Download as pdf or txt
Download as pdf or txt
You are on page 1of 114

Biochar Production Process Optimisation and Product

Characterisation

by

Richa Neupane

Thesis

Submitted in fulfillment of the requirements for the degree of

Master of Applied Science

Central Queensland University

School of Health, Medical and Applied Sciences

Associate Professor Surya P Bhattarai and Professor Kerry B Walsh

September 2022
RHD Thesis Declaration

CANDIDATE’S STATEMENT

By submitting this thesis for formal examination at CQUniversity Australia, I declare that it
meets all requirements as outlined in the Research Higher Degree Theses Policy and
Procedure.

STATEMENT OF AUTHORSHIP AND ORIGINALITY

By submitting this thesis for formal examination at CQUniversity Australia, I declare that all
of the research and discussion presented in this thesis is original work performed by the
author. No content of this thesis has been submitted or considered either in whole or in part,
at any tertiary institute or university for a degree or any other category of award. I also
declare that any material presented in this thesis performed by another person or institute
has been referenced and listed in the reference section.

COPYRIGHT STATEMENT

By submitting this thesis for formal examination at CQUniversity Australia, I acknowledge


that thesis may be freely copied and distributed for private use and study; however, no part
of this thesis or the information contained therein may be included in or referred to in any
publication without prior written permission of the author and/or any reference fully
acknowledged.

PREVIOUS SUBMISSION STATEMENT

This paper has not been submitted for an award by another research degree candidate,
either at CQUniversity or elsewhere.

ACKNOWLEDGEMENT OF FINANCIAL SUPPORT


I gratefully acknowledge the funding received from “Mara Seeds Pty Ltd” through the “Mara
Seeds Industry Scholarship” which has supported this research. The funding sourced from
Australian Government, Business, through the “Innovation Connection Grant” for partial
scholarship support is greatly acknowledged. A full tuition fee waiver from “Central
Queensland University” through a “International Excellence Award” is also acknowledged.

ACKNOWLEDGEMENT OF OTHER SUPPORT


This research was undertaken with in-kind support of the facilities and feedstocks for the
biochar production provided by “Mara Seeds Pty Ltd”.

2
ABSTRACT
Biochar is a carbonaceous solid material produced through heating organic material
in a reduced oxygen environment. The impact of feedstock and pyrolysis conditions
(temperature and time) on biochar properties was assessed using a pilot plant
fluidized bed pyrolyser. Three feedstocks; macadamia nutshell, eucalyptus woodchip
and hemp stem where chosen for their range of woodiness, and as locally available
waste materials. Apatite was used as the fluid bed, given potential for use as a
phosphorus (P) source after roasting. Three different temperatures (350, 450 and
600 ℃) and three residence times (5, 10 and 20 min) were employed. Residence
time had almost no influence on biochar physical and chemical properties, indicating
that the pyrolysis process was completed within five minutes, a result ascribed to the
efficiency of heat transfer in the fluid bed system. For all feedstocks, increasing
pyrolysis temperature was associated with increasing biochar pH, EC, liming
potential, ash content, carbon content, calorific value, water holding capacity, surface
area and aromaticity along with decreasing bulk density, volatile matter, cation
exchange capacity (CEC), zeta potential and functional groups. For example, pH of
nutshell significantly increased from 5.7 to 9.12 and bulk density decreased from
0.36 to 0.22 g/cm3 when temperature increased from 350 to 600 ℃. Feedstock
significantly influence biochar properties, with the two woody materials (nutshell and
woodchip) separating from hemp stem in properties and in a principal component
analysis of infrared spectra. For example, total P, K and Ca was higher in hemp stem
(1.2, 2.04 and 3.12 % w/w) as compared to macadamia nutshell (0.071, 0.084 and
0.146 %) and eucalyptus woodchip (0.81, 0.008, 1.93 %). Infrared spectroscopy is
recommended as a potential tool for quality control of the production process. Thus,
assessed biochar properties were used to match feedstock and pyrolysis conditions
to potential uses, such as improvement of soil water holding capacity, soil CEC,
animal feed and wastewater treatment. After 5 minutes at 600 ℃ under reducing
conditions, apatite P extractability (in 2 % formic acid) decreased from 1.21 to 0.25
% of dry weight, for apatite that contained 12.7 % of total P. However, further work is
recommended to consider the effect of repeated oxidation and reduction cycles on P
extractability, as will occur in a full-scale pyrolysis plant.

3
DEDICATION
To my parents and grandparents, Rishi Ram Neupane, Bebi Dahal Neupane, Krishna
Prasad Dahal, Balkumari Dahal, my supportive husband, Bikash Gautam and my sister and
brother, I would like to dedicate this work. For their tremendous love, continuing support, and
belief that I am capable enough to accomplish everything I work hard for and be at the top in
whatever I do. This is just a gesture of appreciation for your never-ending guidance, thank
you for everything.

“My daughter can reach any destination she intends to achieve, and she is much stronger
than she thinks”.

- My Father

4
ACKNOWLEDGEMENTS
My foremost gratitude to my principal supervisor, Associate Professor Dr. Surya Prasad
Bhattarai for his endless guidance and supervision during the entire master’s degree. I would
like to thank my co-advisor, Prof. Kerry Walsh for his continuous support, valuable
comments and guidance for making this thesis into reality. I would like to acknowledge Prof.
Stephen Joseph, one of the pioneers in biochar research, I feel privileged to work under his
supervision.
I would like to acknowledge CQUniversity for the International excellence award and Mara
Seeds Pty Ltd for the Industry stipend scholarship. This scholarship has assessed me to
deliver full focus and attention to my research works. My heartfelt gratitude to industry
partners, Stuart Larsson, Katina Larsson, and John Winter, for providing all the resources,
guidance, and supervision to commence my research and field works.
I would like to express my deepest gratitude to Miss Sobia Ikram for her continuous support,
guidance, and always being there for me as an elder sister in Rockhampton. I would like to
acknowledge Dr. Pramod Shrestha, Dr. Yadav Bajhgai, Dr. Larelle Fabbro, Alison Craig,
Laura Roden, Dr. Kalpana Pudasaini, Bhima Pandit Bhattarai, Rati Bhattarai, Dipendra
Aryal, Dr. Thakur Bhattarai, who provided help from their side guiding throughout the
research period.
I am very much thankful to the technical team, Ian Learmont, Andrew Bryant, Antony Kodel,
Margaret Stockill, Giselle Weegenaar, Matt Jones, Tania Collins, Judith Couper and Joel
Johnson for guiding throughout the laboratory analysis process.
I express my deep sense of gratitude to the laboratory team of Environmental Analysis
Laboratory, Australian Laboratory Services Pty. Ltd and University of New South Wales for
the analysis of biochar samples.
My profound gratitude to Susan Kinnear, Dean of Graduate research, for her continuous
support throughout the research higher degree.
All my efforts would be meaningless if I miss disclosing my heartfelt gratitude to my parents,
Rishi Ram Neupane and Bebi Dahal Neupane for their love and blessings and my husband,
Bikash Gautam for his love and guidance, and for keeping me motivated in all highs and
lows. I express my deepest regards to my sister Ritu Neupane, brothers Thomas and Nayan
Neupane for providing the emotional support and patience and believing in me for making
this possible.
Last but not the least, I would like to thank all the helping hands for their continuous
coordination directly and indirectly for the completion of my research.

