• No results found

Effects of moisture content during densification of biomass pellets, focusing on polysaccharide substances

N/A
N/A
Protected

Academic year: 2022

Share "Effects of moisture content during densification of biomass pellets, focusing on polysaccharide substances"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

http://www.diva-portal.org

This is the published version of a paper published in Biomass and Bioenergy.

Citation for the original published paper (version of record):

Frodeson, S., Henriksson, G., Berghel, J. (2019)

Effects of moisture content during densification of biomass pellets, focusing on polysaccharide substances

Biomass and Bioenergy, 122: 322-330

https://doi.org/10.1016/j.biombioe.2019.01.048

Access to the published version may require subscription.

N.B. When citing this work, cite the original published paper.

This article is available under the terms of the Creative Commons Attribution License (CC BY). You may copy and distribute the article, create extracts, abstracts and new works from the article, alter and revise the article, text or data mine the article and otherwise reuse the article commercially (including reuse and/or resale of the article) without permission from Elsevier. You must give appropriate credit to the original work, together with a link to the formal publication through the relevant DOI and a link to the Creative Commons user license above. You must indicate if any changes are made but not in any way that suggests the licensor endorses you or your use of the work.

Permanent link to this version:

http://urn.kb.se/resolve?urn=urn:nbn:se:kau:diva-71411

(2)

Biomass and Bioenergy

journal homepage:www.elsevier.com/locate/biombioe

Research paper

Effects of moisture content during densification of biomass pellets, focusing on polysaccharide substances

Stefan Frodeson

a,∗

, Gunnar Henriksson

b

, Jonas Berghel

a

aEnvironmental and Energy Systems, Department of Engineering and Chemical Science, Karlstad University, SE-651 88, Karlstad, Sweden

bChemical Engineering, Department of Engineering and Chemical Science, Karlstad University, SE-651 88, Karlstad, Sweden

A R T I C L E I N F O

Keywords:

Wood pellets Densification Cellulose Hemicellulose Xylan Glucomannan

A B S T R A C T

In this study, we pelletized four different pure polysaccharides represented cellulose - Avicel, hemicelluloses - locus bean gum mannan and beech xylan and other polysaccharides - apple pectin, and three woods - pine, spruce and beech. All were pelletized at 100° in a single pellet press unit with different level of moisture content from 0 to 15%. The maximal friction force and work required for compression and friction was analyzed together with the pellet density and hardness. The results showed that xylan pellets completely changed in color at 10%

moisture content, and this also occurred to some extent with pectin pellets. The color of both Avicel and locus bean gum pellets were not affected at all. During compression, the results showed that water does not affect compression up to 5 kN, while above 5 kN water decreases the energy need for densification of Avicel, locus bean gum and woods. Above 5 kN the energy needs for compressing xylan and pectin increases with increased moisture content. The hardest pellets were produced from Avicel, while locus bean gum produced the weakest pellets. The study concludes that there is a significant difference in how water affects the two hemicelluloses, glucomannan and xylan, during densification.

1. Introduction

Increased utilization of biomass for new products, novel applica- tions, and energy purpose is necessary for a successful transition to- wards a sustainable bio-economy. However, the use and handling of large quantities of biomass poses challenges. Such challenges include high costs for low-density transports, the non-uniform size of biomass particles leading to difficulties with volume dose and molds may impair quality deterioration on the material. By drying the biomass and com- pressing it into pellets, the majority of these problems can be con- trolled, but it is important that the produced pellet retains its shape through all transports. To secure this, strong bonds are required, however, there is insufficient knowledge about how the bonding me- chanisms in the pellets correlate to the characteristics of the biomasses [1–3]. This means that pellet producers strive for a chemical compo- sition of the raw material as equal as possible, and this disadvantage the optimal utilization of all available raw materials. For an increased utilization of all types of biomasses, a pellet producer should be able to manage a varied raw material flow.

A traditional pelletizing process, from wet raw material to manu- factured pellets, includes several pre-treatment steps, such as drying, grinding, and conditioning, as well as post-treatment steps, such as

cooling and storage. The actual densification, when the roller wheels compress the biomass, can be divided into three subprocesses (Fig. 1) [4]: 1) The compression step where roller compresses the feedstock into a thin layer; 2) The flow step when the compressed layer under pressure flows into the die channels and is partly compressed further from the sides in the cone; 3) the friction step when the compressed feedstock is in the die channel [4].

What happens between the fibers during the densification process is described by Mani et al. (2004) as follows: particles rearranged them- selves to form a closely-packed mass where most of the particles retain the majority of their original properties, although energy is dissipated due to inter-particle and particle-to-wall friction. As the compaction pressure increases, particles are forced against each other while un- dergoing elastic and plastic deformation. This increases the inter-par- ticle contact area and as a result, bonding forces such as van der Waal's forces become effective. At higher pressures, internal pores within particles rupture until the density of the compacted bulk approaches its true or solid density [5]. Therefore, when the particles are reaching the die channel, the main part of the density increase is reached, leaving the final friction component. This is important for two reasons: i) to generate the right temperature so that strong bonds can be generated within the pellets, and ii) to create a backpressure sufficient for the

https://doi.org/10.1016/j.biombioe.2019.01.048

Received 13 November 2018; Received in revised form 30 January 2019; Accepted 31 January 2019

Corresponding author.

E-mail addresses:Stefan.Frodeson@kau.se(S. Frodeson),ghenrik@kth.se(G. Henriksson),Jonas.Berghel@kau.se(J. Berghel).

(3)

rollers to create their pressure [4,6]. The friction is generated in an active part of the channel [4], often referred to as the die press length.

The length of the active part is based on specific feedstock [6,7] and the backpressure increases exponentially as the pellet increases in length [8]. Thus, it is important to elucidate how and why different materials generate backpressure. As the backpressure and friction vary depending on the chemical composition of the biomass, pellet producers target a feedstock with a chemical composition that is as uniform as possible.

In order to cope with the transition to an expanded stream of bio- materials, one needs to know more about different types of biomasses and their pellet-forming abilities. A way to learn more is to study the components within the biomass chemical composition and to gain knowledge of their relative importance in the actual densification step.

