Next Article in Journal
A Multi-Objective Optimization Approach for Corrective Switching of Transmission Systems in Emergency Scenarios
Previous Article in Journal
Enhanced Production of Bioethanol by Fermentation of Autohydrolyzed and C4mimOAc-Treated Sugarcane Bagasse Employing Various Yeast Strains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physical and Compression Properties of Pellets Manufactured with the Biomass of Five Woody Tropical Species of Costa Rica Torrefied at Different Temperatures and Times

by
Johanna Gaitán-Alvarez
1,
Roger Moya
1,*,
Allen Puente-Urbina
2 and
Ana Rodriguez-Zuñiga
1
1
Instituto Tecnológico de Costa Rica, Escuela de Ingeniería Forestal, Apartado 159-7050, Cartago, Costa Rica
2
Instituto Tecnológico de Costa Rica, Centro de Investigación y de Servicios Químicos y Microbiológicos (CEQIATEC), Escuela de Química, Apartado 159-7050, Cartago, Costa Rica
*
Author to whom correspondence should be addressed.
Energies 2017, 10(8), 1205; https://doi.org/10.3390/en10081205
Submission received: 29 June 2017 / Revised: 5 August 2017 / Accepted: 8 August 2017 / Published: 14 August 2017
(This article belongs to the Section L: Energy Sources)

Abstract

:
The purpose of the biomass torrefaction and pelletizing processes is to increase its energy properties, be environmentally friendly, decrease shipping costs, and make handling easier. The objective of the present work is to evaluate the density, internal density variation by X-ray densitometry, moisture content, water absorption, and compression force of torrefied biomass pellets of five wood species (Cupressus lusitanica, Dipterix panamensis, Gmelina arborea, Tectona grandis, and Vochysia ferruginea) under three torrefaction temperature conditions (light, middle, and severe) and three torrefaction times (8, 10, and 12 min). The results showed that the bulk density of the pellets was 0.90–1.30 g/cm3. The density variation of the pellets was higher with torrefaction at 250 °C. The moisture content decreased with increasing torrefaction temperature from 3% to 1%. Water absorption was lower in the pellets torrefied at 250 °C. The compression force was lower in the pellets torrefied at 250 °C with approximate loads of 700 N. Based on the above results, it was concluded that pellets made with biomass torrefied at 200 °C have better energy properties and evaluated properties. According to these results, pellets fabricated with the torrefied biomass of tropical species can be used in stove, gas, and hydrogen production because the pellet presents adequate characteristics.

1. Introduction

Massive utilization of energy in recent years has become a worldwide phenomenon, with fossil fuel consumption being the main source of energy [1]. However, this is a costly source of energy, and the environmental issues the process carries with it cannot be ignored [2]. As a result, research on new forms of energy has been conducted that promotes the development of new sustainable technologies using renewable raw material [3]. Biomass can be considered one of those new and renewable products [4].
Lignocellulosic biomass is one of the renewable sources with a greater capacity for energy production; it can be obtained from residuals from various agricultural and forest industries [5,6]. However, biomass presents a series of disadvantages that make it difficult to use as a source of energy: its moisture content, hygroscopic nature, low energetic density, and storage difficulties, among others [7,8]. For this reason, various processes have been implemented to improve the natural characteristics of the biomass [8].
Torrefaction is one of the treatments for improving biomass conditions [9], which enhances its physical characteristics, such as moisture absorption, and increases its calorific potential [10]. The process consists of the application of low temperature, ranging from 200 to 300 °C in an oxygen-free environment, to prevent biomass material from burning due to oxygen flammability [11]. Although each type of biomass has its own torrefaction temperature ranges, some authors have determined three torrefaction temperature categories for species from temperate climates mainly: light torrefaction, with temperatures ranging from 200 to 235 °C; middle torrefaction, with temperatures ranging from 235 to 275 °C; and severe torrefaction, with temperatures ranging from 275 to 300 °C [12]. The torrefaction process parameters are similar to those used in the roasting of coffee beans. The products of the torrefaction process are solid, liquid, and non-condensable gas. The main product of the torrefaction process is the solid, which is called torrefied biomass [10]. Proskurina et al. [9] indicate that: (i) torrefied biomass presents lower ash content and is more environmentally friendly than fossil fuel from the perspective of net greenhouse gas emissions; (ii) these characteristics make torrefied biomass promising not only for final applications but also from a logistical point of view; and (iii) the high-energy density of torrefied biomass decreases shipping costs and makes handling easier.
Once torrefied, the biomass can be densified by manufacturing pellets or briquettes [13]. To produce pellets, the biomass is pressed mechanically to compress the wood’s cell structure and make it denser [14,15,16]. Thus, the energy density increases considerably, the moisture content decreases, and transport and storage costs are reduced [17].
Despite all of the positive characteristics of densification (pelletization), biomass in the form of pellets shows certain unfavourable conditions, including low mechanical and compaction resistance, which leads to problems related to handling and durability [18].
Both techniques, torrefaction and pelletization, have been combined for several years to improve biomass processing and performance [14]. It has been demonstrated that pelletized torrefied biomass is less prone to moisture absorption [14]. However, pellets manufactured with torrefied biomass are more difficult to compress, and therefore the energy consumption needed in the process is higher [19]. In addition, the densities and mechanical strengths are lower compared to pellets manufactured with untorrefied biomass [14]. Various techniques have been used to evaluate the quality of pellets that provide information as to their internal structure. Among those techniques are X-rays [20] and compression force [21,22].
On the other hand, woody biomass presents suitable conditions as source of energy [23], while torrefaction can increase its energy efficiency [24]. However, not enough information is available yet on the behaviour of tropical woods—either from plantations or from natural forests—regarding torrefaction [25,26].
Specifically in Costa Rica, the utilization of biomass as a source of energy started several years ago [27], including pellet utilization [28]. Among the major species used for biomass are Cupressus lusitanica, Dipteryx panamensis, Gmelina arborea, Tectona grandis, and Vochysia ferruginea [26]. These tree species are produced in timber forest plantations, and their wastes are used for pellet production [22,25]. Nonetheless, studies on bioenergy production showing the energy potential of these species or the likelihood of increasing that potential are scarce [25,26].
Given the above conditions, the present study aims to evaluate pellet production from the torrefied biomass of five potential woody biomass species (C. lusitanica, D. panamensis, G. arborea, T. grandis, and V. ferruginea) under three temperature conditions (light, middle, and severe) and torrefaction times. The evaluation takes into account density, internal variation of the density by X-ray densitometry, humidity after pelletizing, water absorption, and compression force. This study will increase the biomass energy efficiency of five tropical species through torrefaction, and so this process can be used to increase the economic value of the plantations of these species. Additionally, the best conditions of temperature and time of torrefaction will be established for a torrefaction process with the highest energy efficiency.