5
Table of Contents

ABSTRACT.............................................................................................................................. 3
ACKNOWLEDGEMENTS ..................................................................................................... 5
Table of Contents .................................................................................................................. 6
List of Tables ........................................................................................................................ 10
List of Figures ...................................................................................................................... 11
CHAPTER 1. INTRODUCTION .......................................................................................... 16
Biochar ............................................................................................................................... 16
Project background ........................................................................................................ 17
Thesis aims and objectives .......................................................................................... 18
Thesis structure .............................................................................................................. 19
Research publication ..................................................................................................... 19
CHAPTER 2. LITERATURE REVIEW .............................................................................. 20
Context ............................................................................................................................... 20
History of biochar ........................................................................................................... 20
Biochar uses .................................................................................................................... 21
Biochar production......................................................................................................... 22
Pyrolysis method ........................................................................................................ 22
Pyrolysis reactor types ............................................................................................. 24
Temperature ................................................................................................................. 26
Residence time ............................................................................................................ 26
Quenching of biochar ................................................................................................ 27
Feedstocks........................................................................................................................ 27
Commercial pyrolysis systems in Australia ............................................................ 28
Biochar characterisation and functionalities .......................................................... 29
Attributes....................................................................................................................... 29
C sequestration and fuel value: carbon content ................................................ 30
Soil improvement: CEC, zeta potential, mineral content, density, porosity,
pH, and liming potential ............................................................................................ 30
Absorption of toxins: Porosity and surface area ............................................... 32
Polycyclic aromatic hydrocarbons and Polychlorinated biphenyls ............. 33
Functional groups....................................................................................................... 34
Apatite ................................................................................................................................ 34

6
Bed properties ............................................................................................................. 34
Fertiliser potential ....................................................................................................... 35
Conclusion ........................................................................................................................ 36
CHAPTER 3. MATERIALS and METHODS .................................................................... 37
Introduction ...................................................................................................................... 37
Biochar production......................................................................................................... 37
Feedstock material ..................................................................................................... 37
Pyrolysis plant ............................................................................................................. 38
Biochar characterisation............................................................................................... 40
Sample preparation .................................................................................................... 40
Bulk density .................................................................................................................. 43
pH and EC ..................................................................................................................... 43
Liming potential........................................................................................................... 43
Water holding capacity .............................................................................................. 44
Proximate analysis ..................................................................................................... 44
Calorific value .............................................................................................................. 45
Carbon and nitrogen content................................................................................... 45
Total elemental analyses .......................................................................................... 45
Water extractable elemental analyses................................................................... 46
Available phosphorus ................................................................................................ 46
Cation exchange capacity ........................................................................................ 46
Surface morphology by scanning electron microscopy-electron dispersive
x-ray spectroscopy (SEM-EDX) ............................................................................... 47
Surface area and pore volume determination ..................................................... 47
Zeta potential................................................................................................................ 47
Polycyclic aromatic hydrocarbons and Polychlorinated biphenyls ............. 48
Surface functional groups by fourier transform infrared (FTIR)
spectroscopy ............................................................................................................... 48
Statistical analysis .......................................................................................................... 48
CHAPTER 4. RESULTS ...................................................................................................... 50
Feedstock and apatite characteristics ...................................................................... 50
Physical and chemical characters ......................................................................... 50
FTIR................................................................................................................................. 54
Apatite after pyrolysis ................................................................................................... 55
Biochar characteristics ................................................................................................. 57

7
Morphology................................................................................................................... 57
Biochar yield ................................................................................................................ 61
Bulk density .................................................................................................................. 62
pH .................................................................................................................................... 62
Electrical conductivity ............................................................................................... 63
Liming potential........................................................................................................... 64
Water holding capacity .............................................................................................. 64
Moisture ......................................................................................................................... 65
Volatile matter .............................................................................................................. 66
Ash .................................................................................................................................. 66
Fixed Carbon ................................................................................................................ 67
Calorific value .............................................................................................................. 67
Carbon............................................................................................................................ 68
Nitrogen ......................................................................................................................... 68
Hydrogen ....................................................................................................................... 69
Sulphur .......................................................................................................................... 69
Total and water extractable elements ................................................................... 70
Cation exchange capacity ........................................................................................ 75
Zeta potential................................................................................................................ 75
Surface area.................................................................................................................. 76
Total polycyclic aromatic hydrocarbons (PAHs) and polychlorinated
biphenyls (PCBs) ........................................................................................................ 77
FTIR spectroscopy...................................................................................................... 78
CHAPTER 5. DISCUSSION and INTERPRETATION ................................................... 85
Apatite ................................................................................................................................ 85
Biochar properties in context of literature ............................................................... 85
Feedstocks........................................................................................................................ 87
Pyrolysis time .................................................................................................................. 88
Pyrolysis temperature ................................................................................................... 88
Proximate analysis (moisture, volatile matter, ash, and fixed carbon
content) .......................................................................................................................... 88
Calorific value and CHNS composition................................................................. 89
Bulk density, pH, EC and elemental concentration ........................................... 89
Elemental constituents .............................................................................................. 90
Cation exchange capacity and zeta potential ..................................................... 90

8
Surface area and morphology ................................................................................. 90
PAHs and PCBs ........................................................................................................... 91
Prediction of production conditions using FTIR ................................................ 91
Designer biochar ............................................................................................................. 92
Choice of measurement attributes ............................................................................. 95
CHAPTER 6. CONCLUSION and RECOMMENDATIONS ........................................... 97
REFERENCES ...................................................................................................................... 98