This would thereby increase knowledge about which components have a major or minor impact. A general description of biomass chemical composition is that it can be divided into two main groups: macro- and low-molecular substances. In this case, low-molecular substances are both organic and inorganic, such as extractives and ashes, and macro- molecular substances are lignin, polysaccharides, and proteins. The group of polysaccharides includes cellulose and polyoses [9]. The polyoses within this study are categorized as hemicelluloses (xylan and glucomannan) and other polysaccharides, such as pectin, starch, and galactan. There is also a deviation between the polysaccharides, cor- related to the number of side chains connected to the functional groups.

None or few side chains means that the polysaccharide is rather linear or stiff, whereas numerous side chains means that the polysaccharide is rather flexible or highly branched [9]. The chemical composition of different biomasses is well defined in the literature [9–11], where it can be found that the hemicellulose of hardwood is mainly xylan, whereas that of softwood is more glucomannan.

With the objective of obtaining a greater understanding of the role pure biomass components play in the pelletizing process, the authors in an earlier study [6] had tested polysaccharides, lignin, protein, and extractives together with reference material for both softwood and hardwood. These were all pelletized in a single pellet press unit at dry conditions. One conclusion drawn was that there is a difference be- tween stiff polysaccharides in comparison with the flexible ones, and that the difference between xylan and glucomannan can explain the difference between the pelletization of hardwood and softwood. How- ever, in that specific set of tests, all substance was pelletized in an

almost completely dry condition, and when it comes to pelletizing biomasses, water has been found to be necessary [7,12,13].

Studies have postulated that an optimal moisture content (MC) for the compression of woody biomass in general is 6–12%, with a varia- tion between different types of biomasses [2,14–20]. This variation in MC also limits the possibilities of handling a variated biomass stream;

thus, if a pellet producer adds a new biomass, variation in optimal MC hampers the production capacity. If the MC is too low, no pellets are being produced – powder simply pours out of the die. If the MC is lower than optimal, the friction increases, which leads to an excessive tem- peratures in the die, and leads to rapid thermal decomposition of the pellet surface and to subsequent problematic consequences with the production process [21]. A high level of MC also negatively affects the pellet quality [2,15,18,22–25]. How water influences variation in pellet quality can be explained by the fact that: the incompressibility of water affects the agglomeration of the particles [26], the excessive internal pressure from steam is generated inside the pellets [15,21], water blocks the hydrogen bonds, which leads to occupied locations of par- ticle-particle binding and, therefore, a weakening of the hydrogen bonds [4,18,21,27]. There is also a difference if the feedstock is fresh or is being used after having been stored, and studies have shown that fresh raw material in general requires a lower level of MC, around 7–8%

[28], whereas stored material requires approximately 11–13% [15,28].

How the MC influences the strength of the bonds within the pellets has not yet been clearly established, despite the existence of some theories. According to Kaliyan et al. (2009) water may be acting as a bridge between the particles and may introduce cohesive forces to improve the bonding strength, which is possible if the distance between the particles is in the range of 0.5–0.74 nm [12], which corresponds to twice the hydrogen bond length plus the OeH length of water [13].

Water can also act as a plasticizer and increase the molecular mobility of the amorphous polymers, which increases the flow ability of hemi- celluloses and other low glass transition extractives, thereby improving bonding in pellets [2,19,22,23]. When the MC is too low, the funicular state will not be fully developed [26] and the gaps where the water molecules could bridge will not be filled, resulting in a limited bonding area [18,27].

However, as this variation in optimal MC causes pellet producers to strive for a uniform feedstock, more knowledge is required regarding how moisture affects densification. Earlier results from the authors [6]

showed that polysaccharides have a high level of impact on the pelle- tizing process. As this study was based on dry conditions, there are gaps in knowledge how water affects densification of specific biomass components.

The objective of this study is to increase the understanding of the role water plays during the densification process, with a focus on polysaccharides. The aim is to investigate the maximum friction force needed to overcome the backpressure, the work required for compres- sion and friction, pellet solid density, and pellet hardness. Four poly- saccharides, representing stiff and flexible polysaccharides, and three wood species, representing softwood and hardwood, were pelletized with varied amounts of water.

2. Method 2.1. Materials

In total, seven different materials were evaluated, two representing stiff polysaccharides (cellulose – test C1 and glucomannan – test HC2) and two flexible polysaccharides (xylan – test HC3 and pectin – test OP4), two softwoods (Scots pine Pinus sylvestris – SW5 and Norway spruce Picea abies – SW6), and one hardwood (Beech, Fagus sylvatica – HW7) (Table 1). The amounts of water were tested at six different MC levels – 0.0, 5.0, 7.5, 10.0, 12.5, and 15.0%.

The method for moistening involved 90 g of each material that was placed in 45 °C for 48 h, to ensure equal starting positions, and to Fig. 1. The densification process is divided into three subprocesses: 1) com-

pression, 2) flow, and 3) friction.

S. Frodeson et al.

(4)

ensure the material was as dry as possible without risking high tem- perature chemical impact. Following this, the material was divided into two parts, whereby 40% of the total amount of material was moistened in two stages: first to 7.5%, and then to 12.5%. The rest of the material was moistened in three stages, to 5.0%, 10.0%, and 15.0%. All test material was moistened according to the same method: the dry material was placed in a 0.25 L square plastic bottle, and water was added by spraying water into the bottle while it was being weighed, until the amount of water corresponded to the actual MC. Once the correct amount of water had been added, the bottle was sealed, and blended by being rotated in a vertical mode for 10 min at a rotational speed of 60 rpm. Next, the material was sealed in plastic bags and stored for 48 h. After this period of storage, 13 g of the material was sealed in a plastic bag until the tests were carried out. The rest of the material was remoistened to reach the next MC level by repeating the same process described above. For the test at 0% MC, 8 g of material was dried in 103 °C for 24 h, and the test was carried out directly after the samples were removed from the 103 °C, and kept in a plastic bag to prevent the absorption of moisture from the environment. The MC (%) of all sam- ples was measured according to the method described in SS-EN 14774-1 [29] on a wet basis.