2. Results

2.1. Pellet Appearance with Torrefied Biomass

Biomass torrefaction at different temperatures is shown clearly in the pellet colour (Figure 1). Lightly torrefied pellets (200 °C during 8 min (200-8), 200 °C during 10 min (200-10), and 200 °C during 12 min (200-12)) show a natural colour almost identical to that of nontorrefied biomass (Figure 1a). The pellet colour at the middle torrefaction temperature (225 °C during 8 min (225-8), 225 °C during 10 min (225-10), and 200 °C during 12 min (225-12)) turns darker than lightly torrefied biomass, taking a greyish shade (Figure 1b). Finally, pellets under the severe torrefaction temperature (250 °C during 8 min (250-8), 250 °C during 10 min (250-10), and 250 °C during 12 min (250-12)) present a charcoal-like black colour (Figure 1c).

2.2. Pellet Density

The pellet density, Dpellet, of the five species torrefied varied between 0.90 g/cm3 and 1.30 g/cm3 (Figure 2a–e). Although the pellets manufactured from C. lusitanica biomass showed differences in the values of Dpellet, this parameter is statistically different only in biomass torrefied at 200-12 (Figure 2a). D. panamensis pellets showed no statistical differences in the Dpellet of biomass torrefied at the various torrefaction temperatures and times (Figure 2b). In G. arborea pellets, the Dpellet increases when manufactured with biomass torrefied at 250-8 and 250-10, showing statistical differences with pellets torrefied under other conditions (Figure 2c). T. grandis biomass pellets presented a significant increment in the Dpellet when manufactured with biomass torrefied at 250-12 (Figure 2d). The pellets manufactured with V. ferruginea biomass under torrefaction conditions of 225-10 and 250-8 presented a significant diminution in the Dpellet, while pellets manufactured with biomass torrefied at 250-8 and 250-12 showed a significant increase in the Dpellet (Figure 2e).
As for the density variation in longitudinal direction measured by coefficient of variation in longitudinal direction (CVlong), pellets manufactured with C. lusitanica biomass showed greater statistical CVlong with biomass torrefied at 225-12 and 250-12 (Table 1). The remaining torrefaction conditions were statistically similar. The pellets made with biomass from D. panamensis and T. grandis showed no statistical differences in CVlong between the various biomass torrefaction conditions (Table 1). On the other hand, the CVlong of the pellets made with torrefied biomass from G. arborea at 225-8 and 250-12 presented statistical differences greater than with the rest of the conditions. As for the remaining torrefaction conditions, the CVlong showed no statistical differences between the various types of torrefied biomass (Table 1). Regarding biomass from V. ferruginea, the pellets manufactured with biomass torrefied at 225 °C and the three times tested showed the highest statistically significant CVlong, followed by pellets with biomass torrefied at 200 °C and 250 °C and 8 and 10 min. The statistically lower CVlong was observed in pellets from biomass at 250-12 (Table 1).
The density variation profiles (Figure 3) show the behaviour of CVlong. Three variation patterns were observed in each species: (1) a first pattern showing a low variation of the density, normally corresponding to pellets manufactured with lightly torrefied biomass at 200 °C and the three times (Figure 4a–e); (2) a second pattern showing mean density variation, generally corresponding to pellets manufactured at 225 °C (Figure 4a–e); and (3) the last pattern characterized by a high variation of the pellet density. In this last case, irregularities were observed all along the pellet (areas of high and low density very close to each other). This pattern of variation was observed in pellets manufactured with biomass torrefied at 250 °C (Figure 3a–e).

2.3. Moisture Content and Moisture Absorption

The pellet moisture content (MCpellet), of biomass pellets from C. lusitanica torrefied at 200-8, 200-10, and 250-8 were statistically different from the rest of the pellets torrefied under the other conditions (Table 1). As for D. panamensis, the MCpellet of pellets manufactured with biomass torrefied at 200-8 and 200-10 presented significant differences, followed by pellets at 200-12, which also were statistically different (Table 1). The highest MCpellet in G. arborea was observed in pellets manufactured with biomass torrefied at 200 °C and the three times, and 225-10, with the latter being statistically different from those manufactured with biomass at 225-8, 225-12, 250-10, and 250-12. On the other hand, the statistically lower MCpellet was observed in pellets manufactured with biomass torrefied at 250-8 (Table 1). As for T. grandis, the pellets manufactured with biomass torrefied at 200-8 and 200-10 presented similar statistical behaviour, followed by another group of pellets from biomass torrefied at 200-12, 225 °C for the three times, and 250-10. Meanwhile, the pellets from biomass at 250-8 and 250-12 presented an MCpellet statistically lower (Table 1). As for pellets manufactured with biomass from V. ferruginea, those torrefied at 200 °C and the three times, 250-10, and 250-12 showed statistically similar behaviour. The other conditions showed statistically similar behaviour, although differing from the previous (Table 1).
The values of the parameter percentage of water absorption, Wabs, of pellets with different types of torrefied biomass were in a range of 6% and 13% (Table 1). Three groups may be made according to the previous value: a first group with statistically higher values of Wabs, consisting of pellets manufactured with biomass torrefied at 200-8, 200-10, 225-10, 225-12, and 250-8; a second group with pellets from biomass torrefied at 200-12 and 225-8, and a third group including biomass torrefied at 250-10 and 250-12 (Table 1). As for D. panamensis concerning Wabs, the pellets manufactured with biomass torrefied at 225-8 were statistically different. Likewise, the pellets with biomass torrefied at 250-12 showed the lowest Wabs (Table 1). As regards to G. arborea pellets, those manufactured under the condition of 225-8 showed the highest Wabs being statistically different, followed by pellets manufactured with biomass under 250-10 and 250-12, which also showed statistical differences, with the lowest statistical values of Wabs (Table 1). T. grandis pellets from biomass at 200-10 and 250-10 were statistically different, showing low values of Wabs; however, the statistically lower value of Wabs was observed with biomass torrefied under the condition 250-12 (Table 1). Meanwhile, V. ferruginea pellets manufactured with biomass torrefied at 225-8 presented statistically higher values of Wabs; the pellets with biomass torrefied at 250-8 presented statistically lower values of Wabs, and the remaining conditions do not present statistical differences (Table 1).

