Next Article in Journal
A Novel Check-List Strategy to Evaluate the Potential of Operational Improvements in Wastewater Treatment Plants
Next Article in Special Issue
Solar-hybrid Thermochemical Gasification of Wood Particles and Solid Recovered Fuel in a Continuously-Fed Prototype Reactor
Previous Article in Journal
Electric Vehicle Aggregate Power Flow Prediction and Smart Charging System for Distributed Renewable Energy Self-Consumption Optimization
Previous Article in Special Issue
Fast Pyrolysis of Lignin Pretreated with Magnesium Formate and Magnesium Hydroxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Porous Structure Development and Ash on the Steam Gasification Reactivity of Biochar Residues from a Commercial Gasifier at Different Temperatures

Department of Chemical Engineering, KTH Royal Institute of Technology, SE-100 44 Stockholm, Sweden
*
Author to whom correspondence should be addressed.
Submission received: 18 August 2020 / Revised: 15 September 2020 / Accepted: 19 September 2020 / Published: 23 September 2020
(This article belongs to the Special Issue Thermochemical Conversion of Biomass and Waste)

Abstract

:
The present study aims at investigating the effects of porous structure development and ash content on the observed reactivity during steam gasification of biochar residues from a commercial gasifier. The experiments were conducted at a temperature range of 700 to 800 °C using biochar, derived from entrained flow gasification of biomass, under isothermal conditions using a thermogravimetric analyzer. The pore size distribution, surface area and morphology of char samples were determined by N2 physiosorption and scanning electron microscopy (SEM). The results showed that the gasification temperature does not affect the porous structure development considerably. The total surface area of char exhibits a threefold increase, while the total pore volume increase ranges between 2.0 and 5.3 times, at all temperatures. Both properties are directly proportional to the observed reactivity, especially at conversions up to 70%. Catalytic effects of the mineral matter of the char (mainly potassium) become predominant at the later stages of conversion (conversion greater than 70%).

Graphical Abstract

1. Introduction

With the increasing energy demand and the environmental impact by the use of fossil fuel, biomass waste feedstocks, such as forestry residue and agricultural waste, are becoming an attractive resource for renewable energy [1]. A feasible technique for primary conversion of biomass waste feedstocks into an intermediate gaseous feedstock (suitable for further upgrading to useful end user products, such as chemicals, heat, and power) is gasification [2]. The biomass conversion includes several consecutive steps, such as drying, pyrolysis, gasification, or partial combustion of the residual char, as well as homogeneous gas phase reactions. The gasification of char, produced in the pyrolysis step, consists of a set of heterogeneous reactions with gasifying agents, such as CO2, H2O, and O2, and is generally the rate-determining step [3]. An understanding of the mechanism and kinetics of char gasification is important to provide useful data for gasification reactor design.
Reports on gasification and combustion reactivity and kinetics of coal and coal char are extensive [4,5,6,7], as compared to a more limited number of studies on lignocellulosic char (or biochar) reactivity. There are many similarities between the two feedstocks, but the biochar is far more reactive (4–10 times) as a result of their physicochemical properties [8]. Information on the reactivity and kinetics of lignocellulosic char can be found in the comprehensive review by Di Blasi [9].
Char gasification is generally influenced by a number of different parameters such as temperature, particle size, char pore structure, and intrinsic char reactivity [9]. Furthermore, the particle temperature history of char generation influence its reactivity [10]. For example, Septien et al. [11] reported that increased pyrolysis heating rates (from 1 °C/s to 1000 °C/s) result in higher gasification rates and also to a more reactive solid char. The higher reaction rate (i.e., char reactivity) may be attributed to a change in morphology [12] or a change in textural properties due to the formation of a larger pore volume at high heating rates, mainly consisting of mesopores and macropores, compared to a lower pore volume of mainly micropores at a low heating rates [13]. Textural biomass-char properties are also significantly changed during the gasification process, expressed as an evident increase in porosity and surface area [14,15,16,17,18,19,20,21,22]. Most studies on pore structure development of biomass char use CO2 as the gasification agent [14,15,16,17,22] and only a few studies focus on char steam gasification, especially wood-derived chars, at different temperatures [18,19,23].
Another important factor influencing the reactivity is the mineral content and composition in the char. Alkali and alkaline earth metals (AAEMs) such as potassium, sodium, calcium, and magnesium, act as catalysts and accelerate the char gasification process, favoring heterogeneous char-steam, char-O2, and char-CO2 reactions [24,25,26,27,28,29,30,31,32,33].
The majority of studies of biochar reactivity generally use biochar samples produced in a pyrolysis step at lab-scale either directly in thermobalances, including thermogravimetric analysis (TGA) instruments [29,34] and fixed bed reactors [33], in drop tubes [31,35] or fluidized beds [28,36]. Detailed studies on reactivity of char, collected at commercial or demonstration scale gasification processes, so-called semi-char, are rarely found in the literature and corresponds to a few cases of coal semi-chars [35,37,38,39,40] and one case investigating different biomass semi-chars, collected from circulating fluidized bed (CFB) gasifiers [41]. The main results from the coal semi-char investigations are somewhat contradictory to each other, with studies observing a lower [37,40] and higher [35] gasification reactivity of char collected from fluidized bed gasifiers, compared to corresponding pyrolyzed coal chars. An example where coal semi-char from Texaco entrained flow slurry gasifiers is investigated is reported by Gu et al. [38]. They concluded that the chemical reactivity of two collected coal semi-chars was much higher compared to corresponding pyrolyzed chars. In case of biomass gasification, Meng et al. [41] investigated semi-char of wood pellets, willow, and dried distiller’s grains with solubles (DDGS) collected from the downcomer in a pilot scale steam-O2 blown CFB gasifier. Although not explicitly investigating the char reactivity, the calculated activation energy range of collected CFB semi-chars was lower, compared to the corresponding pyrolyzed chars, suggesting a higher reactivity of the semi-chars.
It is evident that the reactivity of a char differs depending on the conditions present in a real-scale gasification plant and also between chars produced in the laboratory and real chars. Most studies investigating steam biomass char gasification relate the observed reactivity to char structure development [17,19,20,22] or ash content [9,21,42,43,44,45,46] individually and not in synergy. Mechanistic studies on how both parameters contribute to the char reactivity of real biomass semi-char, during its transformation under steam gasification conditions, are important to enable the development of kinetic models. The present study, aims at evaluating a residual biomass semi-char, collected at a demonstration scale entrained flow gasification process (4 MWth biomass) for combined heat and power production and compare the char reactivity properties with char equivalents in the literature. The biomass char was gasified to different degrees of conversion with steam in a thermogravimetric analyzer (TGA), where each sample of partially gasified char was characterized in terms of textural and morphological characteristics, as well as of ash contents on the surface. In order to evaluate the combined effects of the structural development and ash components on char reactivity, during steam gasification, ash-leached char samples were prepared and compared with the original char.

