Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Accepted Manuscript

Short Communication

Effects of pyrolysis temperature and heating time on biochar obtained from the
pyrolysis of straw and lignosulfonate

Jie Zhang, Jia Liu, Rongle Liu

PII: S0960-8524(14)01613-7
DOI: http://dx.doi.org/10.1016/j.biortech.2014.11.011
Reference: BITE 14217

To appear in: Bioresource Technology

Received Date: 25 September 2014


Revised Date: 3 November 2014
Accepted Date: 4 November 2014

Please cite this article as: Zhang, J., Liu, J., Liu, R., Effects of pyrolysis temperature and heating time on biochar
obtained from the pyrolysis of straw and lignosulfonate, Bioresource Technology (2014), doi: http://dx.doi.org/
10.1016/j.biortech.2014.11.011

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Effects of pyrolysis temperature and heating time on biochar obtained from the

pyrolysis of straw and lignosulfonate


Jie Zhang1, Jia Liu2 , Rongle Liu1*

1 The Graduate School Chinese Academy of Agricultural Sciences, Beijing, China

2 Soil and Fertilizer & Resources and Environment Institute, Jiangxi Academy of Agricultural Sciences,

Nanchang, China

*
Corresponding author. E-mail: rlliu@caas.net.cn. Tel: +86 10 82106606.
Postal address: The Graduate School Chinese Academy of Agricultural Sciences, No. 12 Zhongguancun
South Street, Haidian District, Beijing, China. 100081.

Contributed equally to this work.
Effects of pyrolysis temperature and heating time on biochar obtained from the

pyrolysis of straw and lignosulfonate

Abstract

In this study, the effects of pyrolysis temperature and heating time on the yield and

physicochemical and morphological properties of biochar obtained from straw and

lignosulfonate were investigated. As pyrolysis temperature increased, pH, ash content,

carbon stability, and total content of carbon increased while biochar yield, volatile

matter, total content of hydrogen, oxygen, nitrogen and sulfur decreased. The data from

scanning electron microscope image and nuclear magnetic resonance spectra indicated

an increase in porosity and aromaticity of biochar produced at a high temperature. The

results showed that feedstock types could also influence characteristics of the biochar

with absence of significant effect on properties of biochar for heating time.

Key words: Straw; Lignosulfonate; Biochar; Pyrolysis temperature; Heating time;

1. Introduction

Biochar is a predominantly stable, recalcitrant organic carbon compound, which can

be obtained when biomass is heated to temperatures usually between 300 ºC and 1000

ºC, under low (preferably zero) oxygen concentrations (Verheijen et al., 2010). The

conversion of biomass into biochar has been receiving greater attention from

government regulation agencies and the general public (Inyang et al., 2010). According

to the International Biochar Initiative Organization, it is estimated that by the year 2050

about 80% of all crop and forestry residues may be converted to biochar and energy

2
(Kolodynska et al., 2012).

The properties and functions of those biochar are highly depending on the feedstock

materials and production condition (Bird et al., 2011; Yao et al., 2011). Although

almost all carbonous biomass can be converted into biochar through thermal pyrolysis, a

life cycle assessment of pyrolysis biochar systems suggested that it is more

environmentally and financially viable to make biochar from waste biomass (Roberts et

al., 2010). In this case, agricultural residues and other industrial waste have been

proposed as good feedstock materials to make biochar (Cao and Harris, 2010).

The annual worldwide production of wheat straw as agricultural waste was estimated

to be approximately 540 million tons in 2007 (Reddy and Yang, 2007). The straw might

be left on the field, burned, fed animals or used as industrial raw materials. As

lignosulfonate is the main component of paper mill waste, huge amount of

lignosulfonate was generated and the disposal of waste (liquid, solid and suspended

matter) generated during the paper manufacturing process contributed to a very high

impact on the environment, but less of them were utilized. Moreover large proportion of

waste have been disposed of by burning and discharging, resulting in not only a waste

of resource but also a serious environmental problem. Conversion of straw and

lignosulfonate into biochar through pyrolysis has further advantages of energy and

environment. Therefore, this study examined the physicochemical properties of biochar

produced under various pyrolysis temperatures and heating time using wheat straw and

lignosulfonate as the feedstock. Such understanding is essential for development of

3
biochar application and technology for the production of agricultural and industrial

waste with improved value.

2. Materials and methods

2.1 Feedstock preparation

Two species of feedstock were selected for the production and evaluation of biochar.

Wheat straw and lignosulfonate was collected from a paper mill producing paper

products from neutral sulphite chemical straw pulp. The feedstock was dried at 80 ºC

for 24 h to remove moisture.