2.4. Compression Force

The compression force, Cforce, of pellets from torrefied biomass showed irregular behaviour in all species. C. lusitanica biomass pellets under conditions of 200-8, 250-8, and 250-10 showed the highest Cforce, while pellets under conditions of 225 °C with all times and 250-12 showed statistically lower values of Cforce (Table 1, Figure 5a). As for D. panamensis torrefied biomass, pellets under 225 °C for the three times presented the lowest Cforce and were statistically different from the conditions at 200 and 250 °C. This situation was also observed in G. arborea, where pellets torrefied at 200 and 250 °C for the three times presented differences, while pellets manufactured with biomass torrefied at 225 °C for the three times showed statistically lower values of Cforce (Table 1). With T. grandis, the pellets presenting greater Cforce were torrefied at 225-8 and those at 250 °C for the three torrefaction times, were statistically different with respect to pellets that supported lower Cforce (Table 1). In the case of V. ferruginea, the pellets under conditions of 200-10, 200-12, and 250-8 presented the highest and statistically different values of Cforce, while the pellets with the statistically lowest values of Cforce were torrefied at 225-8 and 225-12 (Table 1). Cforce variation with the deformation of the pellet during the compression test reflects the previous behaviour of the compression force at the different torrefaction temperatures of the different species. The load-deformation curve of pellets manufactured with biomass torrefied at 250 °C presents low load values and little deformation in all of the species, while pellets made with biomass torrefied at 200 °C and 225 °C, especially under conditions of 250-8, 250-10, and 250-12, present high load value and high deformation (Figure 4a–e).

2.5. Effect of Different Properties of Pellets Made with Torrefied Biomass

Table 2 presents the Pearson correlation matrix of the different variables evaluated in each of the five species studied. The Dpellet parameter of the different torrefied biomasses had a significant negative effect on the CVlong of the five species, with the torrefied biomass of V. ferruginea showing a higher correlation coefficient (Table 2). The Dpellet parameter is positively and significantly correlated in D. panamensis, G. arborea, and T. grandis (Table 2). The relationship between Dpellet and Wabs was positive and significant in G. arborea and negative in V. ferruginea (Table 2). As regards to the relationship between Dpellet and MCpellet, it affects the five species studied, with a positive correlation in V. ferruginea, whereas in the other four species this correlation is negative (Table 2).
As for CVlong, it was negatively correlated to Cforce in pellets manufactured with the torrefied biomass of C. lusitanica and V. ferruginea, with the latter species showing a higher correlation coefficient (Table 2). CVlong was affected positively and significantly in D. panamensis and V. ferruginea, while negatively in T. grandis biomass (Table 2). The correlation between CVlong and MCpellet was significant and negative in biomass from D. panamensis and V. ferruginea, where the first species shows a greater correlation coefficient (Table 2).
The Wabs affected Cforce negatively in four of the five species studied, with G. arborea as the only one unaffected by this parameter. The biomass of D. panamensis presented the highest correlation coefficient between these two variables (Table 2). In the case of MCpellet, Cforce was statistically affected in the biomass of C. lusitanica, G. arborea, and V. ferruginea, showing a positive correlation and a negative correlation in the case of T. grandis biomass (Table 2). Finally, the relationship between Wabs and MCpellet was only statistically significant in the biomass of V. ferruginea (Table 2).

3. Discussion

3.1. Appearance of Pellet with Torrefied Biomass

Torrefaction caused a change in pellet colour in all species and temperature conditions studied. This change occurs because the increase in the torrefaction temperature causes a chemical change in the polymers that compose the biomass, especially the lignin, so that the pellet becomes darker and its surface loses gloss and softness [29]. This change in colour increases with increasing torrefaction temperature [30], tending to a darker colour. Severe (250-8, 250-10, and 250-12) biomass torrefaction temperature, therefore, causes a darker colour in the pellets (Figure 1). This color change with the increasing of temperature agrees with the color change occurred in pellets fabricated with Japanese cedar [31] and branches of birch and spruce [32].

3.2. Pellet Density

The Dpellet in nontorrefied biomass is positively correlated to the density of the wood of each species [21]. Therefore, torrefied biomass of species with low density, such as G. arborea, present low Dpellet. In contrast, the torrefied biomass of high-density species such as D. panamensis show higher Dpellet values (Figure 2).
In relation to pellet manufacturing with torrefied biomass, Dpellet increases with increasing biomass torrefaction temperature [24], due to the fact that during this thermal process and the application of forces to compress the biomass, water and chemical compounds with low melting points are lost, since they act as binding agents when softened at 100 °C [33], and thus produce increased Dpellet [25]. Again, the increasing of bulk density with torrefaction temperature agrees with the results found by Mišljenović et al. [32] for pellets fabricated with biomass torrefied with branches of birch and spruce and pellets fabricated with biomass of Olea uropaea [34].
This behaviour could be observed in the present work in some types of torrefied biomass pellets from wood species such as G. arborea (Figure 2c), T. grandis (Figure 3d), and V. ferruginea (Figure 2e), where Dpellet increased with torrefaction temperature and time, especially in severe torrefaction temperatures (250-8, 250-10, and 250-12). However, this behaviour is not evident in C. lusitanica (Figure 3a) or D. panamensis (Figure 2b), where Dpellet is uniform, even in light and severe torrefaction conditions, for in the five types of biomass there is little change in the Dpellet (Figure 2a–e).
The variability of the Dpellet between the different species and torrefaction conditions can be solved by conditioning the torrefied biomass and the pellet itself to moisture content (MC) between 3.5% and 10.5% [35,36]. The correlation that was found between Dpellet and MCpellet (Table 2) confirms the importance of the MC in the Dpellet (decrease of Dpellet with the increase of MCpellet), so that by achieving uniform MCs, we can decrease the density variability.
Another important aspect in the present work is that Dpellet values are higher than those reported for biomass pellets without torrefaction [21,25]. This increase in Dpellet occurs through the use of compression temperatures between 70 °C and 280 °C, as there is an improvement in the compaction of the material, causing the pellets to have higher densities than pellets made with nontorrefied biomass [14,24,37]. However, increasing the Dpellet, either by temperature or Cforce [14,24], requires higher energy consumption for pellet manufacture [35], so its implementation requires financial assessment [38].
CVlong is related to several factors, including the nature of the biomass fibre, the particle size, the MC of the particles, the temperature of the particles, the compression speed, the size and shape of the pellet, and the temperature of compression [39]. Although these factors were not studied in the present work (Table 2), we can infer some results of CVlong from the nature of the biomass fibre, which is different among the species. Woods that have thicker cell walls produce few internal variations of Dpellet and vice versa [40,41]. Here, we verified that species such as D. panamensis and T. grandis (Table 1) show uniform CVlong, whereas thinner cell wall species, such as C. lusitanica, G. arborea, and V. ferruginea, present higher CVlong (Table 1).
It was also observed that the CVlong of the torrefied biomass pellets of all of the species present greater irregularities in the pellets elaborated with biomass torrefied at 250 °C (Figure 4a–e). Such behaviour is because, given the properties of the roasted material, at high temperatures (middle and severe torrefaction) a severe structural and chemical degradation of the material occurs [12], which is a reason why the material is more structurally unstable and presents greater difficulties in compression [12], which is reflected in the density variations (Figure 4a–e).
The most significant correlation occurred between Dpellet and CVlong in the five species. This is a negative relationship (Table 2), which means that the smaller the Dpellet, the greater the variation of density because compaction was not uniform [12,25]. Also, as indicated, a better compaction was observed in pellets made with torrefied biomass in the light temperature condition, since the lowest CVlong (Figure 1a–e) was observed under this condition.