All pure polysaccharides originate from biomasses and are equiva- lent materials commonly used in research studies to reflect their properties in a biomass [30–32]. However, they may have undergone small structural changes compared with their original state, during extraction to pure components. All the wood was sawn with a table saw and then milled in a Culatti Mikro Hammer Mill (DFH 48, Limmat- strasse, Zurich) with a sieve size of 2.0 mm to attain a powder with uniform particle size. All pure components were in powder form and pelletized as they were delivered. Particle distribution control was carried out on all materials and the main part size of all particles was less than 0.35 mm. The particle distribution showed an exceedingly small deviation, so that particle size variations was assumed to not have an effect on the study.

2.2. The densification

Pellets were produced in a single pellet unit located at the Environmental and Energy Systems at Karlstad University, Sweden. The single pellet press is fully described in Ref. [6] with the two deviations being that the hole diameter for this study was 8.4 mm, and that a 10.0 mm-long steel piston plug with a diameter of 8.0 mm was used at the bottom of the die hole to allow a longer distance to measure fric- tion. The temperature of the die was set at 100 °C and, after equilibrium was reached, the steel piston plug was placed at the bottom the die hole, followed by a 13 mm-long nylon plug with a diameter of 8.4 mm positioned above the steel plug. Next, a 1 g sample of the test material was added and finally, once again a nylon plug of the same size was placed above the sample. The two nylon plugs ensured an equal tem- perature at both ends of the pellets.

The samples were compressed by the piston at a velocity of 30 mm per minute, to a desired pressure of 14 kN. After a retention time of

10 s at full pressure, the piston ejected the pellets just after the bottom plate was removed and the steel plug had fallen out. Six pellets were produced for the 0% MC test and ten pellets were produced for all other test series. Directly after pelletizing, each pellet was cooled to ambient temperature with a small fan, and then stored in a sealed plastic bag at ambient room temperature for further testing and analysis.

2.3. Measuring

During densification, the force was logged three times per second.

The compression and friction work were calculated by integrating the force and time determined from logged data. In order to see in greater detail how moisture content affects compression, the compression phase was divided into two steps: time to increase the force from 1 to 5 kN, step 1, and from 5 to 14 kN, step 2. The friction work was de- termined in three steps over a total distance of 17.0 mm and began when the force reached 0.5 kN. The first step, which may have included some retardation of the nylon plug, corresponded to the first 1 mm; the second step, when the pellet was certainly moving inside the die, was 9 mm; finally, the last step involved the phase wherein some of the nylon plug left the die. The compression was depicted as Wcomp(J/g) and the friction work as Wfric(J/g) and Wfric(J/cm2), and pellet surface was based on cold pellet length and the hole diameter. The maximum friction force needed for the piston to expel the pellets was read as the highest value generated in the process. This value, depicted as Fmax

(kN), represents the maximal potential backpressure.

The pellets produced were analyzed by measuring pellet solid density (g/cm3), firstly by sanding the ends of the pellets, and then by measuring their volume and weight. The pellet hardness (kg) was measured using a KAHL motor-driven hardness tester (K3175-0011, Reinbek, Germany), with a 3.5 mm spring installed for the 0–100 kg range. All results are presented as the average value of ten pellets for solid density and six pellets for hardness.

3. Results

Pellets, produced from four different polysaccharides and three woods at different MC levels, are shown inFig. 2. Of all material at 0%

MC, only pellets made from Avicel, beech xylan, pine and beech wood were of sufficient quality that hardness tests could be performed. Of these, only Avicel generated a pellet with a hard surface, although Avicel pellets at 0% MC were much longer than remaining Avicel tests when water was added. Pellets from locus bean gum, pectin, and spruce produced at 0% MC fell apart, and could not be tested for hardness.

Locus bean gum and pectin at 0% MC was of such low strength that density not could be tested. As shown inFig. 2, the xylan pellets un- derwent a change in color and toward a more melted structure when the MC was increased, a change that started at the upper part of the pellet. At about 10% MC, the entire xylan pellets had turned dark brown compared with their original color of light beige and their structure had become glassier and more porous based on visual ob- servations. Similarly, some pectin pellets began to undergo structural Table 1

Tested polysaccharide substances and wood references.

Test Name Origin and comment

C1 Avicel Sigma-Aldrich (Darmstadt, Germany), Avicel®PH-101 product number 11365; short chain highly ordered cellulose from hydrolyzed hardwood chemical pulp

HC2 Locust bean gum mannan Sigma-Aldrich, product number G0753. Galactomannan polysaccharide from locust bean seeds with large similarities to the hemicellulose glucomannan

HC3 Beech xylan Virginiacare Limited (Waghaeusel, Germany), product number: 9014-63-5, Beech (Fagus sylvatica) wood (secondary xylem) OP4 Apple pectin Sigma-Aldrich, product number 93854; pectin from apples, probably rather similar to wood pectin

SW5 Pine wood Scots pine (Pinus sylvestris) representative reference from softwood SW6 Spruce wood Norway Spruce (Picea abies) representative reference from softwood HW7 Beech wood European beech (Fagus sylvatica) representative reference from hardwood

(5)

changes when MC was increased to test series 12.5% MC.

3.1. Densification

The compression process was divided into two steps. As shown in Fig. 3, during compression in step 1, the energy input appears to be independent of the MC. This differs from the range of step 2, where MC has a greater effect on the energy required for compression. The energy demand for compressing xylan increases with the addition of water until around 9–10% MC, where it begins to plateau and then decrease (Fig. 3a). As seen inTable 2, pectin also showed increased compression time requirement when water was added, but did not reach a max- imum.

Fig. 4shows typical compression curves at three different MC levels, and as shown inFig. 4b, the compression time for xylan increased with increased MC. Locus bean gum shows decreased compression times with increased MC (Fig. 4a). Spruce shows a greater decrease of com- pression time in comparison to beech, with increased MC (Fig. 4c and d).