9
List of Tables
Table 1. List of commercial biochar producers in Australia ........................................... 28
Table 2. Selected biochar uses and desired properties ................................................. 29
Table 3. Examples of the use of biochar as a feed additive. ......................................... 33
Table 4. Standard methods for the analysis of physicochemical characteristics ....... 41
Table 5. Chemical characteristics of three feedstocks. Weight units are based on
weight after oven drying at 105 ℃ ...................................................................................... 51
Table 6. Total elemental constituents (acid-digested) the feedstocks and apatite.
Units are presented in terms of weight of oven (40 ℃ for 24 – 48 h) dried material .. 51
Table 7. Change in phosphorus availability in 2 % v/v formic acid solution with
increasing pyrolysis temperature. Units are presented in terms of weight of oven (40
℃ for 24 – 48 h) dried material. There were two replicates per treatment, with a
maximum variation between replicates of 5% of the average value. The % of total P
and Extractable P are calculated for raw apatite. ............................................................ 55
Table 8. Total elemental analysis of apatite with temperature and residence time
variation. Units are presented in terms of weight of oven dried material (40 ℃ for 24
– 48 h) ..................................................................................................................................... 56
Table 9. ANOVA significance testing of impact of feedstock, pyrolysis temperature
and time on biochar yields ................................................................................................... 61
Table 10. Total elemental constituents (acid-digested) of biochar produced at 600 ℃
for 20 min. Units are presented in terms of weight of oven (40 ℃ for 24 – 48 h) dried
material ................................................................................................................................... 70
Table 11. Water extractable elemental content of biochar from different feedstock
produced at varying pyrolysis time and temperatures..................................................... 72
Table 12. Evaluation of the water and total elemental analysis of feedstocks with
biochar at 600 ℃ and 20 min for nutshell, woodchip and stem ..................................... 73
Table 13. PAHs and PCBs of biochar from three feedstocks produced at three
pyrolysis temperatures and three durations ...................................................................... 77
Table 14. Features of mid-infrared spectra of biochar of three feedstocks with
indication of the trend in size of each feature with increasing pyrolysis temperature.
Peak wavenumber for each peak is given for biochar for each feedstock in the order
nutshell, woodchip and stem. ‘-‘ indicates that the peak was not present ................... 78
Table 15. Similar reports of biochar properties in literature ........................................... 87
Table 16. Optimisation of biochar produced from fluidised bed reactor to their
specific uses........................................................................................................................... 93

10
List of Figures
Figure 1. (a) Effect of aerobic thermal treatment on phosphorus solubility in neutral
ammonium citrate solution for apatite deposits from three locations. Data was
sourced from Francisco et al. (2007) to create this figure. b) Phosphorus solubility in
neutral ammonium citrate aerobic thermochemical treatment of sewage sludge ash
mixed with Na and K sulphates. Data was sourced from Herzel et al. (2021) to create
this figure ................................................................................................................................ 36
Figure 2. Fluidised bed batch pyrolyser (pilot plant) at Mara Seeds Soft Agriculture
Company, Mallanganee, NSW, Australia (Winter, 2018) ............................................... 38
Figure 3. Experimental setup sketch of fluidised bed pilot plant pyrolyser (with
permission of John Winter) .................................................................................................. 39
Figure 4. Feedstock materials (macadamia nutshell, eucalyptus woodchip and hemp
stem) and fluidised bed material (ground apatite). Scale bar is 10 mm ....................... 50
Figure 5. Backscattered electron SEM images of feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) and apatite. Scale bar is 100 µm .................... 53
Figure 6. Representative EDX analysis of apatite .......................................................... 53
Figure 7. Mid-infrared spectra of apatite, macadamia nutshell, eucalyptus woodchip
and hemp stem feedstocks .................................................................................................. 54
Figure 8. PCA of macadamia, eucalyptus and hemp feedstocks (Feedstock samples
were oven dried at 105℃ for 24 hours and ground to size <75 µm) ............................. 55
Figure 9. Images biochar produced using pyrolysis duration and temperatures (a)
macadamia nutshell (b) eucalyptus woodchip and (c) hemp stem. Scale bar is 10
mm ........................................................................................................................................... 58
Figure 10. SEM images of macadamia nutshell, eucalypt woodchip and hemp stem
biochar produced at 350, 450 and 600 ℃. Differences in brightness are ascribed to
differences in charging between samples. Red arrows point to apatite fragments.
Blue arrows highlight pores on sclereid cell walls. Scale bar is 50 µm ........................ 60
Figure 11. Yield (% w/w) of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a
feedstock. Bars represent means (n=3) with associated SE ......................................... 61
Figure 12. Bulk density of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a
feedstock. Bars represent means (n=3) with associated SE ......................................... 62
Figure 13. pH of biochar from three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at p < 0.05 within a feedstock.
Bars represent means (n=3) with associated SE ............................................................. 63
Figure 14. EC of biochar from three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at p < 0.05 within a feedstock.
Bars represent means (n=3) with associated SE ............................................................. 63

11
Figure 15. Liming potential of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a
feedstock. Bars represent means (n=3) with associated SE ......................................... 64
Figure 16. Water holding capacity of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at P < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 65
Figure 17. The moisture content of macadamia nutshell, eucalyptus woodchip and
hemp stem biochar produced at three pyrolysis temperatures and three residence
times. Data presented as means with associated SE (n=3). Different letters within a
graph denote significance difference (P < 0.05) .............................................................. 65
Figure 18. Volatile matter content of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at P < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 66
Figure 19. Ash content of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a
feedstock. Bars represent means (n=3) with associated SE ......................................... 66
Figure 20. The fixed carbon content of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at P < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 67
Figure 21. Calorific values of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a
feedstock. Bars represent means (n=3) with associated SE ......................................... 68
Figure 22. Carbon content (%) of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at P < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 68
Figure 23. Nitrogen content (%) of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at P < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 69
Figure 24. Hydrogen content (%) of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
two residence times .............................................................................................................. 69
Figure 25. Sulphur content (%) of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
two residence times .............................................................................................................. 70
Figure 26. The cation exchange capacity of biochar from three feedstocks
(macadamia nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis
temperatures and three residence times. Different letters signify statistical difference
at P < 0.05 within a feedstock. Bars represent means (n=3) with associated SE ...... 75

12
Figure 27. Zeta potential values of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and
three residence times. Different letters signify statistical difference at p < 0.05 within
a feedstock. Bars represent means (n=3) with associated SE ...................................... 76
Figure 28. The surface area of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and two
residence times (n = 1) ......................................................................................................... 76
Figure 29. Mid-infrared spectra of biochar produced using three feedstocks at all
combinations of three pyrolysis temperatures and three pyrolysis durations. Each
line is an average of three spectra per condition ............................................................. 79
Figure 30. Scores biplot for Principal Component Analyses of FTIR spectra of
biochar for each of three feedstocks: (a) macadamia nutshell, (b) eucalypt woodchip
and (c) hemp stem. For each feedstock, biochar was produced at three pyrolysis
temperatures and three durations (n = 27) ....................................................................... 81
Figure 31. Loadings plot for PC1 of Principal Component Analyses of FTIR spectra
of biochar produced from each of (a) nutshell, (b) woodchip and (c) stem. For each
feedstock, biochar was produced at three pyrolysis temperatures and three durations
(n = 27).................................................................................................................................... 82
Figure 32. Scores biplot for Principal Component Analyses of FTIR spectra for
biochar produced from (a) nutshell (Mac), woodchip (Euc), and stem (Hem), and (b)
nutshell and woodchip. For each feedstock, biochar was produced at three pyrolysis
temperatures and three durations (n = 27). In (a), red ovals outline the samples
derived from woody materials (nutshell and woodchip) and from stem material ........ 83
Figure 33. Loadings plot for (a) PC1 and (b) PC2 of a Principal Component Analysis
of FTIR spectra of biochar produced from nutshell and woodchip. For each
feedstock, biochar was produced at three pyrolysis temperatures and three durations
(n = 27).................................................................................................................................... 84
Figure 34. Correlation coefficients (R) between pairs of measured attributes for the
81 samples (three pyrolysis temperature*three pyrolysis
times*3feedstocks*3replicates). Colors are keyed to R values (-1 to +1) .................... 96