2.2 Biochar production

The feedstock was placed in a cylindrical stainless steel reactor of 14.5 cm height and

7.5 cm internal diameter with a lid and pyrolyzed in a muffle furnace (2116, Thermo

Scientific, America). Three different peak temperature and time, i.e. 200, 400, 600 ºC

and 1 h, 2 h, 4 h were adapted to carbonize each feedstock. These subjected

temperatures were selected based on the results of earlier reports (Lee et al., 2013).

Then the reactor was cooled to room temperature inside the furnace. Biochar thus

produced was immediately weighed and at least triplicate aliquots were produced at

each temperature and time, indicating that weight loss in same manner was reproducible.

The biochar from all aliquots at each temperature and time was combined, mixed and

lightly crushed by hand to pass 1 mm for analyses.

2.3 Biochar characteristic

A range of physicochemical properties of the straw (S), lignosulfonate (L), straw

4
biochar (SB) and lignosulfonate biochar (LB) were determined. The pH of the biochar

was measured by adding biochar to deionized water in a mass ratio of 1:10. The

solution was measured by a pH meter (PP-20, Sartorius, Germany). Volatile matter and

ash content were determined by a modified ASTM method (D-1762-84) involving

measurement of weight loss following combustion of about 10 g of biochar in a ceramic

crucible at 900 ºC for 6 min and 750 ºC for 2 h, respectively. Fixed carbon content was

determined by difference. Elemental carbon, hydrogen, oxygen, nitrogen and sulfur

concentrations were determined by Elemental analyzer (Vari pyro cube, elementar,

Germany). Scanning electron microscope (SEM) imaging analysis of the biochar was

conducted using a Hitachi S-570 scanning microscope (Japan). The solid state 13C

nuclear magnetic resonance (NMR) spectra was measured on a Bruker Avance III 400

NMR spectrometer (Germany) conducting at a spinning speed of 5 kHz and a contact

time of 1 ms, with a 1H 90º pulse length of 4 µs and a recycle delay of 0.8 s. The

chemical shift regions 0-45, 45-95, 95-165 and 160-200 ppm were referred to alkyl

carbon, oxygen-alkyl carbon, aryl carbon and carboxylic carbon, respectively.

3. Results and discussion

3.1 Proximate analysis of biochar

The yield of biochar decreased as pyrolysis temperature increased (Table 1) because

the thermal volatiles were further cracked into low molecular weight liquid and gas

rather than biochar as the temperature increased (Thangalazhy-Gopakumar et al., 2010).

But Novak et al. (2009) attributed the decrease in biochar yield with increased

5
temperature to dehydration of hydroxyl groups and thermal degradation of cellulose and

lignin structures. At 200 ºC, feedstock lost little mass, even after 4 h, which shows that

limited pyrolysis occurs at this temperature. A large decrease in yield occurred at 400

ºC, agreeing with the observation of Shinogi and Kanri (2003). The LB was higher than

SB at the same pyrolysis temperature and heating time. High yield of LB may be

ascribed to high ash and low volatile matter in L compared to S. The feedstock (25 ºC)

was acidic and the pH of the 200 ºC treatment showed a pH slight reduction. The

cellulose and hemicelluloses were decomposed around 180-250 ºC, producing organic

acids and phenolic substances that lowered pH of the products (Abe et al. 1998).

Heating time did not have a significant effect on pH, but the pH of biochar increased

with increasing temperature. These increases in pH were mainly due to the organic

functional groups such as -COOH and -OH decreased with increasing pyrolysis

temperature (Yuan et al.,2011). The pH increase of biochar above 400 ºC may be

attributed to the carbonates formation (such as CaCO3and MgCO3) and inorganic alkalis

(such as K, Na, Ca and Mg) of L (Yuan et al., 2011). Considering the pH of biochar,

they may also be used as soil amendments to reduce soil acidity. The fixed carbon

content of biochar increased by 1.1-1.9 times (SB) and 1.1-1.5 times (LB) compared

with that in the feedstock, indicating the carbon stability of biochar enhancing. The

fixed carbon content of SB was higher than that in LB. From a carbon sequestration

perspective, SB was more suited to preserve carbon than LB at the same condition.

Volatile matter content decreased with charring temperature and duration indicating

6
progressive loss of more volatile component with charring. The ash content is a measure

of the non-volatile matter and non-combustible component of the biochar (Angın, 2013).

The ash content remarkably increased at 400 ºC and 600 ºC. Contrastingly, the increase

in ash with temperature was not observed for biochar at 200 ºC. The increase in ash

content should result from progressive concentration of minerals and destructive

volatilization of lignocelluloses matters as temperature increased (Tsaia et al.,2012).

However, the ash content of LB was much higher (up to 59.69%) than that in SB (up to

34.31%), probably due to the presence of impurities, such as fine pulp solids, clays,

small fibers and so on in the paper mill waste.

3.2 Elemental analysis of biochar

Pyrolysis temperature showed more significant effect on biochar’s elemental

compositions than heating time (Table 1). The data showed that carbon content of

biochar increased with temperature, but the loss of hydrogen, oxygen and nitrogen was

recorded. The carbon content of LB was lower than that of SB, and they corresponded

to their feedstock instead of heating time, which showed that feedstock type was the

determinant. Losses in hydrogen and oxygen at high pyrolysis temperature were

attributed to the cleavage of weak bonds within the feedstock structure. For LB, more

than 30% of hydrogen was removed from the feedstock when pyrolyzing at 200 ºC.

Furthermore, pyrolyzing at 600 ºC resulted in total hydrogen losses exceeding 85%. The

nitrogen content of LB is relatively higher than that in SB, this may be attributable to

ammonium sulfide pulping process, while very small amounts of sulfur was observed in

7
SB. Though it has been reported by many authors that pyrolysis lead to reduce nitrogen

content of biochar (Song and Guo, 2012). Interestingly, compared with its feedstock,

the SB had lower content of nitrogen, indicating that low temperature could conserve

nitrogen in straw. This was due to incorporating of nitrogen into complex structures that

are resistant to heating and not easily volatilized. In consistent with our results, Calvelo

Pereira et al. (2011) reported that nitrogen enrichment relative to the original feedstock

was recorded upon pyrolysis of woody material. The sulfur of biochar decreased as

pyrolysis temperature increased because the pyrolysis decomposition of organic sulfur

during pyrolysis at 400 ºC.

3.3 SEM analysis of biochar

The SEM images revealed that with increasing temperature, the particles of biochar

became more and retained less evidence of original cell structures. The SEM analysis

showed increase in degradation of feedstock, but at low temperatures, there was

incomplete decomposition. As the temperature increased, feedstock decomposition

increased and the underlying layers became more exposed resulting in improved

porosity of the biochar, but heating time had no noticeable effect biochar morphology.

SEM images showed the presence of aligned honeycomb-like groups of pores on SB,

most likely the carbonaceous skeleton from the biological capillary structure of

feedstock. The images of SB were in agreement with previous studies using similar

feedstock, for instance, Brodowski et al. (2005) observed primary tissue in maize straw

after heat to 350 ºC but destruction of the original structure at higher temperatures.

8
Comparison of the SEM images of LB showed surface structure changes occurring

during temperature increasing. The biochar produced at 200 ºCwas more inerratic and

flat, while biochar produced at 400 ºC and 600 ºC had several cracks and holes on its

surface, suggesting the surface area for this biochar increased, which provided more

adsorption sites, space for nutrients and water retention (Lehmann et al., 2003). There

were obvious differences of structure between image of SB and LB.

3.4 Solid-state13C NMR analysis of biochar

The solid-state13C NMR spectra analysis was used to investigate their structures. The

results demonstrated that there was a clear relationship between pyrolysis temperature

and the chemical characteristics of aromaticity and aromatic condensation in biochar.

The spectrum of feedstock and biochar at 200 ºC were similar but quite different from

those of biochar at 400 ºC and 600 ºC indicating incomplete decomposition of biochar

at low temperature. This is due to the progressive dehydration, decarbonylation and

decarboxylation reactions (Nishimiya, 1998). The signal alkyl carbon was formed in

feedstock, SB200 and LB400, but was absent at the higher temperature of 600 ºC. The

destroy of aliphatic carbon can partially interpret the rapid decrease of hydrogen/carbon

from 400 ºC to 600 ºC. Several signals for carbohydrates were clearly visible at 45-110

ppm in the NMR spectrum of feedstock and decreased with increasing temperature.

Based on the percentage of total carbon, the aromaticity content (95-165 ppm) of

biochar were low in feedstock and biochar at 200 ºC, but this content of SB, LB sharply

increased at 400 ºC and accounted for 75.79%, 84.25%, respectively. The aliphatic

9
carbon and aromatic carbon content of S was 76.42%(alkyl carbon-10.29%,

oxygen-alkyl carbon-66.13%)and 20.02%, respectively. In contrast, heating to 600 ºC

resulted in higher aromatic carbon content (81.85%) and lower amount of oxygen-alkyl

carbon (12.80%), and no signal in alkyl carbon region. The spectrum of the biochar

produced at 400 ºC and 600 ºC for 2 h were dominated by one signal at 130 ppm which

is characteristic of aromatic carbon. This result has significance for the stability of

biochar in soil.