InTable 2, data from moisture content, maximum friction force, energy requirement for compression, and friction is shown for all tests.

InFig. 5a, it is shown that the highest Fmaxis generated for xylan, while the locus bean gum reaches the lowest value. The flexible poly- saccharides, xylan and pectin, are affected by increased MC correlated to Fmaxand Wfric, while locus bean gum and Avicel appear to be un- affected (Fig. 5a and c). For the wood references, beech is the wood with the highest value for both Fmax(Fig. 5b) and Wfric(Fig. 5d), while data for spruce and pine has similarly lower values.

3.2. Pellet properties

As shown inFig. 6b, to obtain the highest density, an optimal MC for all wood references is approximately 5–8% MC. For hardness the op- timal MC is more scattered, with pine being 5–9%, spruce 6–10%, and beech 8–11% (Fig. 6d). The only polysaccharide that appears to have an optimal MC span for density within the MC range of this study is locus bean gum (Fig. 6a). The density of all the other polysaccharides in- creases or plateaus as the MC rises. The hardest pellets were measured among some of the pectin pellets at test series 15.0% MC, despite the high level of variation shown by the error bars. Hard pellets with the highest pellet density were produced from Avicel, as shown inFig. 6a and c.

Fig. 2. Photographs of the pellets ranging from zero MC to 15% MC; not all pellets are pictured.

Fig. 3. Work for compression divided into two steps, from 1 to 5 kN and 5 to 14 kN: a) showing the two hemicelluloses, locus bean gum and xylan; b) two wood references, spruce and beech.

S. Frodeson et al.

(6)

4. Discussion

In this study, the impact of moisture content during the densifica- tion process was investigated for pellets produced from four different polysaccharides – Avicel, locus bean gum, beech xylan, and pectin, and three wood references – Scots pine (Pinus sylvestris), Norway spruce (Picea abies), and beech (Fagus sylvatica) (Table 1). When drawing comparisons and undertaking evaluations, it is necessary to bear in mind that the polysaccharides here are not in the same form as they would be found in nature.

4.1. Effects of water on material structure and color

There is a perceptible difference between how xylan and pectin are affected by an increase in MC, compared to cellulose (Avicel) and glucomannan (locus bean gum). As shown inFig. 2, beech xylan pellets partially changes color from light beige to dark brown when the MC in the pellets is increased, and its structure undergoes a complete trans- formation around 10% MC. Changes in color and structure could also be seen for pectin. However, xylan and pectin differ with respect to the MC value at which changes begin to manifest. For xylan, the reaction began at the upper end of the pellet, around 5% MC, and by the time the MC reached 10%, all of the pellets had changed both in terms of their color

and structure. The material structure and color changes for pectin pellets began at the core at test series 12.5% MC for some of the pellets and, at test series 15.0% MC, all of the pellets had developed a dark- colored core, this could be seen in both ends of the pellet. During the hardness test when the pellets were crushed in pieces it was also clear that the dark-colored core and chemical change occurred throughout the pellet. The whole pellet, from one end to the other, had a dark- colored core.

The change in color suggests that the change in structure is more than merely a physical structural change. It is most likely a chemical change which resulted in both the color and the structure being altered, and the occurrence of partial vitrification The change in pellet color has been seen as when feedstock is steam-treated before pelletizing, the color changes to darker brown pellets [33,34]. In addition, the glass transition point can be a parameter of the chemical change. It has been shown that the high pressure changes the melting point to a new equilibrium level [35]. When steamed wood darkens, a suggested re- action is that the darkening is caused by reactions similar to the Mail- lard reaction [36] or by thermal reactions in carbohydrates such as the caramelization reactions [34,37]. However, such reactions often happen in the absence of moisture and involve condensation reactions, in contrast to the ones discussed in this paper which are stimulated by water. It is also unclear why only xylan and pectin, and not Table 2

Data for all tests showing MC for substance and pellet, work for compression, maximal friction force, and work for friction.

Substance Testserie (%) Moisture content (%) (wb) Wcomp(J/g) Fmax(kN) Wfric(J/g)

Powder Pellet 1–5 kN 5–14 kN Step 1 Step 2 Step 3

Avicel 0 0.0 0.0 32.9 ± 1.4 96.3 ± 16.8 8.6 ± 1.0 3.4 ± 2.1 65.6 ± 13.0 32.3 ± 6.2

5 5.3 4.4 44.2 ± 1.8 85.6 ± 5.3 5.8 ± 0.4 3.8 ± 1.5 43.9 ± 3.9 24.5 ± 2.4

7.5 6.6 5.9 48.5 ± 2.5 83.4 ± 10.9 5.6 ± 0.6 4.0 ± 1.5 40.2 ± 7.4 27.3 ± 5.7

10 9.6 7.9 45.6 ± 3.5 59.8 ± 12.3 5.2 ± 0.7 5.4 ± 1.6 32.2 ± 8.1 24.7 ± 6.5

12.5 11.2 9.3 49.5 ± 5.2 54.2 ± 6.9 4.7 ± 0.5 5.5 ± 0.6 22.8 ± 4.8 18.7 ± 2.8

15 12.7 10.4 47.1 ± 4.8 48.0 ± 3.1 5.2 ± 0.5 6.0 ± 0.5 29.7 ± 9.0 22.6 ± 5.8

Locus bean gum 0 0.0 0.0 16.8 ± 1.2 38.7 ± 4.2 0.8 ± 0.2 1.9 ± 1.5 5.8 ± 0.7 4.4 ± 0.7