2. Method

2.1. Raw Material and Sample Preparation

The material used in the present work was unreacted biomass char, collected at a demonstration scale entrained flow gasifier, using pine wood with a maximum particle diameter of 0.1 mm (see Supplementary Materials). The unreacted char was collected from the scrubber sedimentation tank, placed downstream of the gasifier. Biomass is gasified at temperatures ranging between 900 and 1150 °C using air as gasification agent at an equivalence ratio of 0.28–0.32.
Prior to use, the char residue was dried in at 105 °C overnight. In order to remove any residual volatile matter, the samples were heated at 950 °C in N2 atmosphere for 3h. The sample was ground to a particle size of 45–120 μm. The resultant char is designated as original char (OC). The results of ultimate and proximate analyses, and mineral content, determined according to the American Society for Testing and Materials (ASTM) methods and ISO 29541:2010, respectively, of both the original biomass fed to the gasifier and the OC are reported in Table 1.
In order to evaluate the effects of ash on char reactivity during steam gasification, ash-leached char samples were also prepared. Deionized water was used as a leaching agent to avoid damaging the char structure and hence affecting the char reactivity. Leaching was performed at 40 °C with a liquid to solid ratio (v/w) of 80 [47] under continuous stirring for 2 and 48 h in order to achieve a different degree of demineralization. After leaching, the samples were washed and dried at 105 °C overnight. The leached samples were named as 2LC and 48LC referring to 2 and 48 h of treatment, respectively. Mineral matter of original and leached chars were analyzed by ICP-SFMS.

2.2. Steam Char Gasification TGA Experiments

Char steam gasification experiments were carried out in isothermal conditions, using a NETZCSH ST490 F3 thermogravimetric analyzer (TGA). In each gasification experiment, approximately 30 ± 2 mg of char powder was loaded in a 3.4 mL ceramic crucible. The sample was initially heated at a rate of 10 °C/min in N2 atmosphere (300 mL/min) to the desired temperature (700, 750, and 800 °C) and maintained for 5 min for temperature stabilization. The gasification took place in a 7.6 mol-% steam atmosphere. The char conversion (X) and reactivity (dX/dt) were calculated according to the following,
X = m 0 m t m 0 m a s h     [ ]
d X d t = 1 m 0 m a s h d m t d t ,     [ min 1 ]
where m0 represents the initial mass of the char at the onset of gasification, mt is the instantaneous mass of the char at time t, and mash is the residual mass of ash.
In order to study the development of the porous structure of char, the gasification reaction was stopped at different conversions aiming at capturing all the reactivity turning points that were observed during complete conversion (at least three to four points).

2.3. Textural Analysis (N2 Physisorption)

The specific surface area and pore size distribution were determined by N2 physisorption (Micromeritics ASAP 2000) based on the Brunauer–Emmett–Teller (BET) method [48,49]. Each sample was degassed at 250 °C overnight prior to analysis. The pore size distribution was determined using Barrett–Joyner–Halenda (BJH) method, while the micropore volume was determined using the t-plot method.

2.4. SEM Analysis

Samples’ surface morphology was investigated by means of scanning electron microscopy (SEM) using a Zeiss GeminiSEM 450 at 15kV. SEM was coupled with energy-dispersive X-ray spectroscopy (EDX). Determination of surface elements concentration was performed using a Si (Li) detector and the Oxford INCA Energy software. The analyses were done in low vacuum mode with carbon tape under the sample. Average mineral content on char sample surface was determined from SEM images with 1000× magnification.

3. Results

All the tests conducted in the present work are presented in Table 2. For OC, textural and morphological characteristics were determined at different conversions. To ensure the repeatability of the runs, the mass loss curves at 700–800 °C at different conversions were compared (see Supplementary Materials (Figures S-2 to S-4)). It was found that the experimental error does not exceed 2%.
The reactivity of the TGA-produced char (from original biomass) and the gasifier-derived char (OC) were determined at 900 °C and are shown in Figure 1. As seen, the temperature history of the particle has an effect on the reactivity of the derived char. The gasifier-derived char (OC) shows almost half of the reactivity of the TGA produced char and shows an increasing reactivity at the later stages of the conversion.