3.3. Moisture Content and Moisture Absorption

MCpellet and Wabs are parameters of high interest, since pellets with low hygroscopicity and that are highly resistant to water can reduce the costs of handling, transport, and storage, prolonging pellet life [12]. Torrefaction increases the hydrophobicity of the material by removing the hydroxyl groups as the biomass decomposes with temperature [42]. Biomass with low values of MC and Wabs has the potential to make torrefied biomass pellets. In the species studied, MCpellet decreased when the torrefaction temperature increased, especially in torrefaction at 250 °C, which presented lower MCpellet in all species (Table 1); this torrefaction condition, therefore, could have potential in pellet manufacturing. The decreasing of water absortion with torrefaction temperature agrees with the results found by Rudolfsson et al. [43] for pellets fabricated with biomass torrefied with scots pine and pellets fabricated with torrefied sawdust [33].
Dpellet was strongly negatively correlated with the MCpellet (Table 2), especially in the torrefied biomas of C. lusitanica, D. panamensis, G arborea, and T. grandis, which means that as the Dpellet decreases, the MCpellet increases (Table 2). This is so because porosity among the particles is greater in pellets with low density, allowing the ingress of moisture [19].
The values of Wabs in the pellets decrease with increasing torrefaction temperature, especially in biomass torrefied at 250 °C (Table 1). This occurs because during torrefaction, the thermal decomposition of the biomass occurs and large voids are formed in the cell wall that reduce their size at the time of densification, decreasing Wabs through the pores as a result [43]. Other authors explain that higher Wabs occurs because the component of biomass that has a greater capacity for moisture absorption is hemicellulose, while lignin shows low moisture absorption tendency [44]. Thus, at the time of an application of thermal treatment to the material, such as torrefaction, degradation occurs in a high percentage of hemicellulose together with an increase in the proportion of lignin, which is the most difficult component to degrade thermally [12]. Thus, torrefaction enhances the hydrophobicity of the material [44], with low Wabs in pellets after severe torrefaction (Table 1).
The effect of Dpellet on Wabs was irregular in the different species (Table 2), having no effect regarding C. lusitanica, D. panamensis, and T. grandis, while showing effects in G. arborea and V. ferruginea. In the first species it affected positively, and it affected negatively in the second, indicating that Wabs is more related to other parameters of the biomass, such as its chemical modification, which determines the capacity to absorb moisture.

3.4. Compression Force

It has been demonstrated that pellets manufactured with torrefied biomass show lower Cforce [45]. This behaviour was observed in the five types of biomass in the present study (Table 1). This diminution occurs because during torrefaction, as commented on above, the decomposition of hemicellulose, lignin, and extractives such as proteins, starches, and carbon hydrates occurs, which reduces the amount of hydrogen and solid bridges among particles, leaving spaces between the pores [34]. As a result, during the test for compression force, these pores or spaces promote particle movement, with the resulting lower Cforce [46]. The effect of temperature of torrefaction on Cforce disagrees with the results obtained by Carone et al. [33] for pellets fabricated with the biomass of Olea uropaea.
Cforce is also influenced by the MCpellet, since the latter plays an important role at the time of pelletization. Moisture attracts the adjacent particles [46]. Such behaviour was observed in this study, for MCpellet decreases with increasing torrefaction temperatures (Table 1); the decrease of MCpellet leads to a decrease of hydrogen bonds and weakens the pelletization [34], thus reducing the Cforce.
Although the above explanations can be applied to the difference in Cforce, there is a tendency in Cforce to diminish with torrefied biomass, especially under the condition 250-12 (Table 1). However, in some species, such as D. panamensis, G. arborea, and V. ferruginea, middle torrefied biomass produces pellets with middle to low Cforce, which increases again at 250 °C. However, there are manufacturing factors intervening here, such as the relationship between this mechanical property with the Dpellet and the CVlong (Table 2).
The Cforce correlates with Dpellet, Wabs, and MCpellet (Table 2). The Cforce in the torrefied biomass of D. panamenis, G. arborea, and T. grandis showed positive correlation (Table 2). This situation was to be expected [21], where the greater the density, the greater is Cforce, due to the greater compaction [46].
It appears that pellets with Cforce show the highest values of Wabs (Table 2). This is because pellets with high density show greater particle compaction, and therefore less cavities for the ingress of moisture are available, and the pellet’s capacity for water absorption is reduced [46].
Lastly, Cforce and MCpellet were positively correlated in the torrefied biomass of C. lusitanica, G. arborea, and V. ferruginea (Table 2), indicating that pellets with high values of MC show high Cforce. However, this effect is produced by torrefied biomass at low temperatures (225 °C), where pellets with low density and high moisture are produced (Table 1), with a resulting collinearity among the variables produced by the temperature of torrefaction.

3.5. Optimal Conditions for the Biomass Torrefaction of Reforestation Species

The previous analyses indicate that pellets manufactured with biomass torrefied at 250 °C, or severe torrefaction, present undesirable characteristics: low Dpellet (Figure 2a–e), greater pellet CVlong (Figure 3a–e), and lower Cforce (Table 1), which are inadequate for energy production from torrefied biomass. Conversely, pellets manufactured with biomass torrefied at 200 °C showed optimal characteristics and conditions to be used for energy generation, such as adequate density values (Figure 2a–e), slight variations of density in longitudinal direction (Figure 3a–e), and high values of Cforce (Table 1, Figure 4a–e).
The appropriate conditions of pellets with torrefied biomass in the light condition in this study are consistent with previous works that indicate that the optimal torrefaction temperatures for pre-treating the biomass are light or lower than 235 °C [47]. In addition, this study agrees in that it is not recommended to use severe torrefaction as pre-treatment, since strong material degradation is caused [12,48]. Light torrefaction causes a significant degradation of hemicellulose, but its effect on cellulose and lignin is slight, with no severe degradation of the biomass [48].
On the other hand, Chen et al. [48] and Nunes et al. [49] indicate that hardwood species endure high torrefaction temperatures, while softwood species do not. This behaviour could be observed in the species in this work, where D. panamensis and T. grandis, with high density, present better behaviour with torrefaction, while C. lusitanica and G. arborea, low density woods, are more prone to early degradation (Table 1); therefore, lower torrefaction temperatures should be used with both.
Finally, the results found in this study were derived from pilot advice adapted to a universal testing machine (Figure 5). However, pellet properties had been determined for similar advice used by Mani et al. [50], Rhen et al. [51], and Carone et al. [33]; they had demonstrated the reproduction of the conditions of the industrial process of pellet fabrication, and they can be implemented in an industrial application [52,53].