5 6.0 5.8 18.4 ± 0.8 31.1 ± 1.5 0.3 ± 0.0 0.9 ± 0.3 2.0 ± 0.3 1.8 ± 0.3

7.5 8.0 8.0 18.3 ± 0.8 30.6 ± 1.7 0.3 ± 0.1 1.0 ± 0.5 1.8 ± 0.5 1.3 ± 0.4

10 10.5 9.8 19.1 ± 1.0 23.4 ± 3.2 0.2 ± 0.0 0.7 ± 0.2 1.5 ± 0.2 0.9 ± 0.1

12.5 12.7 11.4 13.8 ± 1.0 21.2 ± 1.6 0.2 ± 0.0 0.6 ± 0.3 1.3 ± 0.3 0.8 ± 0.2

15 13.3 12.0 12.3 ± 0.6 19.3 ± 0.9 0.2 ± 0.0 0.7 ± 0.1 1.4 ± 0.1 0.9 ± 0.1

Beech xylan 0 0.0 0.0 9.2 ± 4.9 53.4 ± 9.3 7.9 ± 1.4 3.3 ± 2.0 68.1 ± 12.5 36.4 ± 7.8

5 5.1 5.4 16.3 ± 1.9 49.5 ± 4.8 9.8 ± 1.5 3.0 ± 1.6 61.5 ± 7.3 17.0 ± 5.6

7.5 7.3 7.2 12.6 ± 2.4 65.6 ± 11.1 14.6 ± 2.5 1.9 ± 0.5 84.5 ± 20.8 40.7 ± 18.3

10 9.4 9.5 15.3 ± 1.2 88.5 ± 8.0 12.1 ± 1.3 3.1 ± 2.8 83.6 ± 12.5 45.9 ± 19.2

12.5 12.5 11.1 14.2 ± 2.3 85.5 ± 13.0 9.1 ± 0.8 2.9 ± 1.4 70.9 ± 6.5 28.4 ± 3.6

15 12.6 11.3 23.6 ± 6.8 75.2 ± 16.0 7.5 ± 0.9 7.8 ± 1.9 64.7 ± 5.3 33.6 ± 6.7

Apple pectin 0 0.0 0.0 16.0 ± 2.2 48.2 ± 5.5 8.3 ± 2.3 4.0 ± 1.2 66.2 ± 16.2 47.6 ± 13.1

5 4.2 5.6 16.3 ± 3.3 57.1 ± 3.8 8.5 ± 1.1 2.1 ± 0.5 73.1 ± 8.2 51.3 ± 8.6

7.5 8.1 7.6 13.4 ± 2.4 56.7 ± 2.5 9.4 ± 0.6 2.2 ± 0.7 87.0 ± 5.8 52.2 ± 10.9

10 10.9 8.5 17.3 ± 2.5 54.8 ± 6.3 9.2 ± 1.6 3.8 ± 2.1 82.1 ± 16.7 49.9 ± 10.4

12.5 13.8 10.9 15.0 ± 2.3 56.3 ± 8.0 10.7 ± 0.9 3.2 ± 1.4 83.5 ± 10.7 46.8 ± 12.5

15 14.0 12.2 16.2 ± 2.6 64.9 ± 10.2 10.6 ± 0.8 3.6 ± 2.0 73.4 ± 16.9 42.7 ± 18.5

Scots pine 0 0.0 0.0 46.1 ± 3.5 73.2 ± 7.9 4.3 ± 1.0 5.5 ± 0.8 34.2 ± 10.3 17.2 ± 3.7

5 5.4 5.4 47.1 ± 5.4 56.8 ± 6.0 3.6 ± 0.3 4.3 ± 0.9 32.8 ± 5.1 21.3 ± 3.4

7.5 8.0 7.3 51.8 ± 5.5 42.4 ± 3.7 3.1 ± 0.4 4.0 ± 0.6 25.2 ± 4.1 12.2 ± 1.8

10 10.9 8.3 46.7 ± 6.4 34.0 ± 3.4 1.8 ± 0.2 2.7 ± 0.3 13.6 ± 2.2 6.5 ± 0.8

12.5 12.9 9.8 39.5 ± 9.9 25.5 ± 3.6 1.7 ± 0.5 2.5 ± 3.6 12.3 ± 4.6 4.6 ± 1.4

15 13.4 9.8 30.7 ± 4.9 22.8 ± 2.0 2.2 ± 0.3 3.5 ± 0.4 18.4 ± 2.6 10.9 ± 1.2

Norwegien spruce 0 0.0 0.0 22.3 ± 2.3 59.0 ± 2.8 6.8 ± 0.5 2.1 ± 0.6 41.4 ± 8.2 39.7 ± 7.6

5 5.3 4.6 25.3 ± 1.8 51.7 ± 3.8 2.4 ± 0.2 2.2 ± 0.3 19.4 ± 2.6 13.3 ± 1.5

7.5 7.3 6.0 25.0 ± 3.1 58.0 ± 6.4 3.2 ± 0.6 2.2 ± 0.8 24.8 ± 4.3 18.4 ± 4.4

10 5.1 4.6 24.8 ± 1.8 49.9 ± 2.4 2.5 ± 0.4 2.2 ± 0.3 17.2 ± 3.3 15.3 ± 2.4

12.5 11.7 9.8 21.8 ± 2.7 41.6 ± 3.1 1.7 ± 0.6 1.8 ± 0.5 11.7 ± 4.1 10.3 ± 3.9

15 15.3 15.3a 18.7 ± 0.8 24.8 ± 2.0 0.8 ± 0.3 0.8 ± 0.2 5.0 ± 1.6 4.7 ± 2.0

Beech 0 0.0 0.0 24.8 ± 2.1 77.0 ± 5.6 9.0 ± 1.0 2.3 ± 1.2 69.2 ± 8.0 58.9 ± 6.6

5 6.0 5.1 23.3 ± 3.0 75.4 ± 7.6 4.9 ± 0.6 3.0 ± 1.7 46.1 ± 7.4 30.8 ± 4.6

7.5 8.7 7.1 24.8 ± 4.5 70.5 ± 9.0 6.2 ± 0.4 2.2 ± 0.9 47.6 ± 6.3 41.0 ± 3.8

10 11.5 9.1 25.0 ± 2.7 57.6 ± 4.4 4.1 ± 1.0 2.3 ± 0.8 32.3 ± 7.1 26.4 ± 6.2

12.5 11.9 9.7 25.8 ± 3.1 50.9 ± 2.7 3.8 ± 0.4 2.5 ± 0.5 24.2 ± 5.1 24.5 ± 3.1

15 14.9 11.7 22.3 ± 2.7 35.7 ± 3.6 1.7 ± 0.4 1.5 ± 0.4 11.3 ± 2.8 10.6 ± 2.8

a Data for spruce pellets at 15% failed and thus the moisture content of the pellet in the current test series is reported to be the same as for the powder.