3.1. Original Char

The reactivity of char at 700, 750, and 800 °C is shown in Figure 2, Figure 3 and Figure 4, respectively. As shown, there are two distinct regions of conversion where reactivity behavior is different; namely, conversion from 0 to 70% and from 70% to complete conversion. In the first region the reactivity is somewhat stable with decreasing trend while at conversions higher than 70%, the reactivity increases significantly.
Figure 5 illustrates the pore size distribution of the original char at different conversions and temperatures (700, 750, and 800 °C). As shown, at conversions higher than 70%, the mesopores (2–50 nm) fraction start to decrease at the expense of macropores showing a progressive pore growth. The pore volume in the macropore range (>50 nm) is positively correlated with the conversion.
The textural properties of the char at different temperatures and conversions for the original char can be seen in Figure 6. The multipoint BET method was used to calculate the specific surface area, with a relative error between 0.37% and 0.78%. As shown in Figure 6a,b the total pore volume (Vtotal) and the total specific surface area (Stotal) increases 1.3–3.1 and 2.0–5.3 times, respectively, with conversion until a value of approximately 70% at all temperatures, indicating a formation of new pores, as a result of steam-carbon reaction. Similar results are also reported in other studies [17,22]. Stotal, as well as Vtotal, start to decrease after this point. The fraction of micropores increases with conversion monotonically and constitute the majority of the surface area of the char. At 70%, a steep increase in Vmicro and Smicro formation is observed. Figure 6b,d displays the increase of Stotal and Smicro with conversion with a higher increase in the rate of Stotal compared to Smicro until ~70% conversion. Therefore, the mesopores develop at a faster rate than micropores generated until this point. This result is inconsistent with Fu et al. [18], who observed an increase of the rate of Stotal as Smicro kept constant, when the conversion is higher than 10%. The discrepancy can be attributed to the different controlling regime as will be discussed later.
As shown in Figure 6e,f, the mesopores volume fraction increases from 44% to 77% until 70% conversion, while the mesopore area fraction increases from 24% to 45%. At conversions higher than ~70%, the volume fraction of the mesopores decreases from 77% to ~65%. Similarly, the mesopores fraction in the total surface area decreases from 45% to 20%.
At different gasification temperatures, the Smicro/Stotal ratio curve is almost identical to the changes in the reactivity (Figure 2, Figure 3 and Figure 4). In the conversion range from 0 to 70%, the reactivity decreases and so does the Smicro/Stotal ratio. After 70% conversion, the reactivity and Smicro/Stotal ratio starts to increase significantly. This behavior may be related to the generation of new micropores, resulting in an extended surface area for heterogeneous steam-carbon reaction, which leads to increased reactivity.
SEM images (with 1000× magnification) of original char and gasified chars at different carbon conversions, from gasification at 700, 750 and 800 °C in 30 mol-% steam atmosphere, are shown in Figure 7, Figure 8, Figure 9 and Figure 10 (500× and 100× magnifications are presented in the Supplementary Materials (Figures S-5 to S-11). As shown, the surface of the original char is quite porous (as also confirmed by the BET measurements); sparse minerals particles, mainly Ca, is distributed thinly on the char surface (see Supplementary Material for additional information). During the gasification reaction, the carbon is continuously consumed, and pores of gasified chars are expanding and increasing.
As shown in Figure 8, honeycomb-like structures are observed on the char surface. At 25% of char conversion, the surface is more porous than the original char and more small pores are scattered on the char surface. During the conversion of the char, the original pores are enlarged, and new pores are formed (red circle), while the char particle size is gradually decreased.
The surface of the char obtained at 750 °C, shown in Figure 9, displays macroscopic similarities to the one obtained at 700 °C with mineral particles being spread unevenly on the surface. In Figure 10, the surface morphology at 800 °C is shown and is similar to the ones at 700 and 750 °C. Mineral particles (yellow rectangle) become larger, plausibly due to an increased mobility and agglomeration at higher gasification temperatures. N.B. Klinghoffer et al. [36] observed that the agglomeration is clearly visible at 1000 °C.
The agglomerates are formed at higher conversion (green triangle) due to the increased mineral content during char consumption, which is analogous to the results obtained by Yonghui et al. [22] while investigating the morphology of chars exposed in steam atmosphere.
The surface concentration of the major AAEM species (Ca, K, Mg, and Na) at different stages of char conversion is presented in Figure 11. As the char conversion increases from 0% to ~70%, the surface concentration of Ca and Mg increases at 700 °C. However, at 750 and 800 °C, the Ca concentration decreases with temperature plausibly because of formation of larger Ca particles due to agglomeration which resulted in lower exposed Ca. The decrease should not be ascribed to volatilization, since both Ca and Mg are most likely kept as oxides in the sample, especially when the temperature is lower than 1000 °C [50]. Potassium surface concentration is positively correlated with the increase of char conversion with the exception of the low temperature gasification at 700 °C [51], which is at the lower limit of alkali salt vaporization, suggesting K to be highly mobile [50]. At all temperatures, the surface concentration of Na shows no significant change.

3.2. Water Leached Chars

In order to identify the effects of the mineral content on char conversion, gasification experiments with demineralized char were conducted. N2-physisorption (Figure 12) of the water leached char indicate that water leaching does not affect the pore structure of the char samples. The corresponding specific surface areas are 330, 340, and 410 m2/g for the OC, 2LC, and 48LC, respectively. The above is macroscopically verified by SEM microscopy as shown in Figure 7 and Figure 13.
ICP-SFMS analysis indicated that K, Mg, and Ca are the major mineral elements in the OC, while Na, Fe, Al, and P are present in lower concentrations (Figure 14). As seen, the water leaching results in a significant reduction of the amount of K, while Ca and Mg are only slightly decreased. As similar to K, Na is also affected greatly by water leaching. Approximately 77% of Na and 86% of K were washed out after the 48 h treatment. Given that the contents of Fe, Al, and Mn are barely affected by the water treatment and the very low Na content, it is reasonable to ascribe any observed differences in reactivity to K. The importance of K in char reactivity has also been identified by Wang et al. [52].
The reactivities of the water-leached charrs at 700, 750, and 800 °C are shown in Figure 15. At all temperatures, OC shows a constant reactivity up to conversions of approximately 70%, where a significant increase is observed. The mildly leached sample (2LC) shows a declining reactivity up to conversion of approximately 80% where a significant increase in reactivity is observed. The reactivity of 48LC exhibits a monotonic steep decline in reactivity with conversion indicating that the increase in reactivity at the later stages of gasification is mainly attributed to the mineral content.