4. Materials and Methods

4.1. Provenance and Characteristics of the Material

Woody waste biomass of C. lusitanica, D. panamensis, G. arborea, T. grandis, and V. ferruginea from fast growth plantations in different sites in Costa Rica was used. The age of the plantations ranged between 8 and 14 years. The details of provenance of the material is available in Moya et al. [26]. Sawdust from all of the species was collected directly from the sawing process, conditioned to 12% moisture content, and then sieved. After sieving, sawdust particles were used in a proportion of 70% of 450–1000 μm and 30% of 200–425 μm.

4.2. Torrefaction Process

Approximately 500 grams sawdust was taken from each species, and then the material was divided to apply three different times of exposure to torrefaction (8, 10, and 12 min), and three different temperatures (200, 225, and 250 °C), resulting in nine treatments per species. Thus, the torrefaction conditions per species were: 200 °C during 8 min (200-8), 200 °C during 10 min (200-10), 200 °C during 12 min (200-12), 225 °C during 8 min (225-8), 225 °C during 10 min (200-10), 225 °C during 12 min (200-12), 250 °C during 8 min (250-8), 250 °C during 10 min (250-10), and 250 °C during 12 min (250-12). The Thermolyne Furnace 48,000 was used for torrefaction, and was sealed to control the airflow. Every 4–5 min, the air was freed using a stopcock to stop the oxygen from coming in and thus to prevent material ignition at the time the torrefaction process was being conducted [25].

4.3. Pelletizing Process

The torrefied material was pelletized using a pellet device utilized in a universal testing machine of 1TON brand JBA, model 855. The attachment consists of a 6 mm diameter stem and a plunger that compresses the biomass; the pellets’ temperature and pressure were controlled during manufacturing (Figure 5). Pellets were manufactured using 1.09 g of the torrefied material that was placed in the stem of the attachment, and then compressed at a speed of 60 m/min. The biomass was compressed to obtain pellets approximately 2.5 cm long and then left for a period of 2 min at a temperature of 180 °C. Finally, the pellets were extracted from the machine. Fifty-six pellets (6 pellets × 3 temperatures × 3 times) were prepared for each of the species.

4.4. Pellet Density and Variation by X-ray Densitometry

Four hours after manufacturing, the pellets’ length, diameter, and weight were measured to calculate the pellet density (Dpellet) using the mass-volume ratio of the pellet. Following this, the internal variation of Dpellet was calculated by means of X-ray densitometry in the longitudinal direction. For this, an X-ray Quintek Messurement Systems Inc (Knoxville, TN, USA) scanner model QTRS-01X was used. The exposure conditions used were 7 KV tension in the tube and density readings during 1 s every 40 μm. The procedure to determine the density variation in longitudinal direction (CVlong) is presented in detail in Tenorio et al. [21], and consists of finding the coefficient of variation of all the density readings realized every 40 μm.

4.5. Water Absorption Percentage, Moisture Content, and Compression Force of the Pellets

The 25 mm long pellets were cut in half to extract two samples each 12.5 mm long. One sample was used to measure the percentage of water absorption (Wabs) and the pellet moisture content (MCpellet). The second sample was used to measure the compression force (Cforce). For Wabs determination, half of the pellet was placed in a desiccator with a saturated solution of potassium nitrate at 22 °C (21% in equilibrium moisture content) for 8 days to obtain constant weight. The details of this procedure are available in Tenorio et al. [22]. The samples were weighed before and after being placed in the desiccator. Equation (1) was used to determine Wabs. For MCpellet determination, after absorption the pellet was placed into an oven at 103 °C for 24 h and was then weighed. The MCpellet was determined by means of Equation (2).
M o i s t u r e   a b s o r p t i o n   ( % ) = w e i g h t   a t   21 % ( g ) i n i t i a l   w e i g h t ( g ) i n i t i a l   w e i g h t   ( g ) × 100
M o i s t u r e   c o n t e n t   ( % ) = w e i g h t   a f t e r   p e l l e t i n g   ( g ) w e i g h t   o v e r   d r i e d   ( g ) w i e i g h t   a f t e r   p e l l e t i n g   ( g ) × 100
For Cforce determination, a first adjustment was made for all of the pellet samples to have a standard length of 12.5 mm. This test was performed in the lateral direction of the pellet according to the methodology proposed by Aarseth and Prestlokken [54]. In this test, the compression load was applied at a speed of 0.2 mm/s and the maximum load reached by the pellet previous to failure is reported.

4.6. Statistical Analysis

A descriptive analysis (mean, standard deviation, maximum and minimum values) for the variables Dpellet, CVlong, Wabs, and Cforce was performed. Additionally, compliance of these variables with the assumptions of normality and homoscedasticity of variance was verified. Then, an ANOVA variance analysis was performed to verify the existence of significant differences between means of the variables (p < 0.05). The ANOVA model was established considering two independent non-random sources of variation (temperature and time, both at three levels) and the variables were measured as dependent variables. Lastly, the Tukey test was applied to determine the statistical differences between the means. The X-ray density values provided by the equipment were used to construct the density profiles in a longitudinal direction to establish the patterns of variation of Dpellet.

Acknowledgments

The authors thank the Vicerrectoría de Investigación y Extensión of the Instituto Tecnológico de Costa Rica and CONARE for their financial support.