(7)

glucomannan and cellulose, undergo this reaction. One point of com- monality between pectin and xylan – and which differs from the other polysaccharides tested – is that they contain uronic acid residues.

Uronic acids are known to be able to undergo reactions such as beta eliminations [38]. This could be the “starting point” of the reactions leading to color change, but requires further investigation.

The fact that water increases the mobility and fluidity of the hemicelluloses and other low glass transition extractives has been em- pirically verified in earlier studies [2,19,22,23]. However, the results from the research conducted in this study show that it is important to point out which component the hemicellulose consists of. Hemi- celluloses cannot be defined as a homogeneous substance [9], but generalizations can be made about their polysaccharide components.

Beech xylan and locus bean gum mannan exhibited major differences in color and structure changes (Fig. 2). It also appears clear that water plays an important role in this change.

4.2. Effects of water on the compressibility, friction, and backpressure Water's role as a plasticizer and softener has been proven in pre- vious studies [2,19,22,23]. In this study, the compression stage was divided into two steps, between 1–5 kN and 5–14 kN, and there does not appear to be an MC effect for the entire process. As shown inFig. 3, the energy for compression between 1 and 5 kN is not affected by MC and this trend is the same for all tested materials (Table 2). The influence of water between 5 and 14 kN is clear, as for the stiffer polysaccharides – Avicel and locus bean gum – as well as for the wood references, water

decreases the energy requirement. In contrast, the more flexible poly- saccharides – xylan and pectin – have an increase in energy demand as well as time needed for compression with increased MC (Fig. 3and Table 2). One explanation for the deviation compared with the other polysaccharides and the woods may be that the more flexible poly- saccharides, xylan and pectin, require higher energy for compression by virtue of their greater elasticity. The increased demand of energy can also be explained by the structural chemical changes discussed in 4.1 and shown inFig. 2. It is also likely that these deviations explain the increased variation in errors bars that xylan shows in Fig. 3a when water are added and the pressure are above 100 MPA. Regardless of whether the material is of high elasticity or not, it can be inferred from the results that the effect of water as a plasticizer is more pronounced over 100 MPa. The results from this study also show that the required energy as well as time for compression will increase with an increased amount of more flexible polysaccharides.

When it comes to friction and the influence of water, the effect is more diversified, as shown inFig. 5a and c, the polysaccharides do not react in the same way to a rise in MC. As shown inFig. 5c, a rise in MC does not appear to affect the friction for Avicel and locus bean gum, while increased MC increases the energy demand for xylan and pectin until about 8% MC where the friction starts to decline. However, even if the energy need for xylan and pectin is decreasing, is constantly higher then Avicel and locus bean gum. As shown inFig. 5b and d, water acts as a lubricant for all the wood references. Beech is the wood reference which generates the highest maximal backpressure and friction work, independent of MC (Fig. 5c and d). That hardwood generates high Fig. 4. Typical compression curves at 5, 10, and 15% moisture content: a) locus bean gum; b) beech xylan; c) spruce wood; and d) beech wood.

S. Frodeson et al.

(8)

backpressure confirms with the findings of earlier studies [6,8] and hardwood includes higher levels of xylan compared to glucomannan. It can be concluded that there is a difference between the hemicellulose components, xylan and glucomannan, with regards to the backpressure and friction generated, and this difference is affected by the amount of water.

4.3. Effects of water on pellet properties

The effect of water on the pellet density and hardness varies more for polysaccharides than for the wood references. The highest density is obtained for Avicel, followed by locus bean gum, beech xylan, and pectin (Fig. 6a). Of these components, only locus bean gum seems to have a maximum correlation to density – around 8% MC. Avicel seems to plateau, while both beech xylan and pectin still increase in density with increased MC. This is likely to mean that the optimal MC for these polysaccharides, given the highest density, is higher than the range of moisture for this study. This may be an explanation for why xylan-rich materials such as corn and barley exhibit higher optimal MC [39].

Nonetheless, this is not evident in the present study. As shown in Fig. 6b, beech, with a high xylan content in hemicellulose, has the same optimum MC for highest density as the other wood references. The explanation of corn and barley's higher MC optimum is likely due to a combination of a high content of flexible polysaccharides, such as

xylan, and starch together with protein; however, this must be in- vestigated further.

The hardest pellets were produced from Avicel and pectin, while the poorest pellets were from locus bean gum (Fig. 6c). The hardest pellets from woods were from pine and beech, while spruce generated the weakest pellets (Fig. 6d). However, as shown inFig. 6c, there is sub- stantial variation in hardness data for pectin at high MC, meaning that some of the pellets were significantly harder while others were weaker.

For pectin pellets, there was a relationship between the size of the core and the hardness of the pellet. Although, this means that the material changes for pectin generated hard pellets, in contrast to xylan pellets, which generated a pellet whose hardness decreased as the material changed. As shown inFig. 6d, beech wood generated the hardest pellets at an MC range of 7.5–11% and it is likely that xylan's changes in structure act as “solid bridges” between the particle sites and introduce cohesive forces to improve the bonding strength, in beech wood.

However, this has not been analyzed in this study and further work is needed.

4.4. Comments on the planned test series

There were some deviations worth noting in support of other studies as well as to explain results that deviated from the planned test series, including pellet MC (Table 2). For the study, we chose to moisten the Fig. 5. Maximal friction force and total friction work per square centimeter: a) maximal friction force for polysaccharides; b) maximal friction force for wood references; c) friction work for polysaccharides; and d) friction work for wood references.

(9)

samples with a spray bottle and to mix them horizontally. It can be difficult to add the exact amount of water by spray bottle, especially if the samples are as small as 10 g. We attempted to avoid this by adding water in stages with a larger amount of material. However, by the end, the samples were small and a high level of accuracy became important.

