Elsevier

Energy

Volume 207, 15 September 2020, 118324
Energy

Impact of post-torrefaction process on biochar formation from wood pellets and self-heating phenomena for production safety

https://doi.org/10.1016/j.energy.2020.118324Get rights and content

Highlights

  • Self-heating phenomena are investigated using three different scenarios.

  • Three scenarios under different post-torrefaction processes are taken into account.

  • Self-heating occurs when oxidative post-torrefaction with keeping temperature is operated.

  • Oxidative post-torrefaction without keeping temperature intensifies fuel’s calorific value.

  • XPS analysis reveals the depletion of the CO functional group in biomass after torrefaction.

Abstract

The advancement of torrefaction in the industry raises safety issues lately. This work aims to investigate the impact of post-torrefaction upon biochar formation and the self-heating phenomena of torrefied wood pellets. Three operating scenarios are taken into account. Specifically, dried wood pellets are first torrefied in nitrogen at various temperatures (200–300 °C) followed by immediate exposure to N2 or air with/without keeping temperature, namely, the post-torrefaction, to examine the self-heating behavior. Meanwhile, the influences of torrefaction temperature, duration, and self-heating on biochar’s characteristics are analyzed. It is found that self-heating with the highest temperature rise of 61 °C is triggered when the biochars are in an oxidative post-torrefaction process with keeping temperature. Alternatively, oxidative post-torrefaction without keeping temperature can efficiently intensify the pellets’ calorific value up to 42%. The X-ray photoelectron spectroscopy analysis reveals the depletion of the CO functional group. It is concluded that uncontrolled self-heating in large scale production will lead to combustion and make biochar unsafe. Alternatively, if heat accumulation inside the biochar piles can be prevented, the oxidative environment that causes self-heating can open a lot of research opportunities that can help in the advancement of torrefaction technologies for biomass as an alternative source of energy.

Introduction

Currently, the majority of our energy consumption is coming from fossil fuels [1]. Coal plays an important role in fossil fuels and is more efficient than biomass. In addition to combustion, coal gasification is another economical choice to produce synthesis gas [2]. On the other hand, Chen et al. [3] discussed that considering the depletion of resources, there were alternative fuels that could be a potential substitute for coal. Biomass energy or bioenergy is considered as one of the promising resources that can match the characteristics of fossil fuels for GHG emission reduction. As emphasized by Tillman et al. [4], the most carbon-efficient use of biomass materials is by co-firing with coal.

Torrefaction is one of the pretreatment technologies to improve the properties of biomass as fuel. Torrefaction has been a topic of interest for the past years but a very limited number of plants had it on a full commercial scale [5]. Table 1 shows a summary of some torrefaction studies using different biomass materials, reactors, and operating parameters for evaluation. Torrefaction is a mild pyrolysis process that is typically used for upgrading the properties of woody (softwood and hardwood) and algal biomass. It is usually carried out at 200–300 °C without the presence of oxygen [6,7]. The prime product of torrefaction is biochar; biochar is also the by-product of the pyrolysis which has several possible applications from solid fuels to adsorbents. Chen et al. [8] found that biochar produced in a high torrefaction temperature had a comparable energy density to low ranking coals. In comparison to raw biomass, biochar has a better energy density, lower toxicity levels from combustion, higher grindability, and low water affinity [9]. These characteristics and properties made the biochar a good alternative to coal.

For the utilization of fuels such as coal and biomass, their self-heating followed by self-ignition is considered a serious problem for fuel producers and users due to safety issues. The advancement of torrefaction studies also raises safety issues. The study of Cocchi [10] showed that combustible materials were prone to oxidation and might promote self-heating and self-ignition. The self-heating of biochar can affect its properties such as energy yield, enhancement factor, and ash content. Without proper prevention, it can raise safety issues, particularly on large stockpiles, storage, and transportation. Several methods have been conducted to identify and prevent the self-heating of coal. The common tests used in industry and academia to study coal self-heating are outlined in Table 2. In the present study, a horizontal tube furnace was used to identify the onset temperature-time runway of the biochar.