Author Contributions

Roger Moya contributed with the coordination of the Project, the experimental design of the work, and the analysis of all the data obtained. Johanna Gaitán-Álvarez ran the necessary tests to obtain the data, also elaborated the analysis of the data, and helped with the writing of the paper. Allen Puente-Urbina helped with data analysis and experimental design from a chemical analysis perspective. Ana Rodriguez-Zuñiga collaborated with the testing runs as well as with the preparation of the pellets.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fournel, S.; Palacios, J.H.; Morissette, R.; Villeneuve, J.; Godbout, S.; Heitz, M.; Savoie, P. Influence of biomass properties on technical and environmental performance of a multi-fuel boiler during on-farm combustion of energy crops. Appl. Energy 2015, 141, 247–259. [Google Scholar] [CrossRef]
  2. 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]
  3. Kullander, S. Energy from biomass. Eur. Phys. J. Spec. Top. 2009, 176, 115–125. [Google Scholar] [CrossRef]
  4. Kim, S.; Dale, B.E. All biomass is local: The cost, volume produced, and global warming impact of cellulosic biofuels depend strongly on logistics and local conditions. Biofuels Bioprod. Biorefin. 2015, 9, 422–434. [Google Scholar] [CrossRef]
  5. Bahng, M.; Mukarakate, C.; Robichaud, D.; Nimlos, M. Current technologies for analysis of biomass thermochemical processing. Anal. Chim. Acta 2009, 651, 117–138. [Google Scholar] [CrossRef] [PubMed]
  6. Gokcol, C.; Dursunb, B.; Alboyaci, B.; Sunan, E. Importance of biomass energy as alternative to other sources in Turkey. Energy Policy 2009, 37, 424–431. [Google Scholar] [CrossRef]
  7. Wang, G.; Luo, Y.; Deng, J.; Kuang, J.; Zhang, Y. Pretreatment of biomass by torrefaction. Chin. Sci. Bull. 2011, 56, 1442–1448. [Google Scholar] [CrossRef]
  8. Medic, D.; Darr, M.; Shah, A.; Potter, B.; Zimmerman, J. Effects of torrefaction process parameters on biomass feedstock upgrading. Fuel 2012, 91, 147–154. [Google Scholar] [CrossRef]
  9. 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]
  10. Wu, M.; Schott, D.; Lodewijks, G. Physical properties of solid biomass. Biomass Bioenergy 2011, 35, 2093–2105. [Google Scholar] [CrossRef]
  11. Van der Stelt, M.; Gerhauser, H.; Kiel, J.; Ptasinski, K. Biomass upgrading by torrefaction for the production of biofuels: A review. Biomass Bioenergy 2011, 35, 3748–3762. [Google Scholar] [CrossRef]
  12. 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]
  13. Du, S.W.; Chen, W.H.; Lucas, J.A. Pretreatment of biomass by torrefaction and carbonization for coal blend used in pulverized coal injection. Bioresour. Technol. 2014, 161, 333–339. [Google Scholar] [CrossRef] [PubMed]
  14. Li, H.; Liu, X.; Legros, R.; Bi, X.T.; Jim Lim, C.; Sokhansanj, S. Pelletization of torrefied sawdust and properties of torrefied pellets. Appl. Energy 2012, 93, 680–685. [Google Scholar] [CrossRef]
  15. Stelte, W.; Holm, J.; Sanadi, A.; Barsberg, S.; Ahrenfeldt, J.; Henriksen, U. Fuel pellets from biomass: The importance of the pelletizing pressure and its dependency on the processing conditions. Fuel 2011, 90, 3285–3290. [Google Scholar] [CrossRef] [Green Version]
  16. 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. Biorefin. 2011, 5, 683–707. [Google Scholar] [CrossRef]
  17. Filbakk, T.; Skjevrak, G.; Høibø, O.; Dibdiakova, J.; Jirjis, R. The influence of storage and drying methods for Scots pine raw material on mechanical pellet properties and production parameters. Fuel Process. Technol. 2011, 92, 871–878. [Google Scholar] [CrossRef]
  18. Kaliyan, N.; Morey, R.V. Factors affecting strength and durability of densified biomass products. Biomass Bioenergy 2009, 33, 337–359. [Google Scholar] [CrossRef]
  19. Wang, C.; Peng, J.; Li, H.; Bi, X.T.; Legros, R.; Lim, C.J.; Sokhansanj, S. Oxidative torrefaction of biomass residues and densification of torrefied sawdust to pellets. Bioresour. Technol. 2013, 127, 318–325. [Google Scholar] [CrossRef] [PubMed]
  20. Kotwaliwale, N.; Singh, K.; Kalne, A.; Jha, S.N.; Seth, N.; Kar, A. X-ray imaging methods for internal quality evaluation of agricultural produce. J. Food Sci. Technol. 2014, 51, 1–15. [Google Scholar] [CrossRef] [PubMed]
  21. Tenorio, C.; Moya, R.; Tomazello-Filho, M.; Valaert, J. Application of the X-ray densitometry in the evaluation of the quality and mechanical properties of biomass pellets. Fuel Process. Technol. 2015, 132, 62–83. [Google Scholar] [CrossRef]
  22. Tenorio, C.; Moya, R.; Tomazello-Filho, M.; Valaert, J. Quality of pellets made from agricultural and forestry crops in Costa Rican tropical climates. BioResources 2015, 10, 482–498. [Google Scholar] [CrossRef]
  23. Arias, B.; Pevida, C.; Fermoso, J.; Plaza, M.G.; Rubiera, F.; Pis, J.J. Influence of torrefaction on the grindability and reactivity of woody biomass. Fuel Process. Technol. 2008, 89, 169–175. [Google Scholar] [CrossRef]
  24. Peng, J.H.; Bi, X.T.; Sokhansanj, S.; Lim, C.J. Torrefaction and densification of different species of softwood residues. Fuel 2013, 111, 411–421. [Google Scholar] [CrossRef]
  25. Gaitán-Álvarez, J.; Moya, R. Characteristics and properties of pellet fabricated with torrefacted biomass of Gmelina arborea and Dipterix panamensis at different time. Rev. Chapingo Ser. Cienc. For. Ambient. 2016, 23, 325–337. [Google Scholar] [CrossRef]
  26. Moya, R.; Rodríguez-Zúñiga, A.; Puente-Urbina, A. Thermogravimetric and devolatilisation analysis for five plantation species: Effect of extractives, ash compositions, chemical compositions and energy parameters. Thermochim. Acta 2017, 647, 36–46. [Google Scholar] [CrossRef]
  27. Tenorio, C.; Moya, R. Evaluation of different approaches for the drying of lignocellulose residues. BioResources 2012, 7, 3500–3514. [Google Scholar]
  28. Aragón-Garita, S.; Moya, R.; Bond, B.; Valaert, J.; Tomazello Filho, M. Production and quality analysis of pellets manufactured from five potential energy crops in the Northern Region of Costa Rica. Biomass Bioenergy 2016, 87, 84–95. [Google Scholar] [CrossRef]
  29. González-Peña, M.M.; Hale, M.D. Colour in thermally modified wood of beech, Norway spruce and Scots pine. Part 1: Colour evolution and colour changes. Holzforschung 2009, 63, 385–393. [Google Scholar] [CrossRef]