Another choice of method could be to use a climate chamber with a different equilibrium of moisture content. Regardless, both methods will require both accuracy when weighing and time for equilibrium distribution. The reason for the deviations from the planned test series associated with MC in pellets in this study is likely correlated to the handling of samples. Some pellets had a lower MC than planned and it is likely that the samples had dried somewhat during the pelletizing process, during storage, or while their properties were being tested. It is also worth noting that the control of the single pellet production cor- related to the speed was manually operated, which means that it was difficult to shut it down directly at 14 kN, and this force was slightly exceeded before it flared out. However, none of these deviations are expected to affect the results and the conclusions, which means that all data is reliable.

4.5. Conclusions

The results have shown that there is variation in how the poly- saccharides are affected by changes in the MC, with the difference

between glucomannan and xylan being the most important. The dif- ference within the hemicellulose's components – xylan and gluco- mannan – can explain the difficulty of changing biomaterials during pelletization. When water is added to glucomannan and xylan, the energy and time need for compression as well as backpressure decreases and increases for each, respectively. The results also showed that the flexible polysaccharides – xylan and pectin – change in terms of color and structure when water was added, which is not the case for the stiffer polysaccharides – Avicel and locus bean gum. Based on these results, conclusions can be drawn that biomasses with high degrees of flexible polysaccharides are more sensitive to the amount of MC. This knowledge increases the understanding of which components in the biomass have major or minor effects during the densification process.

However, more research must be carried out before it is possible to fully understand the roles of all components within the chemical composi- tion of biomasses.

Funding

This work was supported by the Faculty of Health, Science and Technology at Karlstad University, Sweden.

Fig. 6. Pellet solid density and hardness at different MC values for all tests: a) Pellet solid density for Avicel, locus bean, xylan, and pectin; b) Pellet solid density for pine, spruce, and beech; c) Hardness for Avicel, locus bean, xylan, and pectin; and d) Hardness for pine, spruce, and beech.

S. Frodeson et al.

(10)

Acknowledgements

The authors would like to acknowledge Lars Pettersson at Environmental and Energy Systems at Karlstad University for labora- tory support.

References

[1] S. Mani, L.G. Tabil, S. Sokhansanj, Effects of compressive force, particle size and moisture content on mechanical properties of biomass pellets from grasses, Biomass Bioenergy 30 (7) (2006) 648–654.

[2] W. Stelte, A.R. Sanadi, L. Shang, J.K. Holm, J. Ahrenfeldt, U.B. Henriksen, Recent developments in biomass pelletization – a Review, BioResources 7 (3) (2012) 4451–4490.

[3] Á. Ramírez-Gómez, Research needs on biomass characterization to prevent hand- ling problems and hazards in industry, Part. Sci. Technol. 34 (4) (2016) 432–441.

[4] N.P.K. Nielsen, Importance of Raw Material Properties in Wood Pellet Production – Effects of Differences in Wood Properties for the Energy Requirements of Pelletizing and the Pellet Quality, Forest & Landscape Denmark, Faculty of Life Sciences, University of Copenhagen, Copenhagen, 2009.

[5] S. Mani, L.G. Tabil, S. Sokhansanj, Evaluation of compaction equations applied to four mass species, Can. Biosyst. Eng. 46 (3) (2004) 55–61.

[6] S. Frodeson, G. Henriksson, J. Berghel, Pelletizing pure biomass substances to in- vestigate the mechanical properties and bonding mechanisms, BioResources 13 (1) (2018).

[7] N.P.K. Nielsen, D.J. Gardner, T. Poulsen, C. Felby, Importance of temperature, moisture content, and species for the conversion process of wood residues into fuel pellets, Wood Fiber Sci. 41 (2009) 414–425.

[8] J.K. Holm, U.B. Henriksen, J.E. Hustad, L.H. Sørensen, Toward an understanding of controlling parameters in softwood and hardwood pellets production, Energy Fuels 20 (6) (2006) 2686–2694.

[9] D. Fengel, G. Wegener, Wood: Chemistry, Ultrastructure, Reactions, 2nd Berlin, New York, (1989).

[10] S.V. Vassilev, D. Baxter, L.K. Andersen, C.G. Vassileva, An overview of the chemical composition of biomass, Fuel 89 (5) (2010) 913–933.

[11] S.A. Rydholm, Pulping Processes, Robert Krieger Publishing Co, Malabar, 1985.

[12] N. Kaliyan, R. Vance Morey, Factors affecting strength and durability of densified biomass products, Biomass Bioenergy 33 (3) (2009) 337–359.

[13] R. Samuelsson, S.H. Larsson, M. Thyrel, T.A. Lestander, Moisture content and sto- rage time influence the binding mechanisms in biofuel wood pellets, Appl. Energy 99 (2012) 109–115.

[14] Y. Huang, M. Finell, S. Larsson, X. Wang, J. Zhang, R. Wei, L. Liu, Biofuel pellets made at low moisture content – influence of water in the binding mechanism of densified biomass, Biomass Bioenergy 98 (2017) 8–14.

[15] T. Filbakk, G. Skjevrak, O. Høibø, J. Dibdiakova, R. Jirjis, The influence of storage and drying methods for Scots pine raw material on mechanical pellet properties and production parameters, Fuel Process. Technol. 92 (5) (2011) 871–878.

[16] I. Obernberger, G. Thek, Physical characterisation and chemical composition of densified biomass fuels with regard to their combustion behaviour, Biomass Bioenergy 27 (Compendex) (2004) 653–669.

[17] Q.N. Nguyen, A. Cloutier, A. Achim, T. Stevanovic, Effect of process parameters and raw material characteristics on physical and mechanical properties of wood pellets made from sugar maple particles, Biomass Bioenergy 80 (2015) 338–349.

[18] N.P.K. Nielsen, D.J. Gardner, T. Poulsen, C. Felby, Importance of temperature, moisture content, and species for the conversion process of wood residues into fuel pellets, Wood Fiber Sci. 41 (4) (2009) 414.