13
List of Abbreviations
AC : Ash Content
Al : Aluminium
ALS : Analytical Laboratory Services
ANOVA : Analysis of Variance
As : Arsenic
ASTM : American Standard for Testing and Materials
B : Boron
BC : Biochar
BD : Bulk Density
BET : Brunauer – Emmett – Teller
C : Nitrogen
Ca : Calcium
Cd : Cadmium
CEC : Cation Exchange Capacity
CH4 : Methane
Cl : Chlorine
CO2 : Carbon Dioxide
CQU : Central Queensland University
Cu : Copper
CV : Calorific value
DIW : Deionised Water
EAL : Environmental Analysis Laboratory
EBC : European Biochar Certificate
EC : Electrical Conductivity
FC : Fixed Carbon
Fe : Iron
FTIR : Fourier Transform Infrared Spectroscopy
GHG : Green House Gases
GRDC : Grain Research and Development Corporation
H : Hydrogen
IBI : International Biochar Initiative

14
ICP-MS : Inductively Coupled Plasma – Mass Spectrometry
ICP-OES : Inductively Coupled Plasma – Optical Emission Spectrometry
K : Potassium
LP : Liming Potential
MC : Moisture Content
Mg : Magnesium
Mn : Manganese
N : Nitrogen
Na : Sodium
N2O : Nitrous Oxide
O : Oxygen
P : Phosphorus
PAHs : Polyaromatic Hydrocarbons
Pb : Lead
PCA : Principle Component Analysis
PCBs : Polychlorinated Biphenyls
PV : Pore Volume
S : Sulphur
SA : Surface Area
SEM – EDX : Scanning Electron Microscopy – Energy Dispersive X-Ray
SOC : Soil Organic Carbon
Syngas : Synthetic Gas
UNSW : University of New South Wales
VM : Volatile Matter
WHC : Water Holding Capacity
ZP : Zeta Potential

15
CHAPTER 1. INTRODUCTION
Biochar

The increasing issues of environmental degradation associated with greenhouse gas


emissions, and climate change pose a significant global threat to sustainability of the
ecosystem. Agricultural and biological wastes, landfill and incineration contribute to these
problems. Therefore, immediate and effective management of these wastes is a must. A
large number of agricultural waste materials can be converted to as biochar, which has use
as a fuel source, soil ameliorant, animal feed etc, thus minimising the waste disposal on one
hand and delivering a significant energy source on the other hand (Sohi et al., 2010).

Biochar is the product obtained by heating biomass in the absence or limited supply of air to
>250℃, and the production process is pyrolysis (Lehmann & Joseph, 2015a). Thus, biochar
differs from charcoal and other carbon (C) products, in terms of its intended application in
soil and environmental management.(IBI, 2018b). Biochar is not a new concept, being a
2500-year-old practice since the slashing and charring forest trees resulted in dark, fertile
‘Terra Preta’ soils in the Amazon basin (Glaser et al., 2000; Neves et al., 2003; Smolker,
2015).

Biochar can provide carbon sequestration, retaining carbon for 100s to 1000s of years
(GRDC, 2013). Interest in biochar applications has been growing because of its potential
role in greenhouse gas reduction and carbon sequestration (GRDC, 2013), and the
synthesis of new materials for environmental remediation (Conte et al., 2021). Biochar can
be an excellent adsorbent which enables its application in the removal of contaminants from
water and soil (Siddiqui et al., 2019). Biochar can also act as a nutrient bank by retaining the
nutrients in soils and thus reduce losses in leaching and gaseous emission (Singh et al.,
2010b). Biochar is known to enhance crop productivity by altering the physicochemical and
biological properties of soil (GRDC, 2013). Biochar can also reduce the pesticide residual
effects which are harmful to plants and soils, thus improving soil health and productivity
(Zhang et al., 2013).

Similarly, biochar applications have been extended to animal industries particularly as an


animal feed additive for improving the feed conversion efficiency, toxin absorption, adsorbent
of smells or odours from animal and poultry houses (Prasai et al., 2018). When biochar was
supplemented in animal diet, improved animal growth, improved immunity, reduction in
pathogen loads and reduction of methane production by cattle were reported (Man et al.,
2020).

16
Biochar use in different areas such as animal feed additive, contaminant removal, catalyst
and energy source has been an emerging topic in research (Gelardi & Parikh, 2021). Thus,
research to biochar properties directly impacting biochar quality for its potential uses is of
utmost importance and a user guide for its essential qualities should be established.
However, currently biochar usage is relatively small, largely due to cost of production, lack of
consistent quality and lack of knowledge of the properties of biochar. Some of these quality
constraints can be addressed by developing the biochar specifications for respective areas
and achieving this specification by optimisation of the production process and choice of
feedstocks. High cost of biochar for soil amendment, lack of specification for feed quality
biochar, and difficulties associated with waste management applications have been some of
the major challenges hindering adoption of biochar in agricultural, animal and environmental
sectors.

Project background

In 2012, the Mara Seeds Pty Ltd evolved from a seed production company into SOFT
Agriculture Pty Ltd, with a vision of “carbon smart” farming. The company produced a
diversified range of organic fertilisers and other products for plant and livestock applications
(SOFT, 2012). Biochar production began in 2011 with a purpose-built rotary hearth reactor.
In this reactor, feedstock was fed continuously into the top of the reactor and passed down
through a series of floors, each with a rotating arm to move the material forward. Hot biochar
exited from the reactor base and was immediately quenched in water. Crop residues, waste
plantation timber and green waste were used as feedstock, and biochar was used as a soil
amendment for soyabean, winter cereal and pasture production (Goodwin, 2011). Later,
production was undertaken using an auger reactor under anaerobic conditions producing
highly porous and absorptive biochar. These systems were inefficient as syngas and bio-oil
were burnt for thermal heating.