  30. Shang, L.; Nielsen, N.P.K.; Dahl, J.; Stelte, W.; Ahrenfeldt, J.; Holm, J.K.; Henriksen, U.B. Quality effects caused by torrefaction of pellets made from Scots pine. Fuel Process. Technol. 2012, 101, 23–28. [Google Scholar] [CrossRef]
  31. SawaiI, T.; Katayama, I.; Ida, T.; Imamura, H.; Kajimoto, T. Estimation of energy properties of torrefied Japanese cedar with colorimetric values. Mech. Eng. J. 2017, 4, 16–00320. [Google Scholar] [CrossRef]
  32. Mišljenović, N.; Bach, Q.V.; Tran, K.Q.; Salas-Bringas, C.; Skreiberg, Ø. Torrefaction influence on pelletability and pellet quality of Norwegian forest residues. Energy Fuels 2014, 28, 2554–2561. [Google Scholar] [CrossRef]
  33. Carone, A.; Pantaleo, A.; Pellerano, A. Influence of process parameters and biomass characteristics on the durability of pellets from the pruning residues of Olea europea L. Biomass Bioenergy 2011, 35, 402–410. [Google Scholar] [CrossRef]
  34. Cao, L.; Yuan, X.; Li, H.; Li, C.; Xiao, Z.; Jiang, L.; Zeng, G. Complementary effects of torrefaction and co-pelletization: Energy consumption and characteristics of pellets. Bioresour. Technol. 2015, 185, 254–262. [Google Scholar] [CrossRef] [PubMed]
  35. Peng, J.H.; Bi, H.T.; Sokhansanj, S.; Lim, J.C. A study of particle size effect on biomass torrefaction and densification. Energy Fuels 2012, 26, 3826–3839. [Google Scholar] [CrossRef]
  36. 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] [PubMed]
  37. Pyle, O. Process for Producing Solid Industrial Fuel 1972. U.S. Patent 3950143A, 13 April 1976. [Google Scholar]
  38. Agar, D.A. A comparative economic analysis of torrefied pellet production based on state-of-the-art pellets. Biomass Bioenergy 2017, 97, 155–161. [Google Scholar] [CrossRef]
  39. Bissen, D. Biomass Densification, Document of Evaluation; Agricultural Utilization Research Institute, Zachry Energy Corporation: Minneapolis, MN, USA, 2009. [Google Scholar]
  40. Bergström, D.; Israelsson, S.; Öhman, M.; Dahlqvist, S.A.; Gref, R.; Boman, C.; Wästerlund, I. Effects of raw material particle size distribution on the characteristics of Scots pine sawdust fuel pellets. Fuel Process. Technol. 2008, 89, 1324–1329. [Google Scholar] [CrossRef]
  41. Larsson, S.H.; Thyrel, M.; Geladi, P.; Lestander, T.A. High quality biofuel pellet production from pre-compacted low density raw materials. Bioresour. Technol. 2008, 99, 7176–7182. [Google Scholar] [CrossRef] [PubMed]
  42. Kymäläinen, M.; Rautkari, L.; Hill, C.A. Sorption behaviour of torrefied wood and charcoal determined by dynamic vapour sorption. J. Mater. Sci. 2015, 50, 7673–7680. [Google Scholar] [CrossRef]
  43. Rudolfsson, M.; Borén, E.; Pommer, L.; Nordin, A.; Lestander, T. Combined effects of torrefaction and pelletization parameters on the quality of pellets produced from torrefied biomass. Appl. Energy 2017, 191, 414–424. [Google Scholar] [CrossRef]
  44. Acharjee, T.C.; Coronella, C.J.; Vasquez, V.R. Effect of thermal pretreatment on equilibrium moisture content of lignocellulosic biomass. Bioresour. Technol. 2011, 102, 4849–4854. [Google Scholar] [CrossRef] [PubMed]
  45. Kambo, H.S.; Dutta, A. Strength, storage, and combustion characteristics of densified lignocellulosic biomass produced via torrefaction and hydrothermal carbonization. Appl. Energy 2014, 135, 182–191. [Google Scholar] [CrossRef]
  46. Ahn, B.J.; Chang, H.S.; Lee, S.M.; Choi, D.H.; Cho, S.T.; Han, G.S.; Yang, I. Effect of binders on the durability of wood pellets fabricated from Larix kaemferi C. and Liriodendron tulipifera L. sawdust. Renew. Energy 2014, 62, 18–23. [Google Scholar] [CrossRef]
  47. Keipi, T.; Tolvanen, H.; Kokko, L.; Raiko, R. The effect of torrefaction on the chlorine content and heating value of eight woody biomass samples. Biomass Bioenergy 2014, 66, 232–239. [Google Scholar] [CrossRef]
  48. Chen, W.H.; Kuo, P.C. A study on torrefaction of various biomass materials and its impact on lignocellulosic structure simulated by a thermogravimetry. Energy 2010, 35, 2580–2586. [Google Scholar] [CrossRef]
  49. Nunes, L.J.R.; Matias, J.C.O.; Catalao, J.P.S. Wood pellets as a sustainable energy alternative in Portugal. Renew. Energy 2016, 85, 1011–1016. [Google Scholar] [CrossRef]
  50. Mani, S.; Tabil, L.G.; Sokhansanj, S. Effects of compressive force, particle size and moisture content on mechanical properties of biomass pellets from grasses. Biomass Bioenergy 2006, 30, 648–654. [Google Scholar] [CrossRef]
  51. Rhen, C.; Rolf, G.; Sjöström, M.; Wäterlund, I. Effects of raw material moisture content, densification pressure and temperature on some properties of Norway spruce pellets. Fuel Process. Technol. 2005, 87, 11–16. [Google Scholar] [CrossRef]
  52. Kumar, L.; Koukoulas, A.A.; Mani, S.; Satyavolu, J. Integrating torrefaction in the wood pellet industry: A Critical Review. Energy Fuels 2016, 31, 37–54. [Google Scholar] [CrossRef]
  53. 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]
  54. Aarseth, K.A.; Prestlokken, E. Mechanical properties of feed Pellets: Weibull analysis. Biosyst. Eng. 2003, 84, 349–361. [Google Scholar] [CrossRef]
Figure 1. Torrefied pellets at different temperatures and times of five types of biomass from woody tropical species: (a) light torrefaction; (b) Middle torrefaction; (c) Severe torrefaction.
Figure 1. Torrefied pellets at different temperatures and times of five types of biomass from woody tropical species: (a) light torrefaction; (b) Middle torrefaction; (c) Severe torrefaction.
Energies 10 01205 g001
Figure 2. Density of pellets torrefied at different temperatures and times of five types of biomass from woody tropical species: (a) Cupressus lusitanica; (b) Dipteryx panamensis; (c) Gmelina arborea; (d) Tectona grandis; (e) Vochysia ferruginea.
Figure 2. Density of pellets torrefied at different temperatures and times of five types of biomass from woody tropical species: (a) Cupressus lusitanica; (b) Dipteryx panamensis; (c) Gmelina arborea; (d) Tectona grandis; (e) Vochysia ferruginea.
Energies 10 01205 g002
Figure 3. Density variation profiles in longitudinal direction of pellets torrefied at different temperatures and times of five types of biomass from woody tropical species: (a) density variation in Cupressus lusitanica; (b) density variation in Dipteryx panamensis; (c) density variation in Gmelina arborea; (d) density variation in Tectona grandis; (e) density variation in Vochysia ferruginea.
Figure 3. Density variation profiles in longitudinal direction of pellets torrefied at different temperatures and times of five types of biomass from woody tropical species: (a) density variation in Cupressus lusitanica; (b) density variation in Dipteryx panamensis; (c) density variation in Gmelina arborea; (d) density variation in Tectona grandis; (e) density variation in Vochysia ferruginea.
Energies 10 01205 g003
Figure 4. Compression force variation in relation to the deformation of pellets torrefied at different temperatures and times for five types of biomass from woody tropical species: (a) compression force in Cupressus lusitanica; (b) compression force in Dipteryx panamensis; (c) compression force in Gmelina arborea; (d) compression force in Tectona grandis; (e) compression force in Vochysia ferruginea.
Figure 4. Compression force variation in relation to the deformation of pellets torrefied at different temperatures and times for five types of biomass from woody tropical species: (a) compression force in Cupressus lusitanica; (b) compression force in Dipteryx panamensis; (c) compression force in Gmelina arborea; (d) compression force in Tectona grandis; (e) compression force in Vochysia ferruginea.
Energies 10 01205 g004
Figure 5. Universal test machine (a) and the pellet device built for pellet manufacturing (b).
Figure 5. Universal test machine (a) and the pellet device built for pellet manufacturing (b).
Energies 10 01205 g005
Table 1. Moisture content, moisture absorption, and compression force of pellets torrefied at different temperatures and times for five types of biomass from woody tropical species.
Table 1. Moisture content, moisture absorption, and compression force of pellets torrefied at different temperatures and times for five types of biomass from woody tropical species.
SpeciesTemperature (°C)Time (minutes)Moisture Content (%)Moisture Absorption (%)Compression Force (N)Density Variation (%)
Cupressus lusitanica20082.71 A12.15 A929.09 A4.25 A
103.44 A12.46 A609.21 B4.67 A
121.28 B17.92 B323.20 C4.59 A
22581.65 B16.32 B416.83 C5.09 A
101.56 B12.22 A379.01 C5.25 A
120.90 B11.58 A370.91 C6.42 B
25083.14 A10.34 A752.48 A4.31 A
100.96 B9.75 C819.17 A4.70 A
121.73 B7.31 C672.33 B6.99 B
Dipteryx panamensis20085.05 A10.08 A757.92 A3.81 A
105.08 A9.22 A702.55 A2.84 A
121.82 B9.41 A804.96 A3.00 A
22581.19 C11.76 B341.63 B3.35 A
101.19 C10.89 A407.86 B4.03 A
120.73 C10.55 A409.95 B3.49 A
25080.67 C11.15 A832.50 A3.77 A
100.41 C9.28 A840.95 A4.47 A
120.43 C7.19 C682.79 A4.66 A
Gmelina arborea20084.19 A11.84 A579.62 A4.08 A
105.51 A11.17 A497.68 A5.18 A
123.98 A11.51 A582.56 A3.42 A
22582.69 B13.15 B191.78 B9.10 B
105.23 A11.44 A187.77 B3.22 A
122.02 B11.82 A164.57 B3.74 A
25080.92 C12.09 A394.78 C4.72 A
102.35 B10.62 C457.50 C4.85 A
122.71 B8.19 D332.06 C7.71 B
Tectona grandis20083.93 A11.32 A296.59 A3.64 A
104.64 A9.41 B268.53 A3.14 A
121.68 B11.98 A250.78 A3.18 A
22582.15 B10.45 A620.32 B3.35 A
101.27 B10.12 A344.14 A3.92 A
121.24 B10.74 A428.75 A3.56 A
25080.61 C9.82 A597.15 B3.62 A
102.25 B8.94 B506.88 B3.99 A
121.18 C7.21 C589.91 B3.49 A
Vochysia ferruginea20083.09 A9.82 A1052.41 A8.24 A
103.68 A9.82 A1468.97 B8.26 A
124.21 A9.69 A1434.28 B9.33 A
22580.87 B13.02 B623.97 C10.57 B
100.49 B11.26 A850.87 A12.55 B
120.70 B9.92 A730.27 C11.18 B
25080.60 B9.46 A1388.33 B6.74 A
102.18 A10.30 A1162.28 A6.68 A
123.08 A6.97 C867.46 A2.11 C
Table 2. Correlation matrix of the evaluated variables of torrefied pellets at different temperatures and times for five types of biomass from woody tropical species.
Table 2. Correlation matrix of the evaluated variables of torrefied pellets at different temperatures and times for five types of biomass from woody tropical species.
Cupressus Lusitanica
VariablesDensityCoefficient of VariationCompression ForceMoisture AbsorptionMoisture Content
Density1.00----
Coefficient of variation−0.32 **1.00---
Compression force0.04 NS−0.19 *1.00--
Moisture absorption0.14 NS0.04 NS−0.33 **1.00-
Moisture content−0.16 *−0.08 NS0.27 *−0.14 NS1.00
Dipteryx Panamensis
VariablesDensityCoefficient of VariationCompression ForceMoisture AbsorptionMoisture Content
Density1.00----
Coefficient of variation−0.18 *1.00---
Compression force0.21 *0.13 NS1.00--
Moisture absorption−0.07 NS−0.38 **−0.45 **1.00-
Moisture content−0.17 *−0.31 **0.15 NS0.04 NS1.00
Gmelina Arborea
VariablesDensityCoefficient of VariationCompression ForceMoisture AbsorptionMoisture Content
Density1.00----
Coefficient of variation−0.21 *1.00---
Compression force0.17 *−0.10 NS1.00--
Moisture absorption0.18 *−0.10 NS−0.06 NS1.00-
Moisture content−0.42 **1.10 NS0.42 **−0.15 NS1.00
Tectona Grandis
VariablesDensityCoefficient of VariationCompression ForceMoisture AbsorptionMoisture Content
Density1.00----
Coefficient of variation−0.29 *1.00---
Compression force0.16 *−0.03 NS1.00--
Moisture absorption−0.01 NS−0.16 *−0.19 *1.00-
Moisture content−0.27 *0.00 NS−0.28 *0.09 NS1.00
Vochysia Ferruginea
VariablesDensityCoefficient of VariationCompression ForceMoisture AbsorptionMoisture Content
Density1.00----
Coefficient of variation−0.55 **1.00---
Compression force−0.03 NS−0.23 *1.00--
Moisture absorption−0.37 **0.39 **−0.24 *1.00-
Moisture content0.32 **−0.26 *0.32 **−0.19 *1.00
Note: * significance at 95%, ** significance at 99%, NS significance not present.

Share and Cite

MDPI and ACS Style

Gaitán-Alvarez, J.; Moya, R.; Puente-Urbina, A.; Rodriguez-Zuñiga, A. Physical and Compression Properties of Pellets Manufactured with the Biomass of Five Woody Tropical Species of Costa Rica Torrefied at Different Temperatures and Times. Energies 2017, 10, 1205. https://doi.org/10.3390/en10081205

AMA Style

Gaitán-Alvarez J, Moya R, Puente-Urbina A, Rodriguez-Zuñiga A. Physical and Compression Properties of Pellets Manufactured with the Biomass of Five Woody Tropical Species of Costa Rica Torrefied at Different Temperatures and Times. Energies. 2017; 10(8):1205. https://doi.org/10.3390/en10081205

Chicago/Turabian Style

Gaitán-Alvarez, Johanna, Roger Moya, Allen Puente-Urbina, and Ana Rodriguez-Zuñiga. 2017. "Physical and Compression Properties of Pellets Manufactured with the Biomass of Five Woody Tropical Species of Costa Rica Torrefied at Different Temperatures and Times" Energies 10, no. 8: 1205. https://doi.org/10.3390/en10081205

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