According to the study of Jones et al. [11], coal and biomass share similar properties including self-heating propensity. The self-heating phenomenon usually starts at a low temperature and gradually increases with the heat released by the reaction. Cruz et al. [12] showed that at higher temperatures, the degree of self-heating was higher with enough oxygen exposure. The self-heating phenomenon occurs when there are exothermic reactions [13] like oxidation reactions. A temperature rise occurs when the heat release rate is faster than the heat dissipation rate [6,14]. The dissipation rate can be increased by purging of inert gas to overcome the heat release from the biochar. The volume of the piled material affects the self-heating propensity of biomass, depending on the dimension of the storage which affects the dissipation rate, and the biomass can be stored safely with minimal controlled storage conditions [15]. Large volume of biochar (greater than 1 ton), on the other hand, may experiences a self-heating phenomenon because of the lower dissipation rate of heat through the large stockpile boundaries. The boundary serves as a thermal blanket which hinders the heat dissipation rate. Moisture content and humidity also affect the self-heating propensity phenomenon. Beamish et al. [16] explained that self-heating could be significantly delayed by the level of moisture content because the heat released by oxidation was dissipated by the evaporation of moisture. Kiel et al. [17] had identified the self-heating phenomenon as the cause of a fire incident in the storage after the torrefaction process [18]. Hazard and risk issues are now considered in that the increase in temperature can lead to heat generation, and spontaneous ignition can lead to self-combustion, the critical state of self-heating. Evangelista et al. [6] observed that self-combustion was a known scenario for coal. With coal and biomass having some similar properties, self-combustion can also be observed in biochars. Verhoeff et al. [19] observed that in small scale experiments, self-ignition could occur.

According to the literature review, the recognition of self-heating in an industrial scale is crucial, but the literature for this topic is limited so far [20]. Past studies are limited to the intrinsic effect of self-heating on biochar’s elemental composition, and to determine the O/C and H/C ratio difference [[21], [22], [23]]. Studies for self-hating oxidation kinetics and its effect on energy yield, parameters causing the elevation of self-heating to self-ignition, and analysis of self-heating on large scale samples are still not available. For this reason, this work aims to identify the existence of the self-heating phenomenon that can be possibly observed in biochar discharging and storage stages after biomass is torrefied and exposed to air in a scaled-up setting. Potential safety hazard arises when biochar exhibits thermal runaway. This thermal runaway can lead to combustion and thereby causes a fire. This study specifically aims to compare the existence of self-heating under different scenarios and identify its effect on the higher heating value (HHV) which can be used to evaluate the quality of biochar as a fuel.

Section snippets

Raw materials and experimental set-up

The wood pellets used in the experiment were composed of various forest residues. The pellets were obtained from Jinding Green Energy in Taiwan. To provide a basis for preparing samples in the whole experiment, the pellets were dried in an oven at 105 °C for 24 h. To provide a basis for the experiment, then the pellets were trimmed to have average length and diameter of about 25 mm and 9 mm, respectively, to standardize the volume and surface area of the samples for experiments and analysis.

Experimental apparatus

The

Properties of raw wood pellets

The proximate analysis (dry basis) and HHVs of the raw wood pellets are shown in Table 3. The volatile matter (VM) content is 80.79 wt%, accounting for the largest share in the pellets. In contrast, the fixed carbon (FC) content is 17.81 wt%, and the ash content is 1.40 wt%. On account of the high VM and low FC contents of the raw wood pellets, the HHV of the fuel is 18.33 MJ/kg. This value is located in the normal HHV range of raw biomass (16–22 MJ/kg) [28].

Self-heating phenomena

The temperature profiles in wood

Conclusions

In this study, detecting the occurrence of self-heating has been conducted and the properties of the biochar have been analyzed to realize the effect of the self-heating on biochar formation. Light, mild, and severe torrefaction of wood pellets in N2 followed by three different post-torrefaction processes termed Scenarios 1–3 by exposing them to N2 and air with/without keeping the temperature is practiced. In the post-torrefaction processes using N2 (Scenario 1) and air (Scenario 2) without

CRediT authorship contribution statement

Emmanuel Arriola: Formal analysis, Data curation, Writing - original draft. Wei-Hsin Chen: Funding acquisition, Conceptualization, Project administration, Supervision, Resources, Validation, Writing - review & editing. Yi-Kai Chih: Formal analysis, Data curation. Mark Daniel De Luna: Data curation. Pau Loke Show: Writing - original draft.

Declaration of competing interest

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

Acknowledgments

The authors would like to acknowledge the financial support of the Ministry of Science and Technology, Taiwan, R.O.C., under the grant numbers MOST 108-3116-F-006-007-CC1, MOST 109-3116-F-006-016-CC1, and MOST 109-2221-E-006-040-MY3. The first author also acknowledges the ERDT Scholarship from the Department of Science and Technology, Philippines, for this research. The authors would like to extend gratitude to Jui-Chin Lee, XPS Technician in charge of NCKU Instrument Center.

References (67)

  • B. Evangelista et al.

    Reactor scale study of self-heating and self-ignition of torrefied wood in contact with oxygen

    Fuel

    (2018)
  • J.M. Ashman et al.

    Some characteristics of the self-heating of the large scale storage of biomass

    Fuel Process Technol

    (2018)
  • B.B. Beamish et al.

    Effect of moisture content on the R70 self-heating rate of Callide coal

    Int J Coal Geol

    (2005)
  • S. Krigstin et al.

    Comparative analysis of bark and woodchip biomass piles for enhancing predictability of self-heating

    Fuel

    (2019)
  • F. Restuccia et al.

    Quantifying self-heating ignition of biochar as a function of feedstock and the pyrolysis reactor temperature

    Fuel

    (2019)
  • J. Idris et al.

    Improved yield and higher heating value of biochar from oil palm biomass at low retention time under self-sustained carbonization

    J Clean Prod

    (2015)
  • X.J. Lee et al.

    Biochar potential evaluation of palm oil wastes through slow pyrolysis: thermochemical characterization and pyrolytic kinetic studies

    Bioresour Technol

    (2017)
  • S.-W. Park et al.

    Torrefaction and low-temperature carbonization of woody biomass: evaluation of fuel characteristics of the products

    Energy

    (2012)
  • H. Konno

    Chapter 8 - X-ray photoelectron spectroscopy

  • R.H.H. Ibrahim et al.

    Physicochemical characterisation of torrefied biomass

    J Anal Appl Pyrol

    (2013)
  • R. Galhano dos Santos et al.

    Estimation of HHV of lignocellulosic biomass towards hierarchical cluster analysis by Euclidean’s distance method

    Fuel

    (2018)
  • K. Weber et al.

    Properties of biochar

    Fuel

    (2018)
  • H. Wang et al.

    Analysis of the mechanism of the low-temperature oxidation of coal

    Combust Flame

    (2003)
  • Y. Shi et al.

    Modelling imbibition data for determining size distribution of organic and inorganic pores in unconventional rocks

    Int J Coal Geol

    (2019)
  • W.-H. Chen et al.

    Torrefaction and co-torrefaction characterization of hemicellulose, cellulose and lignin as well as torrefaction of some basic constituents in biomass

    Energy

    (2011)
  • S. Gul et al.

    Kinetic, volatile release modeling and optimization of torrefaction

    J Anal Appl Pyrol

    (2017)
  • S.-W. Du et al.

    Pretreatment of biomass by torrefaction and carbonization for coal blend used in pulverized coal injection

    Bioresour Technol

    (2014)
  • C. Zhang et al.

    Torrefaction performance and energy usage of biomass wastes and their correlations with torrefaction severity index

    Appl Energy

    (2018)
  • W.-H. Chen et al.

    Hygroscopic transformation of woody biomass torrefaction for carbon storage

    Appl Energy

    (2018)
  • S. Ceylan et al.

    Pyrolysis kinetics of hazelnut husk using thermogravimetric analysis

    Bioresour Technol

    (2014)
  • L. Wang et al.

    Effect of torrefaction on physiochemical characteristics and grindability of stem wood, stump and bark

    Appl Energy

    (2018)
  • S. Cardona et al.

    Torrefaction of eucalyptus-tree residues: a new method for energy and mass balances of the process with the best torrefaction conditions

    Sustainable Energy Technologies and Assessments

    (2019)
  • A. Alvarez et al.

    Non-oxidative torrefaction of biomass to enhance its fuel properties

    Energy

    (2018)
  • Cited by (19)

    • Complementary effects of torrefaction and pelletization for the production of fuel pellets from agricultural residues: A comparative study

      2022, Industrial Crops and Products
      Citation Excerpt :

      The mechanical strength of torrefied pellets ranges between 1 to 2 MPa. The similar observations were found by literature for torrefied biomass (Arriola et al., 2020; Manouchehrinejad and Mani, 2018; Zhang et al., 2020). However, durability of most of the pellet was found in the range of 97–100%, which is above the European standard (Alakangas and Eija, 2011) and Canadian wood pellet standard (NRC, 2021) for densified materials.

    • Synthesis of 5-hydroxymethylfurfural from glucose, fructose, cellulose and agricultural wastes over sulfur-doped peanut shell catalysts in ionic liquid

      2022, Chemosphere
      Citation Excerpt :

      Biomass utilization boasts of renewability and carbon neutrality, two characteristics missing from fossil fuel feedstocks (Pillejera et al., 2017; Salvilla et al., 2020). Common biomass applications include chemical synthesis, bioenergy, and environmental remediation (Genuino et al., 2018; Arriola et al., 2020). In 2004, the US Department of Energy proposed a list of chemical building blocks derived from carbohydrates such as 5-hydroxymethylfurfural (HMF) as an important step towards establishing a green chemical industry (Galkin and Ananikov, 2019).

    • Effect of in-situ torrefaction and densification on the properties of pellets from rice husk and rice straw

      2022, Chemosphere
      Citation Excerpt :

      Even though both the energy yield and solid yield decreased, the rate of decrease with temperature for energy yield was lower than the decrease for solid yield. These indicate that torrefaction positively influences the heating value and the sustainability of combustion (Arriola et al., 2020). Moreover, by increasing the temperature from 250 °C to 300 °C, the volatile content of RS and RH were both obviously decreased from 71.16% to 52.00%, from 66.37% to 51.42%, respectively, which was likely to harm the quality of subsequent densified pellets.

    View all citing articles on Scopus
    View full text