Mara SOFT then engaged with John Winter to apply the fluidised bed pyrolysis technique
documented in the patent of Winter (2018). This technique involves preheating of a solid
media, such as apatite or ilmenite sand, in a separate oxidation fluid bed and then mixing the
hot sand with biomass in a separate vessel, which can be gas or mechanically fluidised. In
this way the nitrogen from air is kept separate from the syngas, hence the syngas can be
economically converted into useful products, such as hydrogen. This system also allows
scale-up to large plant capacities as it is not limited by heat transfer area.

Mara SOFT purchased two ARTi retorts (https://www.arti.com/) from New Hope coal mine to
liquids conversion trains that were based in Ipswich. These trains had been commissioned
but never used commercially. The plant was dismantled and relocated to the Mara Soft base

17
near Mallanganee, NSW, for re-purposing as a biomass pyrolysis plant. This location is
close to a number of waste organic sources, including timber plantations and cropping
residues. A batch pilot plant was developed to simulate aspects of the full-scale process.
This pilot plant was the foundation of the work completed in this thesis. At the time of writing,
the full-scale plant is in the process of construction.

Mara SOFT commissioned the work of this thesis to address for characterisation of biochar
properties as determined by feedstock and pyrolysis conditions for the various uses of
biochar.

Thesis aims and objectives

The work of this thesis aims to determine optimum pyrolysis conditions (time and
temperature combinations) in a fluidised bed reactor for feedstocks differing in woodiness in
for production of biochar for different applications. Specific objectives include:

• To recommend a set of specifications for biochar production in the context of its


intended use.
• To recommend a choice of methods for assessment of the properties of the resulting
biochar.
• To assess the effect of temperature and residence time on the quality of biochar
produced from a set feedstock.
• To characterise physicochemical properties of the resulting biochar and processed
apatite in relation to their applications in agriculture and animal feeds.

Thus, the study provides a case study on the matching feedstock and pyrolysis conditions
(time and temperature) for a given pyrolysis system to achieve specifications on biochar for
specific applications, with recommendations made for measurement of a reduced set of
attributes in future pyrolysis system characterisation exercises. A fluidized bed pyrolysis
system is employed for its potential for rapid, and thus potentially more cost effective,
biochar production. The impact of the roasting process on phosphate availability of apatite is
also considered, for its potential in providing a secondary output of the pyrolysis system.
Based on the level of correlation between attributes, a recommendation is made for
assessment a reduced number of biochar attributes in future pyrolysis system
characterization exercises.

18
Thesis structure

The thesis is presented in the ‘traditional’ format of Introduction (Chapter 1), Literature
Review (Chapter 2), Materials and Methods (Chapter 3), Results (Chapter 4), Discussion
and Interpretation (Chapter 5) and Conclusion and Recommendations (Chapter 6).

Research publication

The results from the trials presented in the thesis chapter 3, 4, 5 and 6 have been pooled
into a single manuscript “Designing biochar based on substrate and fluidised bed pyrolysis
conditions” and will be submitted into the journal “Biochar”
(https://www.springer.com/journal/42773).

19
CHAPTER 2. LITERATURE REVIEW
Context

This review provides a background on biochar and its production, and then focuses on the
characteristics of biochar required for various applications, including use for animal health
and productivity, wastewater treatment, compost production and soil health. Attention is
given to the definition of attributes that allow characterisation of desired functional attributes,
methods for the analysis of these characteristics, and the definition of pyrolysis production
conditions that impact these characteristics.

The Web of Science, Google Scholar, Elsevier, Scopus, Research Gate, and Science Direct
search engines were interrogated for the years 1998 - December 2021 using the keywords
of “biochar”, “pyrolysis”, “biochar production”, “fluidised bed pyrolyser”, “biochar
characterisation”, “biochar as a feed additive” and “apatite as a modifier in biochar”. Of the
508 articles identified, 220 were selected for reading based on the appropriateness of the
title and abstract. Further research articles were sourced using the keyword of “apatite”,
“phosphate minerals” and “bone char”. No research articles could be tracked with the
combination of apatite and biochar key words in the search.

History of biochar

The term ‘biochar’ was coined in 2005 by Peter Read (Godbey, 2016) but the concept
originated in the pre-Columbian Amazon where the addition of charcoal to soil produced the
“Terra preta” or Black Earths across approximately >350 hectares (Wayne, 2012). These
soils have a carbon (C) content of about 9 % w/w as compared to 0.5 % w/w in the
surrounding soils (Sparks, 2012). Similar practices have occurred in other areas around the
world, creating similar C-rich soils, as reported in Ecuador, West Africa, and South Africa
(Lehmann et al., 2006). Biochar C is recalcitrant, i.e., it will not readily be decomposed, and
so it has a lasting impact on soil properties. Crops planted in Terra Preta soil show a four-
fold increase in plant productivity relative to plantings in the nearby non-biochar amended
soil (Glaser et al., 2002). Yield increase in response to biochar application to soil is generally
reported for highly acidic or nutrient-depleted soils (Dukua et al., 2011).

Biochar and charcoal have traditionally been used in treatment of animals for digestive
disorders, including food poisoning (Schmidt et al., 2019). Totusek and Beeson (1953)
described the use of biochar/charcoal-like materials in pig and poultry feed in the 1880s.
Schmidt et al. (2019) described the use in the 19th and early 20th centuries of ‘cow tonics’
comprised mainly of charcoal for increasing appetite, butter fat content in milk and

20
enhancing milk production. The feeding of charcoal to poisoned patients and animals to
adsorb toxic substances in the gastrointestinal tract (Derlet & Albertson, 1986) or to mitigate
diarrheal symptoms (Mangold, 1936) also has a long history. The use of activated biochar to
reduce the pathogenic toxins from Clostridium botulinum and Clostridium tetani was reported
in the early 1900s (Schmidt et al., 2019).

Biochar has been re-introduced to animal diets a feed additive in recent decades. In Europe,
regulation has followed, with the European Biochar Certificate (EBC) guidelines introduced
in 2006 on the qualities of biochar to be used as feed supplement (Schmidt et al., 2016).

Thus, while production and use of biochar has received increased attention in the last
decade, it has been in traditional use for centuries.

Biochar uses

In recent decades, application of biochar for a range of uses has been reported, including for
soil improvement, carbon sequestration, as an animal feed supplement and for adsorption of
heavy metals in wastewater treatment (Conte et al., 2021).

Biochar application to soil can result in increased water holding capacity (in sandy soil),
increased soil thermal properties and reduced soil tensile strength. Biologically, the addition
of biochar to soil has been shown to increase mycorrhizal fungi colonisation of roots and
improve nitrogen fixation by legumes (GRDC, 2013) and increase earthworm populations
(Oni et al., 2019). However, the impact of addition of biochar to soil on crop production
varies from positive to negative across different reports (GRDC, 2013). This variation will
reflect the nature of the factors limiting crop production at a given site and the properties of
the biochar added.