[19] W. Stelte, J.K. Holm, A.R. Sanadi, S. Barsberg, J. Ahrenfeldt, U.B. Henriksen, Fuel pellets from biomass: the importance of the pelletizing pressure and its dependency

on the processing conditions, Fuel 90 (11) (2011) 3285–3290.

[20] H. Li, L.-B. Jiang, C.-Z. Li, J. Liang, X.-Z. Yuan, Z.-H. Xiao, Z.-H. Xiao, H. Wang, Co- pelletization of sewage sludge and biomass: the energy input and properties of pellets, Fuel Process. Technol. 132 (2015) 55–61.

[21] S.H. Larsson, M. Thyrel, P. Geladi, T.A. Lestander, High quality biofuel pellet pro- duction from pre-compacted low density raw materials, Bioresour. Technol. 99 (15) (2008) 7176–7182.

[22] M. Puig-Arnavat, L. Shang, Z. Sárossy, J. Ahrenfeldt, U.B. Henriksen, From a single pellet press to a bench scale pellet mill — pelletizing six different biomass feed- stocks, Fuel Process. Technol. 142 (2016) 27–33.

[23] W. Stelte, C. Clemons, J.K. Holm, J. Ahrenfeldt, U.B. Henriksen, A.R. Sanadi, Thermal transitions of the amorphous polymers in wheat straw, Ind. Crop. Prod. 34 (1) (2011) 1053–1056.

[24] M.T. Carone, A. Pantaleo, A. Pellerano, Influence of process parameters and bio- mass characteristics on the durability of pellets from the pruning residues of Olea europaea L, Biomass Bioenergy 35 (1) (2011) 402–410.

[25] C. Rhén, R. Gref, M. Sjöström, I. Wästerlund, Effects of raw material moisture content, densification pressure and temperature on some properties of Norway spruce pellets, Fuel Process. Technol. 87 (1) (2005) 11–16.

[26] H. Rumpf, The strength of granules and agglomerate, Agglomeration (1962) 379–418.

[27] L. Back Ernst, The bonding mechanism in hardboard manufacture review report, Holzforschung – Int. J. Biol. Chem. Phys. Technol. Wood (1987) 247.

[28] M. Arshadi, R. Gref, P. Geladi, S.-A. Dahlqvist, T. Lestander, The influence of raw material characteristics on the industrial pelletizing process and pellet quality, Fuel Process. Technol. 89 (12) (2008) 1442–1447.

[29] S.S.I. SIS, SS-EN 14774-1, Solid Biofuels – Determination of Moisture Content – Oven Dry Method – Part 1: Total Moisture – Reference Method, SIS Förlag AB, 118 80, Stockholm, Sweden, Stockholm, Sweden, 2009.

[30] R.L. Kudahettige-Nilsson, H. Ullsten, G. Henriksson, Plastic composites made from glycerol, citric acid, and forest components, BioResources 13 (3) (2018) 6600–6612.

[31] R. Bi, J. Berglund, F. Vilaplana, L.S. McKee, G. Henriksson, The degree of acet- ylation affects the microbial degradability of mannans, Polym. Degrad. Stabil. 133 (2016) 36–46.

[32] A. Abbaszadeh, W. MacNaughtan, G. Sworn, T.J. Foster, New insights into xanthan synergistic interactions with konjac glucomannan: a novel interaction mechanism proposal, Carbohydr. Polym. 144 (2016) 168–177.

[33] P.S. Lam, S. Sokhansanj, X. Bi, C.J. Lim, S. Melin, Energy input and quality of pellets made from steam-exploded Douglas fir (Pseudotsuga menziesii), Energy Fuels 25 (4) (2011) 1521–1528.

[34] P. Lam, S. Sokhansanj, X. Bi, C. Lim, Colorimetry applied to steam-treated biomass and pellets made from western Douglas fir (Pseudotsuga menziesii L.), Transactions of the ASABE 55 (2) (2012) 673–678.

[35] J.S. Tumuluru, C.T. Wright, J.R. Hess, K.L. Kenney, A review of biomass densifi- cation systems to develop uniform feedstock commodities for bioenergy applica- tion, Biofuels, Bioprod. Biorefin. 5 (6) (2011) 683–707.

[36] M. Sehlstedt-Persson, Colour responses to heat-treatment of extractives and sap from pine and spruce, International IUFRO Wood Drying Conference: 25/08/2003- 29/08/2003, Transilvania University of Brasov, Faculty of Wood Industry, 2003, pp. 459–464.

[37] M. Roe, R. Faulks, Color development in a model system during frying: role of individual amino acids and sugars, J. Food Sci. 56 (6) (1991) 1711–1713.

[38] A. Teleman, M. Siika-aho, H. Sorsa, J. Buchert, M. Perttula, T. Hausalo, M. Tenkanen, 4-O-Methyl-β-idopyranosyluronic acid linked to xylan from kraft pulp: isolation procedure and characterisation by NMR spectroscopy, Carbohydr.

Res. 293 (1) (1996) 1–13.

[39] N. Ungureanu, V. Vladut, G. Voicu, M.-N. Dinca, B.-S. Zabava, Influence of biomass moisture content on pellet properties – review, 17th International Scientific Conference Engineering for Rural Development, Jelgava, LATVIA.

References

Related documents

Stöden omfattar statliga lån och kreditgarantier; anstånd med skatter och avgifter; tillfälligt sänkta arbetsgivaravgifter under pandemins första fas; ökat statligt ansvar

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Däremot är denna studie endast begränsat till direkta effekter av reformen, det vill säga vi tittar exempelvis inte närmare på andra indirekta effekter för de individer som

Both Brazil and Sweden have made bilateral cooperation in areas of technology and innovation a top priority. It has been formalized in a series of agreements and made explicit

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

I regleringsbrevet för 2014 uppdrog Regeringen åt Tillväxtanalys att ”föreslå mätmetoder och indikatorer som kan användas vid utvärdering av de samhällsekonomiska effekterna av

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating