google Scholar] [CrossRef]
  64. Dos Santos, R.C.; Carneiro, A.d.C.O.; Vital, B.R.; Castro, R.V.O.; Vidaurre, G.B.; Trugilho, P.F.; Castro, A.F.N.M. Effect of properties chemical and siringil/guaiacil relation wood clones of eucalyptus in the production of charcoal. Ciênc. Florest. 2016, 26, 657–669. [Google Scholar]
  65. Sturion, J.; Pereira, J.; Chemin, M. Wood quality of Eucalyptus viminalis for energy purposes as a function of spacing and felling age. Bol. Pesqui. Florest. 1988, 16, 55–59. [Google Scholar]
  66. Resquin, F.; Navarro-Cerrillo, R.M.; Carrasco-Letelier, L.; Casnati, C.R. Influence of contrasting stocking densities on the dynamics of above-ground biomass and wood density of Eucalyptus benthamii, Eucalyptus dunnii, and Eucalyptus grandis for bioenergy in Uruguay. For. Ecol. Manag. 2019, 438, 63–74. [Google Scholar] [CrossRef]
  67. Kalderis, D.; Kotti, M.S.; Mendez, A.; Gascó, G. Characterization of hydrochars produced by hydrothermal carbonization of rice husk. Solid Earth 2014, 5, 477–483. [Google Scholar] [CrossRef] [Green Version]
  68. Demirbaş, A. Fuel Characteristics of Olive Husk and Walnut, Hazelnut, Sunflower, and Almond Shells. Energy Sources 2002, 24, 215–221. [Google Scholar] [CrossRef]
  69. Torreiro, Y.; Pérez, L.; Piñeiro, G.; Pedras, F.; Rodríguez-Abalde, A. The Role of Energy Valuation of Agroforestry Biomass on the Circular Economy. Energies 2020, 13, 2516. [Google Scholar] [CrossRef]
  70. José, M.J.S.; Alvarez, S.; Garcia, I.; Peñas, F.J. A novel conical combustor for thermal exploitation of vineyard pruning wastes. Fuel 2013, 110, 178–184. [Google Scholar] [CrossRef]
  71. Nunes, L.J.; Loureiro, L.M.E.F.; Sá, L.C.R.; Silva, H.F. Thermochemical Conversion of Olive Oil Industry Waste: Circular Economy through Energy Recovery. Recycling 2020, 5, 12. [Google Scholar] [CrossRef]
  72. Nunes, L.J.; Godina, R.; Matias, J.C.; Catalão, J. Evaluation of the utilization of woodchips as fuel for industrial boilers. J. Clean. Prod. 2019, 223, 270–277. [Google Scholar] [CrossRef]
  73. Hasan, M.F.; Shamsuddin, H.S.; Heiree, M.H.M.; Hanaffi, M.F.M.A.; Rahman, M.R.A.; Rahman, M.; Latiff, Z.; Harif, M.H.M. Torrefaction of densified mesocarp fibre and palm kernel shell. Renew. Energy 2018, 122, 419–428. [Google Scholar] [CrossRef]
  74. Lau, H.S.; Ng, H.K.; Gan, S.; A Jourabchi, S. Torrefaction of oil palm fronds for co-firing in coal power plants. Energy Procedia 2018, 144, 75–81. [Google Scholar] [CrossRef]
  75. Conag, A.T.; Villahermosa, J.E.R.; Cabatingan, L.K.; Go, A.W. Energy densification of sugarcane bagasse through torrefaction under minimized oxidative atmosphere. J. Environ. Chem. Eng. 2017, 5, 5411–5419. [Google Scholar] [CrossRef]
  76. Chen, D.; Chen, F.; Cen, K.; Cao, X.; Zhang, J.; Zhou, J. Upgrading rice husk via oxidative torrefaction: Characterization of solid, liquid, gaseous products and a comparison with non-oxidative torrefaction. Fuel 2020, 275, 117936. [Google Scholar] [CrossRef]
  77. Chiou, B.-S.; Cao, T.; Valenzuela-Medina, D.; Bilbao-Sainz, C.; Avena-Bustillos, R.J.; Milczarek, R.; Du, W.-X.; Glenn, G.M.; Orts, W.J. Torrefaction kinetics of almond and walnut shells. J. Therm. Anal. Calorim. 2017, 131, 3065–3075. [Google Scholar] [CrossRef]
  78. Margaritis, N.; Grammelis, P.; Karampinis, E.; Kanaveli, I.-P. Impact of Torrefaction on Vine Pruning’s Fuel Characteristics. J. Energy Eng. 2020, 146, 04020006. [Google Scholar] [CrossRef]
  79. Volpe, R.; Messineo, A.; Millan, M.; Volpe, M.; Kandiyoti, R. Assessment of olive wastes as energy source: Pyrolysis, torrefaction and the key role of H loss in thermal breakdown. Energy 2015, 82, 119–127. [Google Scholar] [CrossRef]
  80. Phanphanich, M.; Mani, S. Impact of torrefaction on the grindability and fuel characteristics of forest biomass. Bioresour. Technol. 2011, 102, 1246–1253. [Google Scholar] [CrossRef]
  81. Pode, R. Potential applications of rice husk ash waste from rice husk biomass power plant. Renew. Sustain. Energy Rev. 2016, 53, 1468–1485. [Google Scholar] [CrossRef]
  82. Prasara-A, J.; Gheewala, S.H. Sustainable utilization of rice husk ash from power plants: A review. J. Clean. Prod. 2017, 167, 1020–1028. [Google Scholar] [CrossRef]
  83. Raisi, E.M.; Amiri, J.-V.; Davoodi, M.R. Mechanical performance of self-compacting concrete incorporating rice husk ash. Constr. Build. Mater. 2018, 177, 148–157. [Google Scholar] [CrossRef]
  84. Da Silva, L.R.; de Carvalho Gama, K.N.; Salles, P.V.; Braga, F.C.S. Concrete with rice husk ash and construction and demolition wastes. Res. Soc. Dev. 2019, 8, 2684861. [Google Scholar]
  85. Dyjakon, A.; Garcia, D. Implementing Agricultural Pruning to Energy in Europe: Technical, Economic and Implementation Potentials. Energies 2019, 12, 1513. [Google Scholar] [CrossRef] [Green Version]
  86. Boumanchar, I.; Charafeddine, K.; Chhiti, Y.; Alaoui, F.E.M.; Sahibed-Dine, A.; Bentiss, F.; Jama, C.; Bensitel, M. Biomass higher heating value prediction from ultimate analysis using multiple regression and genetic programming. Biomass Convers. Biorefining 2019, 9, 499–509. [Google Scholar] [CrossRef]
  87. Ren, C.; Guo, D.; Liu, X.; Li, R.; Zhang, Z. Performance of the emerging biochar on the stabilization of potentially toxic metals in smelter- and mining-contaminated soils. Environ. Sci. Pollut. Res. 2020, 1–11. [Google Scholar] [CrossRef]
  88. Verslype, N.I.; de Souza Caldas, R.M.; Machado, J.; Martins, F.M.G.; Fernandez, H.M.; Rodrigues, J.I. Sustainable agriculture in temporary and permanent crops in Portugal| Agricultura sustentável em culturas temporárias e permanentes em Portugal. Rev. Geama 2016, 2, 211–219. [Google Scholar]
  89. Pagels, J.; Strand, M.; Rissler, J.; Szpila, A.; Gudmundsson, A.; Bohgard, M.; Lillieblad, L.; Sanati, M.; Swietlicki, E. Characteristics of aerosol particles formed during grate combustion of moist forest residue. J. Aerosol Sci. 2003, 34, 1043–1059. [Google Scholar] [CrossRef] [Green Version]
  90. Zhao, P.; Shen, Y.; Ge, S.; Chen, Z.; Yoshikawa, K. Clean solid biofuel production from high moisture content waste biomass employing hydrothermal treatment. Appl. Energy 2014, 131, 345–367. [Google Scholar] [CrossRef] [Green Version]
  91. Demirbaş, A. Combustion Systems for Biomass Fuel. Energy Sources Part A Recovery Util. Environ. Eff. 2007, 29, 303–312. [Google Scholar] [CrossRef]
  92. Deboni, T.L.; Simioni, F.J.; Brand, M.A.; Lopes, G.P. Evolution of the quality of forest biomass for energy generation in a cogeneration plant. Renew. Energy 2019, 135, 1291–1302. [Google Scholar] [CrossRef]
  93. Samadi, S.H.; Ghobadian, B.; Nosrati, M. Prediction and estimation of biomass energy from agricultural residues using air gasification technology in Iran. Renew. Energy 2020, 149, 1077–1091. [Google Scholar] [CrossRef]
  94. Williams, A.; Jones, J.; Ma, L.; Pourkashanian, M. Pollutants from the combustion of solid biomass fuels. Prog. Energy Combust. Sci. 2012, 38, 113–137. [Google Scholar] [CrossRef]
  95. Shen, G.; Xue, M.; Wei, S.; Chen, Y.; Zhao, Q.; Li, B.; Wu, H.; Tao, S. Influence of fuel moisture, charge size, feeding rate and air ventilation conditions on the emissions of PM, OC, EC, parent PAHs, and their derivatives from residential wood combustion. J. Environ. Sci. 2013, 25, 1808–1816. [Google Scholar] [CrossRef]
  96. Tavares, S.R.D.L.; Dos Santos, T.E. Uso De Diferentes Fontes De Biomassa Vegetal Para A Producão De Biocombustíveis Sólidos. HOLOS 2013, 5, 19–27. [Google Scholar] [CrossRef]
  97. Strzalka, R.; Erhart, T.G.; Eicker, U. Analysis and optimization of a cogeneration system based on biomass combustion. Appl. Therm. Eng. 2013, 50, 1418–1426. [Google Scholar] [CrossRef]
  98. Bazargan, A.; Gebreegziabher, T.; Hui, C.-W.; McKay, G. The effect of alkali treatment on rice husk moisture content and drying kinetics. Biomass Bioenergy 2014, 70, 468–475. [Google Scholar] [CrossRef]
  99. Manzone, M.; Gioelli, F.; Balsari, P. Effects of Different Storage Techniques on Round-Baled Orchard-Pruning Residues. Energies 2019, 12, 1044. [Google Scholar] [CrossRef] [Green Version]
  100. Kalembkiewicz, J.; Chmielarz, U. Ashes from co-combustion of coal and biomass: New industrial wastes. Resour. Conserv. Recycl. 2012, 69, 109–121. [Google Scholar] [CrossRef]
  101. Nunes, L.J.; Matias, J.C.O.; Catalão, J. Biomass combustion systems: A review on the physical and chemical properties of the ashes. Renew. Sustain. Energy Rev. 2016, 53, 235–242. [Google Scholar] [CrossRef]
  102. Van Der Drift, A.; Van Doorn, J.; Vermeulen, J. Ten residual biomass fuels for circulating fluidized-bed gasification. Biomass Bioenergy 2001, 20, 45–56. [Google Scholar] [CrossRef]
  103. Haykiri-Acma, H.; Yaman, S.; Küçükbayrak, S. Gasification of biomass chars in steam–nitrogen mixture. Energy Convers. Manag. 2006, 47, 1004–1013. [Google Scholar] [CrossRef]
  104. Rajamma, R.; Ball, R.J.; Tarelho, L.; Allen, G.C.; Labrincha, J.A.; Ferreira, V. Characterisation and use of biomass fly ash in cement-based materials. J. Hazard. Mater. 2009, 172, 1049–1060. [Google Scholar] [CrossRef] [PubMed]
  105. Thy, P.; Jenkins, B.; Grundvig, S.; Shiraki, R.; Lesher, C. High temperature elemental losses and mineralogical changes in common biomass ashes. Fuel 2006, 85, 783–795. [Google Scholar] [CrossRef]
  106. Deng, L.; Zhang, T.; Che, D. Effect of water washing on fuel properties, pyrolysis and combustion characteristics, and ash fusibility of biomass. Fuel Process. Technol. 2013, 106, 712–720. [Google Scholar] [CrossRef]
  107. Turn, S.Q.; Kinoshita, C.M.; Ishimura, D.M. Removal of inorganic constituents of biomass feedstocks by mechanical dewatering and leaching. Biomass Bioenergy 1997, 12, 241–252. [Google Scholar] [CrossRef]
  108. Gao, X.; Wu, H. Combustion of Volatiles Producedin Situfrom the Fast Pyrolysis of Woody Biomass: Direct Evidence on Its Substantial Contribution to Submicrometer Particle (PM1) Emission. Energy Fuels 2011, 25, 4172–4181. [Google Scholar] [CrossRef]
  109. Sun, X.; Shan, R.; Li, X.; Pan, J.; Liu, X.; Deng, R.; Song, J. Characterization of 60 types of Chinese biomass waste and resultant biochars in terms of their candidacy for soil application. GCB Bioenergy 2017, 9, 1423–1435. [Google Scholar] [CrossRef]
  110. Singh, K.; Zondlo, J. Characterization of Fuel Properties for Coal and Torrefied Biomass Mixtures. J. Energy Inst. 2017, 90, 505–512. [Google Scholar] [CrossRef]
  111. Picchio, R.; Spina, R.; Sirna, A.; Monaco, A.L.; Civitarese, V.; Del Giudice, A.; Suardi, A.; Pari, L. Characterization of Woodchips for Energy from Forestry and Agroforestry Production. Energies 2012, 5, 3803–3816. [Google Scholar] [CrossRef]
  112. Rosales, E.; Ferreira, L.; Sanromán, M.; Tavares, M.; Pazos, M. Enhanced selective metal adsorption on optimised agroforestry waste mixtures. Bioresour. Technol. 2015, 182, 41–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Enes, T.; Aranha, J.T.M.; Fonseca, T.F.; Lopes, D.; Alves, A.; Lousada, J.L. Thermal Properties of Residual Agroforestry Biomass of Northern Portugal. Energies 2019, 12, 1418. [Google Scholar] [CrossRef] [Green Version]
  114. Enes, T.; Aranha, J.T.M.; Fonseca, T.F.; Matos, C.; Barros, A.I.; Lousada, J.L. Residual Agroforestry Biomass—Thermochemical Properties. Forests 2019, 10, 1072. [Google Scholar] [CrossRef] [Green Version]
  115. Nyakuma, B.B.; Oladokun, O.; Ivase, T.J.; Moveh, S.; Otitolaiye, V.O. Energy Potential of Malaysian Forestry Waste Residues. Pet. Coal 2020, 62, 238–243. [Google Scholar]
  116. Garcia, D.P.; Caraschi, J.C.; Ventorim, G.; Vieira, F.H.A. Trends and Challenges of Brazilian Pellets Industry Originated from Agroforestry. Cerne 2016, 22, 233–240. [Google Scholar] [CrossRef] [Green Version]
  117. Batidzirai, B.; Mignot, A.; Schakel, W.; Junginger, M.; Faaij, A. Biomass torrefaction technology: Techno-economic status and future prospects. Energy 2013, 62, 196–214. [Google Scholar] [CrossRef]
  118. Nanou, P.; Carbo, M.C.; Kiel, J.H. Detailed mapping of the mass and energy balance of a continuous biomass torrefaction plant. Biomass Bioenergy 2016, 89, 67–77. [Google Scholar] [CrossRef] [Green Version]
  119. Viana, H.; Rodrigues, A.; Godina, R.; Matias, J.C.O.; Nunes, L.J. Evaluation of the Physical, Chemical and Thermal Properties of Portuguese Maritime Pine Biomass. Sustainability 2018, 10, 2877. [Google Scholar] [CrossRef] [Green Version]
  120. Loureiro, L.M.; Nunes, L.J.; Rodrigues, A.M. Woody biomass torrefaction: Fundamentals and potential for Portugal. Silva Lusit. 2017, 25, 35–63. [Google Scholar]
  121. Nunes, L.J. A Case Study about Biomass Torrefaction on an Industrial Scale: Solutions to Problems Related to Self-Heating, Difficulties in Pelletizing, and Excessive Wear of Production Equipment. Appl. Sci. 2020, 10, 2546. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Obtained materials: (a) rice husk; (b) almond shells; (c) kiwi pruning; (d) vine pruning; (e) olive oil pomace; and (f) pine woodchips.
Figure 1. Obtained materials: (a) rice husk; (b) almond shells; (c) kiwi pruning; (d) vine pruning; (e) olive oil pomace; and (f) pine woodchips.
Cleantechnol 02 00023 g001
Table 1. Correspondence of the four programmable levels with the different torrefaction phases, depending on the temperature and residence time.
Table 1. Correspondence of the four programmable levels with the different torrefaction phases, depending on the temperature and residence time.
LevelTorrefaction PhaseTemperature (°C)Residence Time (min)
1Heating/DryingTroom (18 °C) to T1Rt1
2Drying/TorrefactionT1 to T2Rt2
3TorrefactionT2Rt3
4CoolingT2 to 50 °CRt4
Table 2. Torrefaction tests.
Table 2. Torrefaction tests.
Temperature (°C)Residence Time (min)
18–180 30
180–300 60
300 90
300–50The time needed to safely collect the material
Table 3. Ultimate analysis for the different products.
Table 3. Ultimate analysis for the different products.
ElementsRice HuskAlmond ShellsKiwi PruningVine PruningOlive PomacePine Woodchips
C (%)Dried 48.6058.4049.4048.3056.2050.20
300 °C52.6079.9075.5073.4081.7080.10
H (%)Dried 4.425.245.695.716.835.86
300 °C2.913.773.623.803.663.91
N (%)Dried 0.5740.2420.5590.7481.2100.100
300 °C0.8330.4701.6001.3201.5600.227
O (%)Dried 46.4136.1244.3545.2435.7643.84
300 °C43.6615.8619.2821.4813.0815.76
Table 4. Thermogravimetric composition for the different products.
Table 4. Thermogravimetric composition for the different products.
PropertiesRice HuskAlmond ShellsKiwi PruningVine PruningOlive PomacePine Woodchips
Fixed Carbon (%)Dried 17.5520.7319.5420.0418.8417.82
300 °C45.1073.6368.5763.5773.7872.98
Moisture (%)Dried 11.558.6410.8710.843.512.61
300 °C4.371.262.982.682.261.87
Ashes (%)Dried 15.861.601.322.891.310.27
300 °C33.194.075.037.984.560.94
Volatiles (%)Dried 66.5977.6779.1477.0679.8581.91
300 °C21.7122.3026.4028.4621.6626.08
Table 5. High heating values calculated for the different torrefaction temperatures.
Table 5. High heating values calculated for the different torrefaction temperatures.
PropertiesRice HuskAlmond ShellsKiwi PruningVine PruningOlive PomacePine Woodchips
HHV (MJ/kg)Dried 17.0322.7919.3318.8423.9219.89
300 °C16.5630.6028.5027.6931.3630.92
LHV (MJ/kg)Dried15.2820.7117.0716.5821.2217.57
300 °C15.4129.1127.0726.1929.9129.37
Mass loss (%)56.9068.0066.0063.5071.1069.50

Share and Cite

MDPI and ACS Style

Nunes, L.J.R.; Loureiro, L.M.E.F.; Sá, L.C.R.; Silva, H.F.C. Waste Recovery through Thermochemical Conversion Technologies: A Case Study with Several Portuguese Agroforestry By-Products. Clean Technol. 2020, 2, 377-391. https://doi.org/10.3390/cleantechnol2030023

AMA Style

Nunes LJR, Loureiro LMEF, Sá LCR, Silva HFC. Waste Recovery through Thermochemical Conversion Technologies: A Case Study with Several Portuguese Agroforestry By-Products. Clean Technologies. 2020; 2(3):377-391. https://doi.org/10.3390/cleantechnol2030023

Chicago/Turabian Style

Nunes, Leonel J. R., Liliana M. E. F. Loureiro, Letícia C. R. Sá, and Hugo F. C. Silva. 2020. "Waste Recovery through Thermochemical Conversion Technologies: A Case Study with Several Portuguese Agroforestry By-Products" Clean Technologies 2, no. 3: 377-391. https://doi.org/10.3390/cleantechnol2030023

Article Metrics

Back to TopTop