4. Conclusion

The results of this study indicated that heating time during biochar production did not

have a significant effect on properties of biochar but pyrolysis temperature was found to

greatly influence both physicochemical properties and stability. As pyrolysis

temperature increased, biochar yield, volatile matter decreased. The pH and carbon

content increased with increasing temperature, as the pyrolysis progressed, hydrogen,

oxygen, nitrogen and sulfur were removed. The degree of carbonization for biochar was

accelerated with increasingly pyrolysis temperature from 200 ºC to 400 ºC. The type of

feedstock also affected yield, elemental composition and chemical structure of the

biochar.

References

[1] Verheijen, F., Jeffery, S., Bastos, A.C., van der Velde, M., Diafas, I., 2010. JRC

Scientific and Technical Reports-a critical scientific review of effects on soil

properties, processes and functions. Office for the Official Publications of the

10
European Communities, Luxembourg.

[2] Inyang, M., I., Gao, B., Pullammanappallil, P., Ding, W.C., Zimmerman, A.R.,

2010. Biochar from anaerobically digested sugarcane bagasse. Bioresour. Technol.

101, 8868-8872.

[3] Kolodynska, D., Wnetrzak, R., Leahy, J.J., Hayes, M.H.B., Kwapinski, W., Hubicki,

Z., 2012. Kinetic and adsorptive characterization of biochar in metal ions removal.

Chem. Eng. J. 197, 295-305

[4] Bird, M.I., Wurster, C.M., de Paula Silva, P.H., Bass, A.M., de Nys, R., 2011.

Algal biochar-production and properties. Bioresour. Technol. 102, 1886-1891.

[5] Yao, Y., Gao, B., Inyang, M., Zinnerman, A.R., Cao, X.D., Pullammanappallil, P.,

Yang, L.Y, 2011. Biochar derived from anaerobically digested sugar beet tailings:

Characterization and phosphate removal potential. Bioresour. Technol. 102,

6273-6278.

[6] Roberts, K.G., Gloy, B.A., Joseph, S., Scott, N.R., Lehmann, J., 2010. Life cycle

assessment of biochar systems: estimating the energetic, economic, and climate

change potential. Environ. Sci. Technol. 44, 827-833.

[7] Cao, X.D., Harris, W., 2010. Properties of dairy-manure-derived biochar pertinent

to its potential use in remediation. Bioresource. Technol. 101, 5222-5228.

[8] Reddy, N., Yang, Y., 2007. Preparation and characterization of long natural

cellulose fibers from wheat straw. J. Agric. Food Chem. 55, 8570-8575.

[9] Lee, Y., Pu-Reun-Byul Eum, P.R.B., Ryu, C., Park, Y.K., Jung, J.H., Hyun, S.,

11
2013. Characteristics of biochar produced from slow pyrolysis of Geodae-Uksae 1.

Bioresour. Technol. 130, 345-350.

[10] Thangalazhy-Gopakumar, S., Adhikari, S., Ravindran, H., Gupta, R.B., Fasina, O.,

Tu, M., Fernando, S.D., 2010. Physiochemical properties of bio-oil produced at

various temperatures from pine wood using an auger reactor. Bioresour. Technol.

101, 8389-8395.

[11] Novak, J.M., Busscher, W.J., Laird, D.L., Ahmedna M., Watts, D.W., Niandou,

M.A.S., 2009. Impact of biochar amendment on fertility of a southeastern coastal

plain soil. Soil Sci. 174, 105-112

[12] Shinogi, Y., Kanri, Y., 2003. Pyrolysis of plant, animal and human waste: physical

and chemical characterization of the pyrolysis products. Bioresour. Technol. 90,

241–247.

[13] Abe, F., 1998. The thermochemical study of forest biomass. Forest. Prod. Chem. 45,

1–95.

[14] Yuan, J.H., Xu, R.K., Zhang, H., 2011. The forms of alkalis in the biochar

produced from crop residues at different temperatures. Bioresour. Technol. 102,

3488-3497.

[15] Angın, D., 2013. Effect of pyrolysis temperature and heating rate on biochar

obtained from pyrolysis of safflower seed press cake. Bioresour. Technol. 128,

593–597

[16] Tsaia, W.T., Liu, S.C., Chen, H.R., Chang, Y.M., Yi-Lin Tsai, Y.L., 2012. Textural

12
and chemical properties of swine-manure-derived biochar pertinent to its potential

use as a soil amendment. Chemosphere 89, 198-203.

[17] Song, W., Guo, M., 2012. Quality variations of poultry litter biochar generated at

different pyrolysis temperatures. J. Anal. Appl. Pyrolysis 94, 138–145.

[18] Calvelo Pereira, R., Kaal, J., Camps Arbestain, M., Pardo Lorenzo, R., Aitkenhead,

W., Hedley, M., Macías, F., Hindmarsh, J., Maciá-Agulló, J.A., 2011. Contribution

to characterisation of biochar to estimate the labile fraction of carbon. Org.

Geochem. 42, 1331–1342.

[19] Brodowski, S., Amelung, W., Haumaier, L., Abetz, C., Zech, W., 2005.

Morphological and chemical properties of black carbon in physical soil fractions as

revealed by scanning electron microscopy and energy-dispersive X-ray

spectroscopy. Geoderma 128, 116–129.

[20] Nishimiya, K., 1998. Analysis of chemical structure of wood charcoal by X-ray

photoelectron spectroscopy. J. Wood Sci. 44, 56.

13
Table 1 Physicochemical properties of biochar
Volatile Fixed
Temperature Duration Yield Ash C H O N S
Feedstock pH matter carbon
(ºC) (h) (%) (%) (%) (%) (%) (%) (%)
(%) (%)

S - - - 6.57 10.81 59.33 29.86 35.17 6.56 49.45 0.68 1.47

200 1 84.95 5.34 11.90 54.46 33.63 45.57 5.77 38.45 1.03 1.23
200 2 81.23 5.91 12.29 53.24 34.47 46.33 5.62 36.67 1.01 0.93
200 4 78.24 6.11 12.78 52.49 34.72 46.52 5.50 36.08 1.00 0.66
400 1 37.30 10.82 25.74 33.85 40.41 57.07 3.33 16.50 0.88 0.80
400 2 35.91 10.86 27.54 29.80 42.66 57.59 3.22 15.69 0.79 0.80
400 4 36.65 10.78 28.40 28.84 42.76 57.92 3.04 15.36 0.82 0.74
600 1 32.48 10.93 32.33 17.80 49.87 59.17 1.54 11.03 0.75 0.74
600 2 32.48 10.96 33.57 12.10 54.34 59.09 1.44 10.67 0.78 0.80
600 4 30.89 10.99 34.31 10.31 55.37 60.80 1.33 10.62 0.77 0.77
L - - - 4.89 32.98 41.84 25.18 25.74 5.00 46.38 4.85 12.83
200 1 81.99 4.18 35.86 35.08 29.06 30.33 3.81 34.80 4.76 11.83
200 2 78.93 4.56 36.15 35.06 28.79 31.07 3.61 32.55 4.76 11.11
200 4 74.97 4.37 36.93 34.43 28.64 31.94 3.09 31.83 4.71 10.44
400 1 58.51 9.75 45.61 20.83 33.56 33.58 1.82 27.91 3.65 9.45
400 2 57.80 9.65 47.61 18.26 34.13 33.67 1.75 27.45 3.65 11.38
400 4 57.24 9.35 49.83 17.89 32.28 33.60 1.58 26.91 3.66 11.69
600 1 52.99 10.68 56.13 7.05 36.83 34.52 1.09 25.49 2.79 6.69
600 2 48.12 12.50 58.12 6.32 35.56 34.60 1.10 27.83 2.70 8.59
600 4 43.85 12.95 59.69 5.26 35.05 36.81 1.26 29.75 2.95 6.79

14
Table 2 Quantitative contribution of carbon species
0-45(ppm) 45-95(ppm) 95-165(ppm) 165-200(ppm)
Temperature
Feedstock Alkyl C O-alkyl C Aryl C Carboxylic C
(ºC)
% of total carbon

S - 10.29 66.13 20.02 3.56


200 10.89 53.83 29.37 5.91
400 11.68 6.47 75.79 6.06
600 - 12.80 81.85 5.35
L - 17.58 39.95 34.14 8.33
200 18.44 19.93 52.61 9.03
400 5.86 5.71 84.25 4.18
600 - 6.86 87.17 5.87

15
Research highlights 1: The properties of biochar obtained from lignosulfonate were

first time explored.

Research highlights 2: We studied biochar produced at different pyrolysis temperature

and heating time.

Research highlights 3: The temperature affected biochar properties more

significantly than heating time.

16

You might also like