Depending on feedstock and production, a fraction of biochar in soil, subject to


environmental weathering and biological activity, is able to persist for thousands of years
(Huckfeldt, 2020) while also contributing to improved soil quality. There is therefore potential
for sale of C credits associated with this sequestration. Indeed, Pacific Biochar reported the
first such sale in the US in November 2020 (https://pacificbiochar.com/first-biochar-carbon-
credits/). With the current value of a C credit (1 tonne of CO2) at $ 50 in Australia
(Mazengarb, 2021), such a market could provide a supplementary income for a biochar user.

Biochar is recognised as a potential animal feed supplement, having a beneficial impact on


the digestion and absorption of nutrients and water (Schmidt & Wilson, 2014), with enhanced
health, performance and productivity of animals reported for pigs, poultry and ruminants (Lao
& Mbega, 2020). The mechanism for this improvement is attributed to the adsorption by

21
biochar of toxins and possibly specific pathogens (Galvano et al., 2001). Biochar inclusion in
animal diet is also reported to result in decreased methane emission thus aiding in the
reduction of greenhouse gas emissions (Teoh et al., 2019). As in soil application, benefits
can be variable, and matching of limiting factor for animal production to biochar property is
required. Further, biochar is typically included in diet as a macroconstituent, at levels of
between 1 and 4 %. Such levels result in decreased levels of other constituents, such as
protein, if the same feed materials are used. Other alternative raw materials of higher feed
values can be used to address nutrient dilution, although such inputs are likely to become
more expensive.

Biochar also acts as an absorbent, with absorption of heavy metals and contaminants from
the wastewater and the environment (Uchimiya et al., 2011).

To advance these uses, the characteristics of biochar required for the intended application
should be documented. The production conditions to achieve these characteristics can then
be determined, e.g., through the combination of feedstock and co-additions, and time and
temperature of pyrolysis (Tomczyk et al., 2020; Tripathi et al., 2016).

Biochar production

Pyrolysis method

In the combustion of organic material in an oxygen-rich environment, all carbon can be


oxidised, leaving only minerals (ash). If heated in an anaerobic environment, a purer carbon
product, charcoal, is produced. Incomplete combustion or charring in an anaerobic
environment (pyrolysis) result in charcoal-like material, biochar, that retains carboxylic and
other chemical groups. Pyrolysis is defined as the thermal decomposition of biomass in
limited supply or absence of oxygen, at temperatures between 300 - 700 ℃ (IBI, 2018c).

The pyrolysis of organic materials generates an acidic oil (bio-oil), combustible gas (syngas),
and char (biochar) (Zhu et al., 2018). The conditions of the pyrolysis process influence the
relative yield of these three product classes, with lower temperatures and longer residence
time generally favouring biochar production (Zhu et al., 2018) and higher temperature and
faster reaction time yielding more bio-oil and syngas. A pyrolysis system can have a
condensation collection option to separate bio-oil and bio liquor from the resulting pyrolysis
gases, as well as a system to recover biochar (Dhyani & Bhaskar, 2018). The products of
pyrolysis are biochar, bio-oil, and syngas.

The bio-oil produced is comparable to crude oil because it can be fractionated and upgraded
to form biodiesel and other organic chemicals and transportation fuels (Wang et al., 2013a).

22
However, the process is complex and significant processing problems have been reported
during the conversion. Studies on the cost-effectiveness and readily adaptable mechanism
for utilization of bio-oil could open future markets.

Similarly, the gasification of organic material in the pyrolysis process results in syngas
(synthetic gas), consisting mainly of carbon monoxide and hydrogen, with traces of methane
and other hydrocarbons. Syngas is much less energy-dense than natural gas, as it contains
carbon that has been mostly oxidised (Cheruvu, 2018). Syngas can be catalytically
converted to transportation fuels and chemicals, such as gasoline, diesel, wax, and jet fuel
(Cheruvu, 2018).

Biochar properties are highly affected by pyrolysis conditions (temperature, time and
feedstock). There are many variants on pyrolysis method, with one classification based on
treatment time. Slow pyrolysis involves a slow heating ramp (minutes to hours) of biomass to
a desired temperature followed by a residence time of minutes to hours to ensure all material
has reached the desired temperature (Mohan et al., 2006). Slow pyrolysis systems are
typically batch systems involving loading of a reaction vessel with a relatively large volume of
biomass (e.g., 1 tonne) and heating of the external vessel or the blowing of heated gas
through the vessel, thus requiring a relatively slow heating time. Slow pyrolysis results in
higher biochar yield compared to fast pyrolysis for a given production temperature. Lower
heating rates suppresses thermal cracking leading to higher char production (Yadav &
Jagadevan, 2019). For a given production temperature, slow pyrolysis results in biochar with
more surface charged groups, and thus more active surface chemistry (Tomczyk et al.,
2020) and higher CEC (Ippolito et al., 2020). However, product temperature may be more
variable within the reaction vessel, depending on the mode of heating.

Fast pyrolysis is characterized by a method that achieves a fast-heating rate of


approximately 100 oC per second to approximately 400 – 600 oC. Such a rapid heat transfer
to the feedstock is typically accomplished with fluidised beds, heating blades, or screw mixer
reactors. In a fluidised bed system, the residence time can be as short as 0.5 – 10 s (IBI,
2018c). A short residence time of 0.5 – 10 s minimizes secondary cracking of the pyrolysis
products (Wang et al., 2013b), leading to lower char production (Yadav & Jagadevan, 2019).
More feedstock is converted to bio-oil (53 vs. 30 % w/w) and syngas than biochar (31 vs. 43
% w/w) during fast pyrolysis in a fluidised bed reactor than in slow pyrolysis, respectively
(Mohan et al., 2006; Wang et al., 2013a). Thus, fast pyrolysis is suited to bio-oil production
for use in upgrading to alternative energy fuels.

23
Pyrolysis reactor types

Pyrolysis reactors can be categorised as:

Fixed bed reactors


The fixed bed reactors are characterised by the solids remaining in a fixed position in the
reactor and with the supply of heat, the solids undergo chemical reactions (Andrigo et al.,
1999; Eigenberger, 1992). Fixed bed reactors operate using an external heat source in a
batch mode (Bamido, 2018). Feedstock is typically introduced through a lid in the top of the
vessel, which is then sealed for the duration of the reaction, and product extracted via a port
on the bottom of the vessel. Pyrolysis duration and temperature is under operator control,
with maximum heating rate set by the design of the system. The major drawbacks of fixed
bed reactors lie in inefficient heat transfer to the feedstock and difficulty in regulation of
temperature (Ibrahim, 2020). Thus, fixed-bed reactors are commonly employed for the slow-
pyrolysis aiming to higher biochar yield (Raza et al., 2021). Scale-up is limited by the rate of
heat transfer, hence multiple reactors are required to achieve large capacity.

Auger reactors
Auger reactors are characterised by a helical screw rotating in a closed tubular shell wherein
the rotation of the screw conveys the feedstock material into the reactor for carbonisation
(Campuzano et al., 2019). The pyrolysis process is enhanced with the efficient mixing and
transport of biomass particles effected by the screw drive (Bamido, 2018). Screw auger
systems operate with the external heat exchange. The speed of auger and the inert gas (N2)
flow rate regulates the solid and gas residence time inside the reactor (Campuzano et al.,
2019). Biochar is moved out of the heating zone to a collection point by the screw, while
gases and volatiles are directed to the condenser or afterburner (Raza et al., 2021). Thus,
auger reactors are popular for pyrolyzing wide ranges of feedstocks to obtain the desired
product (biochar, oil or syngas) operating at slow, intermediate and fast conditions, which in
fact, is not suitable for other reactors. However, the potential for blockages through
compaction of biomass inside the feeder might be the drawback for its application in large-
scale plants (Strezov & Evans, 2014).

Rotary kiln reactors


Rotary kiln reactors are the most effective reactor for the pyrolysis solid waste of variable
shape, size and heating capacity operated in batch or continuous modes (Li et al., 1999).
There is direct heat exchange resulting in the partial combustion of the material thus
generating lower quality biochar and syngas is highly diluted with N2. When the feedstock is
fed from a supply or screw feed, the slow rotation and inclination facilitates the transport of

24
feedstock into the kiln (Li et al., 2004). The residence time and operating conditions can be
regulated for the desired yield of the product (Li et al., 2004).

Fluidised bed reactors

Fluidisation refers to the process through which a fluid like state is attained by the solid
particles through suspension in a flowing gas or liquid (Thomas & Fucher, 2000). A fluidised
bed reactor can use a bubbling or a recirculating bed depending on the fluidisation medium
velocity. A bubbling fluidised bed functions at comparatively low gas velocities (< 1 m s-1)
compared to a circulating fluidised bed (3 – 10 m s-1) (Siedlecki et al., 2011). In a bubbling
fluidised bed, hot gas (inert or flue) is pumped through a porous distributor plate, and with
the attainment of minimum fluidisation velocity, the bed particles act like a fluid (Strezov &
Evans, 2014). A circulating fluidised bed differs in terms of use of a higher velocity of gas
flow, resulting in an expanded bed and the circulation of bed particles in a constant loop of
cyclone and loop seal (Strezov & Evans, 2014). In a circulating fluidised bed, the produced
biochar can be separated from the reactor and can be reused to supply heat back to the
reactor for higher bio-oil yield (Siedlecki et al., 2011).

A fluidised bed reactor has the advantage of rapid and uniform heat transfer to the feedstock
materials, resulting in fast processing and production of a less variable product. Both the
bubbling and circulating fluidised bed reactors are highly favoured for high bio-oil yield
(Raza et al., 2021).

Ablative reactors

Ablative reactors involve heat transfer from the hot reactor wall to melt the wood (feedstock)
which is in contact to the reactor surface under high pressure (Bridgwater, 2012). As
mentioned by Bridgwater (2012), the reaction is similar to melting a butter in a hot pan i.e.,
higher melting is favored by pressing down and making a motion on the butter over the hot
pan. The decomposition rate is proportional to the pressure of the material on the heated
surface, the relative velocity of the materials and the reactor surface temperature
(Bridgwater, 2012). Thus, ablative reactors are very efficient in a smaller reactor volume,
with lower operation and capital costs, minimal feed preparation, and with no requirement for
recycling of gas, compared to other reactor types (Peacocke & Bridgwater, 1994). Ablative
reactors allow for fast pyrolysis, achieving higher yield of bio-oil (Peacocke & Bridgwater,
1994).

Entrained flow reactors


The entrained flow reactor consists of a preheated vertical tubular reactor wherein the
feedstock materials are loaded and operating in a co-current flow with carrier gas (Maliutina

25
et al., 2017). Thus, the higher velocity of gas flow facilitates the higher heating of the reactor
resulting in the decomposition of biomass materials (Tchapda & Pisupati, 2015). For the
maximum conversion potential of the reactor and the shortest residence time (few seconds),
a smaller size of feedstock (100 µm) and a temperature > 1000 ℃ is preferred. Entrained
flow reactors operates at extreme temperature and heating rates for better carbon
conversion and cleaner syngas (Tchapda & Pisupati, 2015). The reactor is mainly applicable
for fast pyrolysis when operated at higher pressures and temperature in large plants (Zhang
et al., 2015). However, the insufficient heat transfer between gas and solid particle limits its
use and application (Bridgwater, 2012).

Temperature

Pyrolysis temperature strongly influences biochar yield and properties (Zhao et al., 2017).
Biochar yield is inversely proportional to temperature i.e., with increasing temperature
decreasing biochar yield (Ashwath et al., 2018; Nam et al., 2015; Sun et al., 2017a). This
trend is attributed to the decomposition of cellulose and hemicellulose at a higher
temperature. Physicochemical properties affected by pyrolysis temperature include a
reduction in the amount of nitrogen, hydrogen, and oxygen along with volatile materials (Tag
et al., 2016). A higher carbon content, higher pH, higher surface area, and nutrient/ash
content are achieved by higher temperature pyrolysis (Askeland et al., 2019; Mašek et al.,
2013). Volatile matter content decreases with increasing temperature (Bolan et al., 2021).
However, a decrease in volatile matter is also associated with decreased availability of
organic matter to soil microbes and thus degrades the value of the biochar for use in
improving soil health (Ding et al., 2016).

Residence time

At a given temperature the pyrolysis reaction will proceed through a rapid reaction phase
and then a slow reaction period involving reaction of more recalcitrant material. Residence
times that are too short will obviously result in product that has not completely reacted.
These times will depend on the nature of the reactor, i.e., the speed of heat transfer. The
required reaction time is longer at a lower temperature (Sun et al., 2017a). For example,
yield of biochar produced from shredded cotton stalk in a fixed bed reactor decreased from
70.2 to 64.8 % with the increase of residence time from 60 to 240 min at 200 ℃, however,
the effect was insignificant at higher temperature > 400 ℃ (Makavana et al., 2020). Thus,
increasing residence time at higher temperature leads to secondary cracking rather than
decreasing yield (Mohd Hasan et al., 2019).