4. Discussion

As shown in Figure 1, the gasifier-derived char shows a lower reactivity compared to the TGA-generated char. Albeit heating rate is known to be of importance for production of char of high reactivity both in biomass and coal gasification [9,13,53,54], the results indicate that the char generated at high heating rate (entrained flow gasifier) is not as reactive as the char generated at slow heating rate in the TGA. Another parameter which is of importance is the temperature history of the char particle. The char from entrained flow gasifier has been exposed to temperatures in excess of 1000 °C and is partly preoxidized. The degree of preoxidation and temperature are being reported to greatly affect the reactivities of coal derived chars [55,56], and therefore the observed lower reactivity can be attributed to the high preoxidation temperature the char has been subjected to in the gasifier.
It is evident (Figure 5) that during char conversion, there is a simultaneous generation of micropores and development of mesopores (and to some extent macropores at higher conversions), as a result of micropore enlargement, irrespective of the temperature. The simultaneous and uniform development are indicative of an interconnected porous network [22]. The generation of micropores suggests reaction on the active sites of char on the external surface [57] or opening of the so-called closed pores of the biochar structure [20,58] both being supported by the observed increase in char surface area and pore volume (Figure 6).
Moreover, the continuous and uniform porous network development indicates that the total particle volume is readily accessible inferring that intraparticle diffusion is not limiting the conversion, and thus the observed conversions are under kinetic control.
The reactivity behaviors observed (Figure 15) at conversions up to ~70% are a result of the total char surface area available for reaction with steam. The observed decline in the reactivity for all samples is primarily due to a decrease in the total surface area (and thus active reaction sites) in the reaction system (thermo balance), supporting that the system is under kinetic control regime [59,60]. Increased access of steam to catalytic active sites, which are gradually exposed as conversion progresses (Figure 8, Figure 9 and Figure 10) and their surface density increase with carbon consumption, can ebhance the reactivity at later stages. Diffusion limitations have been reported to be of importance for char particle sizes in the vicinity of 300 µm at temperatures in excess of 800 °C in steam atmosphere [29,61].
The dependence on the char surface area can be elucidated by determining the specific area normalized reactivity at very low conversion with the assumption that the surface area of the initial char has not changed considerably. As listed in Table 3 the obtained normalized reactivities are comparable and the differences observed can be ascribed to the small variation in the initial surface area and porosity of the samples.
As perceived, the initial surface normalized (or intrinsic) reactivity of the char is comparable irrespective of the ash content, indicating that the available char area is more important. The generation of new micropores and enlargement of existing pores (as shown in Figure 5 and Figure 6) during conversion increases the surface area and accessibility of steam, however the consumption of carbon results in, as expected, reduced reactivity. The normalized reactivity is reduced by approximately 63%, 51%, and 25% at ~25% conversion for 700, 750, and 800 °C, respectively. Limitations due to diffusion through ash layer, similar to the shrinking core model, cannot explain the obtained results given that no extensive ash layer is formed (Figure 8, Figure 9 and Figure 10).
The reactivity of OC albeit declining, is the least affected (for conversions up to ~70%) compared to 2LC and 48LC (Figure 15). The observed decline in reactivity is negatively correlated with the mineral content of the char, indicating that with increasing particle conversion, the contribution of the catalytic effects of the mineral content become more important. Peranders et al. [32] observed that there was a steady increase of reactivity with the increase of the alkali content (K) at char conversions up to 70%. Bouraoui et al. [29] reported that the reaction rate is strongly dependent on the K content at conversions higher than 70%, during CO2 gasification of char, while Suzuki, T. et al. [33] observed proportional increase in gasification rates to the K content of cedar wood. The observed behavior can be related to the surface exposed minerals. As can be seen in Figure 11, the K on the char surface increases as the reaction progresses to completion. The different behavior of 48LC is therefore directly correlated to the K content.
At conversions greater than 70%, the char reactivity is increased to various degrees for all samples, indicating the aforementioned dependence on the mineral content (Figure 14 and Figure 15). At all temperatures, the turning point of reactivity curves is observed at similar conversion (66–70%), supporting the char conversion until this point took place under kinetic controlled regime (negligible intraparticle diffusion limitations) [60]. This suggests that there are distinct regions where the superimposed phenomena of micropore structure development and alkali catalyzed reactions are prevailing.
The observed lag in the reactivity increase for 2LC and 48LC, as shown in Figure 15, can be ascribed to the reduced K content (Figure 14) and its associated catalytic activity. The onset of reactivity increase is observed at conversion of 85% and 92% for 2LC and 48LC, respectively.
For OC, the catalytic effect would lead to a continuous increase in reactivity until the conversion reaches 95% (Figure 15). For 48LC, the reaction rate decreases with conversion up to 92%. At this stage, the micropore surface area increases indicating enhanced reaction activity. Figure 16 schematically illustrates the progression of char conversion and the role of K.

5. Conclusions

In the present study, the evolution of char structural features and the effect of ash, during steam gasification of char residues collected from a commercial entrained flow gasifier, were investigated. Thermogravimetric analysis was used to assess the combined effects of pore structure and ash on char reactivity toward steam at a temperature range between 700 and 800 °C.
From the early conversion stages, the char particles are characterized by a broad pore size distribution that includes both micro- and mesopores at a ~1:1 volume ratio. During conversion, new micropores are generated and also existing micropores and mesopores are enlarged indicating that the whole particle volume is accessible to steam, which in turn implies that conversion takes place in the kinetic control regime. The generation of micropores and the subsequent enlargement result in substantial increase in the specific total surface area, as well as the total pore volume of char (1.3–3.1 and 2.0–5.3 times for the surface area and pore volume respectively) up to a certain conversion point, where they pivot and decrease at all temperatures. The temperature has a minor effect on char pore structure development.
At conversions below 70%, the conversion is proportional to the surface area of the char irrespective of the mineral content of the char. Declining reactivity with increasing conversion supports that conversion was taking place in kinetic regime and is proportional to char active sites.
The results indicate that the char generated at high heating rate in an entrained flow gasifier is not as reactive as the char generated at slow heating rate in the TGA. The observed lower reactivity can possibly be attributed to high preoxidation temperature that the char has been subjected in the entrained flow gasifier.
When the conversion reaches 70%, catalytic phenomena mainly due to potassium presence become dominant. From this it is evident that potassium is an important parameter to predict steam gasification reactivity for high conversion, as in the present study is determined to be above 70%, The point of conversion where the reactivity increases is clearly dependent on the K content, as illustrated by the observed differences for the original and water leached chars.
Conclusively, the study showed that the superimposed phenomena of surface area and pore development as well as catalytic reactions due to alkali presence have distinct regions where they prevail with the low conversions governed by noncatalytic reactions and micropores generation, while at high conversions the importance of catalytic reactions is greater.

Supplementary Materials

Author Contributions

S.D. was responsible for conceptualization, data curation, validation, formal analysis, and original draft preparation. E.K. and K.E. were responsible for the conceptualization and supervision of the project. All authors participated in the writing, review, and editing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Swedish gasification Centre (SFC), Swedish Energy Agency (contract no 34721-2).

Acknowledgments

This work was carried out within the Swedish Centre for Biomass Gasification (SFC). The authors are grateful to the Swedish Energy Agency for financial support. The authors appreciate the industrial partner (Meva Energy) for supplying the sample.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Acharya, B.; Dutta, A.; Basu, P. An investigation into steam gasification of biomass for hydrogen enriched gas production in presence of CaO. Int. J. Hydrogen Energy 2010, 35, 1582–1589. [Google Scholar] [CrossRef]
  2. Rauch, R.; Hrbek, J.; Hofbauer, H. Biomass gasification for synthesis gas production and applications of the syngas. WIREs Energy Environ. 2014, 3, 343–362. [Google Scholar] [CrossRef]
  3. Ahmed, I.I.; Gupta, A.K. Kinetics of woodchips char gasification with steam and carbon dioxide. Appl. Energy 2011, 88, 1613–1619. [Google Scholar] [CrossRef]
  4. Laurendeau, N.M. Heterogeneous kinetics of coal char gasification and combustion. Prog. Energy Combust. Sci. 1978, 4, 221–270. [Google Scholar] [CrossRef]
  5. Smith, I.W. The combustion rates of coal chars: A review. Symp. (Int.) Combust. 1982, 19, 1045–1065. [Google Scholar] [CrossRef]
  6. Bews, I.M.; Hayhurst, A.N.; Richardson, S.M.; Taylor, S.G. The order, Arrhenius parameters, and mechanism of the reaction between gaseous oxygen and solid carbon. Combust. Flame 2001, 124, 231–245. [Google Scholar] [CrossRef]
  7. Hurt, R.H.; Calo, J.M. Semi-global intrinsic kinetics for char combustion modeling. Combust. Flame 2001, 125, 1138–1149. [Google Scholar] [CrossRef]
  8. Encinar, J.M.; González, J.F.; Rodríguez, J.J.; Ramiro, M.J. Catalysed and uncatalysed steam gasification of eucalyptus char: Influence of variables and kinetic study. Fuel 2001, 80, 2025–2036. [Google Scholar] [CrossRef]
  9. Di Blasi, C. Combustion and gasification rates of lignocellulosic chars. Prog. Energy Combust. Sci. 2009, 35, 121–140. [Google Scholar] [CrossRef]
  10. Guizani, C.; Jeguirim, M.; Valin, S.; Limousy, L.; Salvador, S. Biomass Chars: The Effects of Pyrolysis Conditions on their Morphology, Structure, Chemical Properties and Reactivity. Energies 2017, 10, 796. [Google Scholar] [CrossRef] [Green Version]
  11. Septien, S.; Escudero Sanz, F.J.; Salvador, S.; Valin, S. The effect of pyrolysis heating rate on the steam gasification reactivity of char from woodchips. Energy 2018, 142, 68–78. [Google Scholar] [CrossRef] [Green Version]
  12. Avila, C.; Pang, C.H.; Wu, T.; Lester, E. Morphology and reactivity characteristics of char biomass particles. Bioresour. Technol. 2011, 102, 5237–5243. [Google Scholar] [CrossRef]
  13. Mermoud, F.; Salvador, S.; Van de Steene, L.; Golfier, F. Influence of the pyrolysis heating rate on the steam gasification rate of large wood char particles. Fuel 2006, 85, 1473–1482. [Google Scholar] [CrossRef] [Green Version]
  14. Senneca, O. Kinetics of pyrolysis, combustion and gasification of three biomass fuels. Fuel Process. Technol. 2007, 88, 87–97. [Google Scholar] [CrossRef]
  15. Gil, M.V.; Riaza, J.; Álvarez, L.; Pevida, C.; Rubiera, F. Biomass devolatilization at high temperature under N2 and CO2: Char morphology and reactivity. Energy 2015, 91, 655–662. [Google Scholar] [CrossRef] [Green Version]
  16. Bai, Y.; Wang, Y.; Zhu, S.; Li, F.; Xie, K. Structural features and gasification reactivity of coal chars formed in Ar and CO2 atmospheres at elevated pressures. Energy 2014, 74, 464–470. [Google Scholar] [CrossRef]
  17. Guizani, C.; Jeguirim, M.; Gadiou, R.; Escudero Sanz, F.J.; Salvador, S. Biomass char gasification by H2O, CO2 and their mixture: Evolution of chemical, textural and structural properties of the chars. Energy 2016, 112, 133–145. [Google Scholar] [CrossRef] [Green Version]
  18. Fu, P.; Hu, S.; Xiang, J.; Yi, W.; Bai, X.; Sun, L.; Su, S. Evolution of char structure during steam gasification of the chars produced from rapid pyrolysis of rice husk. Bioresour. Technol. 2012, 114, 691–697. [Google Scholar] [CrossRef]
  19. Keown, D.M.; Hayashi, J.-I.; Li, C.-Z. Drastic changes in biomass char structure and reactivity upon contact with steam. Fuel 2008, 87, 1127–1132. [Google Scholar] [CrossRef]
  20. Wu, H.; Yip, K.; Tian, F.; Xie, Z.; Li, C.-Z. Evolution of Char Structure during the Steam Gasification of Biochars Produced from the Pyrolysis of Various Mallee Biomass Components. Ind. Eng. Chem. Res. 2009, 48, 10431–10438. [Google Scholar] [CrossRef]
  21. Kajita, M.; Kimura, T.; Norinaga, K.; Li, C.-Z.; Hayashi, J.-I. Catalytic and Noncatalytic Mechanisms in Steam Gasification of Char from the Pyrolysis of Biomass. Energy Fuels 2010, 24, 108–116. [Google Scholar] [CrossRef]
  22. Bai, Y.; Lv, P.; Yang, X.; Gao, M.; Zhu, S.; Yan, L.; Li, F. Gasification of coal char in H2O/CO2 atmospheres: Evolution of surface morphology and pore structure. Fuel 2018, 218, 236–246. [Google Scholar] [CrossRef]
  23. Alvarez, J.; Lopez, G.; Amutio, M.; Bilbao, J.; Olazar, M. Evolution of biomass char features and their role in the reactivity during steam gasification in a conical spouted bed reactor. Energy Convers. Manag. 2019, 181, 214–222. [Google Scholar] [CrossRef]
  24. Davidsson, K.O.; Korsgren, J.G.; Pettersson, J.B.C.; Jäglid, U. The effects of fuel washing techniques on alkali release from biomass. Fuel 2002, 81, 137–142. [Google Scholar] [CrossRef]
  25. Di Blasi, C.; Branca, C.; D’Errico, G. Degradation characteristics of straw and washed straw. Thermochim. Acta 2000, 364, 133–142. [Google Scholar] [CrossRef]
  26. Jenkins, B.M.; Bakker, R.R.; Wei, J.B. On the properties of washed straw. Biomass Bioenergy 1996, 10, 177–200. [Google Scholar] [CrossRef]
  27. Zolin, A.; Jensen, A.; Jensen, P.A.; Frandsen, F.; Dam-Johansen, K. The Influence of Inorganic Materials on the Thermal Deactivation of Fuel Chars. Energy Fuels 2001, 15, 1110–1122. [Google Scholar] [CrossRef]
  28. Aho, A.; DeMartini, N.; Pranovich, A.; Krogell, J.; Kumar, N.; Eränen, K.; Holmbom, B.; Salmi, T.; Hupa, M.; Murzin, D.Y. Pyrolysis of pine and gasification of pine chars-influence of organically bound metals. Bioresour. Technol. 2013, 128, 22–29. [Google Scholar] [CrossRef]
  29. Bouraoui, Z.; Dupont, C.; Jeguirim, M.; Limousy, L.; Gadiou, R. CO2 gasification of woody biomass chars: The influence of K and Si on char reactivity. C. R. Chim. 2016, 19, 457–465. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, Y.; Yin, X.; Wu, C.; Wang, C.; Xie, J.; Zhou, Z.; Ma, L.; Li, H. Effects of metal catalysts on CO2 gasification reactivity of biomass char. Biotechnol. Adv. 2009, 27, 568–572. [Google Scholar] [CrossRef]
  31. Karlström, O.; Dirbeba, M.J.; Costa, M.; Brink, A.; Hupa, M. Influence of K/C Ratio on Gasification Rate of Biomass Chars. Energy Fuels 2018, 32, 10695–10700. [Google Scholar] [CrossRef]
  32. Perander, M.; DeMartini, N.; Brink, A.; Kramb, J.; Karlström, O.; Hemming, J.; Moilanen, A.; Konttinen, J.; Hupa, M. Catalytic effect of Ca and K on CO2 gasification of spruce wood char. Fuel 2015, 150, 464–472. [Google Scholar] [CrossRef]
  33. Suzuki, T.; Nakajima, H.; Ikenaga, N.-O.; Oda, H.; Miyake, T. Effect of mineral matters in biomass on the gasification rate of their chars. Biomass Conv. Bioref. 2011, 1, 17–28. [Google Scholar] [CrossRef] [Green Version]
  34. Kirtania, K.; Axelsson, J.; Matsakas, L.; Christakopoulos, P.; Umeki, K.; Furusjö, E. Kinetic study of catalytic gasification of wood char impregnated with different alkali salts. Energy 2017, 118, 1055–1065. [Google Scholar] [CrossRef]
  35. Kelebopile, L.; Sun, R.; Liao, J. Fly ash and coal char reactivity from Thermo-gravimetric (TGA) experiments. Fuel Process. Technol. 2011, 92, 1178–1186. [Google Scholar] [CrossRef]
  36. Klinghoffer, N.B.; Castaldi, M.J.; Nzihou, A. Influence of char composition and inorganics on catalytic activity of char from biomass gasification. Fuel 2015, 157, 37–47. [Google Scholar] [CrossRef] [Green Version]
  37. Jing, X.; Wang, Z.; Yu, Z.; Zhang, Q.; Li, C.; Fang, Y. Experimental and Kinetic Investigations of CO2 Gasification of Fine Chars Separated from a Pilot-Scale Fluidized-Bed Gasifier. Energy Fuels 2013, 27, 2422–2430. [Google Scholar] [CrossRef]
  38. Gu, J.; Wu, S.; Wu, Y.; Li, Y.; Gao, J. Differences in Gasification Behaviors and Related Properties between Entrained Gasifier Fly Ash and Coal Char. Energy Fuels 2008, 22, 4029–4033. [Google Scholar] [CrossRef]
  39. Zhang, H.; Zhu, Z.; Dong, Q.; Yu, K.; Lu, Q. Structural Properties and Gasification Reactivity of Shenmu Fly Ash Obtained from a 5t/d Circulating Fluidized Bed Gasifier. Procedia Eng. 2015, 102, 1104–1111. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, Y.; Zhang, H.; Zhu, Z. Regasification properties of industrial CFB-gasified semi-char. J. Therm. Anal. Calorim. 2018, 131, 3035–3046. [Google Scholar] [CrossRef]
  41. Meng, X.; Benito, P.; de Jong, W.; Basile, F.; Verkooijen, A.H.M.; Fornasari, G.; Vaccari, A. Steam–O2 Blown Circulating Fluidized-Bed (CFB) Biomass Gasification: Characterization of Different Residual Chars and Comparison of Their Gasification Behavior to Thermogravimetric (TG)-Derived Pyrolysis Chars. Energy Fuels 2012, 26, 722–739. [Google Scholar] [CrossRef]
  42. Marquez-Montesinos, F.; Cordero, T.; Rodrıguez-Mirasol, J.; Rodrıguez, J.J. CO2 and steam gasification of a grapefruit skin char. Fuel 2002, 81, 423–429. [Google Scholar] [CrossRef]
  43. Dupont, C.; Jacob, S.; Marrakchy, K.O.; Hognon, C.; Grateau, M.; Labalette, F.; Da Silva Perez, D. How inorganic elements of biomass influence char steam gasification kinetics. Energy 2016, 109, 430–435. [Google Scholar] [CrossRef]
  44. Dupont, C.; Nocquet, T.; Da Costa, J.A.; Verne-Tournon, C. Kinetic modelling of steam gasification of various woody biomass chars: Influence of inorganic elements. Bioresour. Technol. 2011, 102, 9743–9748. [Google Scholar] [CrossRef] [PubMed]
  45. Hognon, C.; Dupont, C.; Grateau, M.; Delrue, F. Comparison of steam gasification reactivity of algal and lignocellulosic biomass: Influence of inorganic elements. Bioresour. Technol. 2014, 164, 347–353. [Google Scholar] [CrossRef] [PubMed]
  46. Yip, K.; Tian, F.; Hayashi, J.-I.; Wu, H. Effect of Alkali and Alkaline Earth Metallic Species on Biochar Reactivity and Syngas Compositions during Steam Gasification. Energy Fuels 2010, 24, 173–181. [Google Scholar] [CrossRef]
  47. Ohki, A.; Nakajima, T.; Yamashita, H.; Iwashita, A.; Takanashi, H. Leaching of various metals from coal into aqueous solutions containing an acid or a chelating agent. Fuel Process. Technol. 2004, 85, 1089–1102. [Google Scholar] [CrossRef]
  48. Haul, R. S. J. Gregg, K. S. W. Sing: Adsorption, Surface Area and Porosity. 2. Auflage, Academic Press, London 1982. 303 Seiten, Preis: 49.50. Berichte der Bunsengesellschaft für Physikalische Chemie 1982, 86, 957. [Google Scholar] [CrossRef]
  49. Ioannidou, O.; Zabaniotou, A. Agricultural residues as precursors for activated carbon production—A review. Renew. Sustain. Energy Rev. 2007, 11, 1966–2005. [Google Scholar] [CrossRef]
  50. Defoort, F.; Dupont, C.; Durruty, J.; Guillaudeau, J.; Bedel, L.; Ravel, S.; Campargue, M.; Labalette, F.; Da Silva Perez, D. Thermodynamic Study of the Alkali Release Behavior during Steam Gasification of Several Biomasses. Energy Fuels 2015, 29, 7242–7253. [Google Scholar] [CrossRef]
  51. Moud, P.H.; Andersson, K.J.; Lanza, R.; Pettersson, J.B.C.; Engvall, K. Effect of gas phase alkali species on tar reforming catalyst performance: Initial characterization and method development. Fuel 2015, 154, 95–106. [Google Scholar] [CrossRef]
  52. Wang, M.; Shen, Y.; Guo, P.; Kong, J.; Wu, Y.; Chang, L.; Wang, J.; Xie, W. A comparative study on the intrinsic reactivity and structural evolution during gasification of chars from biomass and different rank coals. J. Anal. Appl. Pyrolysis 2020, 149, 104859. [Google Scholar] [CrossRef]
  53. Ashu, J.T.; Nsakala, N.Y.; Mahajan, O.P.; Walker, P.L. Enhancement of char reactivity by rapid heating of precursor coal. Fuel 1978, 57, 250–251. [Google Scholar] [CrossRef]
  54. Cetin, E.; Moghtaderi, B.; Gupta, R.; Wall, T.F. Influence of pyrolysis conditions on the structure and gasification reactivity of biomass chars. Fuel 2004, 83, 2139–2150. [Google Scholar] [CrossRef]
  55. Gomi, K.; Hishinuma, Y. Effect of pre-oxidation on reactivity of chars in steam. Fuel 1982, 61, 77–80. [Google Scholar] [CrossRef]
  56. Tremel, A.; Haselsteiner, T.; Nakonz, M.; Spliethoff, H. Coal and char properties in high temperature entrained flow gasification. Energy 2012, 45, 176–182. [Google Scholar] [CrossRef]
  57. Fatehi, H.; Bai, X.-S. Structural evolution of biomass char and its effect on the gasification rate. Appl. Energy 2017, 185, 998–1006. [Google Scholar] [CrossRef]
  58. Struis, R.P.W.J.; Von Scala, C.; Stucki, S.; Prins, R. Gasification reactivity of charcoal with CO2. Part I: Conversion and structural phenomena. Chem. Eng. Sci. 2002, 57, 3581–3592. [Google Scholar] [CrossRef]
  59. Hodge, E.M.; Roberts, D.G.; Harris, D.J.; Stubington, J.F. The Significance of Char Morphology to the Analysis of High-Temperature Char−CO2 Reaction Rates. Energy Fuels 2010, 24, 100–107. [Google Scholar] [CrossRef]
  60. Naredi, P.; Pisupati, S.V. Interpretation of Char Reactivity Profiles Obtained Using a Thermogravimetric Analyzer. Energy Fuels 2008, 22, 317–320. [Google Scholar] [CrossRef]
  61. Guizani, C.; Escudero Sanz, F.J.; Salvador, S. Influence of temperature and particle size on the single and mixed atmosphere gasification of biomass char with H2O and CO2. Fuel Process. Technol. 2015, 134, 175–188. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Reactivity of thermogravimetric analysis (TGA)-derived char as compared to the gasifier-derived char.
Figure 1. Reactivity of thermogravimetric analysis (TGA)-derived char as compared to the gasifier-derived char.
Energies 13 05004 g001
Figure 2. Micropores development and its reactivity at 700 °C.
Figure 2. Micropores development and its reactivity at 700 °C.
Energies 13 05004 g002
Figure 3. Micropores development and its reactivity at 750 °C.
Figure 3. Micropores development and its reactivity at 750 °C.
Energies 13 05004 g003
Figure 4. Micropores development and its reactivity at 800 °C.
Figure 4. Micropores development and its reactivity at 800 °C.
Energies 13 05004 g004
Figure 5. Pore size distribution of char at different conversions and 700, 750, and 800 °C according to Barrett–Joyner–Halenda (BJH) method.
Figure 5. Pore size distribution of char at different conversions and 700, 750, and 800 °C according to Barrett–Joyner–Halenda (BJH) method.
Energies 13 05004 g005
Figure 6. Textural properties of the OC at different temperatures with different conversions for original char. Left: pore volume versus conversion at different temperatures. Right: Specific surface area versus conversion at different temperatures. (The relative error is minor and is not shown in the figure.) (a) the total pore volume versus conversion; (b) the total specific surface area versus conversion; (c) the micropore volume versus conversion; (d) the micopore surface area versus conversion; (e) the mesopores volume fraction versus conversion; (f) the mesopore area fraction versus conversion.
Figure 6. Textural properties of the OC at different temperatures with different conversions for original char. Left: pore volume versus conversion at different temperatures. Right: Specific surface area versus conversion at different temperatures. (The relative error is minor and is not shown in the figure.) (a) the total pore volume versus conversion; (b) the total specific surface area versus conversion; (c) the micropore volume versus conversion; (d) the micopore surface area versus conversion; (e) the mesopores volume fraction versus conversion; (f) the mesopore area fraction versus conversion.
Energies 13 05004 g006
Figure 7. SEM images (1000× magnification) of the original char.
Figure 7. SEM images (1000× magnification) of the original char.
Energies 13 05004 g007
Figure 8. SEM images (1000× magnification) of char with different carbon conversions prepared from char gasification in H2O at 700 °C.
Figure 8. SEM images (1000× magnification) of char with different carbon conversions prepared from char gasification in H2O at 700 °C.
Energies 13 05004 g008
Figure 9. SEM images (1000× magnification) of char at different carbon conversions prepared from char gasification in H2O at 750 °C.
Figure 9. SEM images (1000× magnification) of char at different carbon conversions prepared from char gasification in H2O at 750 °C.
Energies 13 05004 g009
Figure 10. SEM images (1000× magnification) of char with different carbon conversions prepared from char gasification in H2O at 800 °C.
Figure 10. SEM images (1000× magnification) of char with different carbon conversions prepared from char gasification in H2O at 800 °C.
Energies 13 05004 g010
Figure 11. Surface concentrations of the alkali and alkaline earth metal (AAEM) species as a function of char conversion at different temperatures.
Figure 11. Surface concentrations of the alkali and alkaline earth metal (AAEM) species as a function of char conversion at different temperatures.
Energies 13 05004 g011
Figure 12. Pore size distribution (BJH) of original and the leached char samples.
Figure 12. Pore size distribution (BJH) of original and the leached char samples.
Energies 13 05004 g012
Figure 13. SEM images of 2LC and 48LC.
Figure 13. SEM images of 2LC and 48LC.
Energies 13 05004 g013
Figure 14. Effects of water leaching on the concentrations of main inorganic elements in char sample, dry basis.
Figure 14. Effects of water leaching on the concentrations of main inorganic elements in char sample, dry basis.
Energies 13 05004 g014
Figure 15. Gasification rate versus conversion for OC, 2LC, and 48LC at different temperatures.
Figure 15. Gasification rate versus conversion for OC, 2LC, and 48LC at different temperatures.
Energies 13 05004 g015
Figure 16. Simplified schematic representation of char conversion in steam atmosphere. (a) Progression of char conversion. (b) The role of K at higher char conversion.
Figure 16. Simplified schematic representation of char conversion in steam atmosphere. (a) Progression of char conversion. (b) The role of K at higher char conversion.
Energies 13 05004 g016
Table 1. Ultimate and proximate analyses of biomass and original char.
Table 1. Ultimate and proximate analyses of biomass and original char.
Proximate Analysis (wt%)
BiomassOriginal Char
Moisture7.20.6
Volatile Matter (db) b92.84.53
Ash (db) a0.326.9
Fixed carbon (db) b6.8888.57
Ultimate Analysis (wt% db)
C50.990.6
H6.20.2
N<0.10.43
O b42.61.87
Mineral Matter (mg/kg db)
Na111755
K3405690
Mg1142660
Ca56211,200
Fe45.5720
Al79336
Mn69.91540
P37.8677
Si43.11450
db-dry basis. a Determined at 550 °C, b Determined by difference.
Table 2. Summary of experiments and analyses.
Table 2. Summary of experiments and analyses.
700 °C750 °C800 °C
Conversions (%)0276472100027537391100025457085100
SampleOC
Tests
TGA
BET
SEM/EDS
Samples2LC and 48LC
Tests
TGA
BET
SEM/EDS
Table 3. Specific area normalized reactivity at 3% conversion.
Table 3. Specific area normalized reactivity at 3% conversion.
(dX/dt)/Stotal (g/m2-min)
SamplesOC2LC48LC
T (°C)
7005.26 × 10−67.01 × 10−65.49 × 10−6
7501.45 × 10−51.90 × 10−51.49 × 10−5
8003.30 × 10−53.81 × 10−53.18 × 10−5

Share and Cite

MDPI and ACS Style

Ding, S.; Kantarelis, E.; Engvall, K. Effects of Porous Structure Development and Ash on the Steam Gasification Reactivity of Biochar Residues from a Commercial Gasifier at Different Temperatures. Energies 2020, 13, 5004. https://0-doi-org.brum.beds.ac.uk/10.3390/en13195004

AMA Style

Ding S, Kantarelis E, Engvall K. Effects of Porous Structure Development and Ash on the Steam Gasification Reactivity of Biochar Residues from a Commercial Gasifier at Different Temperatures. Energies. 2020; 13(19):5004. https://0-doi-org.brum.beds.ac.uk/10.3390/en13195004

Chicago/Turabian Style

Ding, Saiman, Efthymios Kantarelis, and Klas Engvall. 2020. "Effects of Porous Structure Development and Ash on the Steam Gasification Reactivity of Biochar Residues from a Commercial Gasifier at Different Temperatures" Energies 13, no. 19: 5004. https://0-doi-org.brum.beds.ac.uk/10.3390/en13195004

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