26
Quenching of biochar

Quenching is the process of cooling of biochar immediately after pyrolysis and before
exposure to an oxygen rich atmosphere to prevent biochar combustion. Quenching of
biochar with water lowers the temperature of biochar, eases handling issues by removing the
dust hazard associated with dry product (Marrero, 2018). Quenching cleans the biochar and
increases the pore volume and surface area of biochar (Schmidt et al., 2014). If nutrient rich
water is used during quenching, the biochar CEC can be ‘charged’ with essential nutrients
(Joseph et al., 2018b). The quenching of biochar increased surface area and pore volume
thus enhancing the adsorption capacity of biochar, and use of cattle urine for quenching
increased the fertiliser value of the product (Schmidt & Wilson, 2014).

Feedstocks

Globally, large amounts of agricultural, animal and farm wastes are either burned or
decomposed, resulting in greenhouse gas emissions and pollution of land and water with
minerals and sometimes with pathogen contamination (IBI, 2018d).

Ideally, the feedstock used for biochar production should not compete with the use for food
production, including indirect uses such as compost production. This situation is unlikely
given the relative economic value of inputs for the two markets. Biochar is commonly
produced from waste organics such as field crop residues, nutshells, fruit pits, bagasse, food
wastes and animal wastes (IBI, 2018d). Feedstocks for pyrolysis should be selected based
on local availability to reduce the cost of collection and transportation (IBI, 2018a).
Feedstocks should be checked for the presence of heavy metals and toxic compounds (IBI,
2018d).

The properties of the feedstock will in part determine the properties of biochar produced,
such as nutrient content and functional groups (IBI, 2018a). For example, biochar derived
from wood is high in carbon and low in mineral content relative to biochar produced from
herbaceous biomasses and manures (Lehmann & Joseph, 2015b). Lignin pyrolysis creates
phenolic compounds which include guaiacol-like, phenol-like, syringol-like, and catechol-like
systems (Liu et al., 2017). Low lignin and nitrogen-containing biomasses are best for the
production of bio-oil and fuel gas (Conte et al., 2021). High lignin-containing feedstocks
produces higher yield (% w/w) of biochar (Cha et al., 2016). Thus, the choice of feedstock
should be based on the required properties of the biochar in its intended use.

27
Commercial pyrolysis systems in Australia

The use of biochar attracted increasing interest in the first decades of the 21st century, with a
number of entrants to the field producing pyrolysis systems in Australia (Table 1). The
relatively high cost of production of biochar resulted in high biochar prices (e.g.,
CQUniversity paid $500 per tonne of biochar from Pacific Pyrolysis Ltd in 2010, while Green
Man Char, purportedly the largest producer in Australia, was reported to sell biochar at
$1600 per ton (Young et al., 2019). These high prices result in little commercial demand.
Further, the harsh conditions of pyrolysis (acidic products, high temperatures) result in rapid
corrosion of the reactor vessels. As a consequence, a number of producers have exited the
market (Table 1).

Table 1. List of commercial biochar producers in Australia

Commercial Producers in Australia Location Pyrolysis reactor Years of


used operation
Pacific Pyrolysis Pty Limited Somersby, Continuous flow 2006 - 2010
(https://www.agriculture- NSW slow pyrolysis rotary
xprt.com/companies/pacific-pyrolysis-pty- kiln reactor
limited-39055)
Pyrotech Energy Melbourne, Circulating fluidised 2016 –
(https://pyrotechenergy.com/) VIC bed reactor present
Renewable Carbon Resources Australia Charleville, Traditional charcoal 2005 -
(RCRA) (https://rcra.com.au/) Qld pit manufacture present

Grayson Australia Bayswater Not Available 1992 -


(https://www.graysonaustralia.com/biochar) VIC present

Earth Systems Hawthorn, batch pyrolysis 1993 -


(https://www.esenergy.com.au/ ) VIC present
Rainbow Bee Eater Melbourne, Not Available 2008 -
(https://www.rainbowbeeeater.com.au/) VIC present

Charman (http://www.charman.com.au/) Pyalong, Rotary kiln reactor 2012 -


VIC, present

Ecoreps (https://www.ecoreps.com.au/) Welland, SA Not Available 2009 -


present
Energy Farmers Australia Pty Ltd Geraldton, Auger reactor 2010 -
(https://www.energyfarmers.com.au/) WA present
Biochar Industries Uki, NSW TFOD (top fed open 2011 -
(http://biocharindustries.com/) draft) kiln present
Mara Seeds Pty Ltd Mallanganee, Multi-hearth reactor 2005 -
(https://maraseeds.com.au/) NSW present

28
Biochar characterisation and functionalities

Attributes

Biochar is typically characterised in terms of (i) porosity, (ii) density, (iii) pH, (iv) cation
exchange capacity, (v) water holding capacity, (vi) fixed carbon content, carbon to nitrogen
ratio, (vii) macronutrients, (viii) zeta potential, (ix) calorific value, (x) surface area and (xi)
functional groups (Oni et al., 2019). The potential value of such attributes is summarised in
Table 2.

Table 2. Selected biochar uses and desired properties

Beneficial applications Biochar desirable References


properties
Soil amendment
The higher the porosity the higher the retention of water thus Surface area (> 150 (Asai et al., 2009b)
increasing the soil water holding capacity and improving in yield. m2/g)

Biochar addition reduces the density of soil and facilitates soil Bulk density (0.06 – 0.7 (Blanco-Canqui, 2017)
aeration. g/cm3)

Alkaline biochar can treat acidic soil minimizing Al toxicity with pH (>7) (Schmidt et al., 2016;
the liming effect. Tsai et al., 2012)
(Shetty & Prakash,
2020)
Addition of a high CEC biochar to a low CEC soil will improve Cation exchange (Ding et al., 2016;
nutrient retention. capacity (5 to 50 Munera et al., 2017; Wu
cmol/kg) et al., 2012)
Addition of a high-water holding capacity biochar to a low water Water holding capacity (Yu et al., 2013)
holding capacity soil (sand) will improve plant available water.
Recalcitrant carbon is sequestered carbon and will remain in the Recalcitrant carbon (Chan & Xu, 2012;
soil for thousands of years. content (> 50 %) Rehrah et al., 2014;
Schmidt et al., 2016)
Volatile and non-recalcitrant carbon can be used by soil Higher surface area (Melas, 2014)
microbes. The higher surface area and porosity provides (150 m2/g)
favourable habitat for microbial activity and growth
The mineral content of biochar has fertiliser value. It is highly Mineral content, ash (Ding et al., 2016)
related to the feedstock; thus, biochar tends to increase the content
available nutrients and increase soil fertility.
Lower zeta potential is associated with higher surface negative Zeta potential (Jatav et al., 2021; Yang
charge, and thus electrostatic attraction of heavy metals and (- 20 to - 50 mV) et al., 2020; Yuan et al.,
pollutants from soil and water. 2015)

Energy fuel source


A higher calorific value increases suitability of biochar as a fuel Calorific value (Dąbrowska et al., 2018;
source. (24 to 31 MJ/kg) Suman, 2020; Yang et
al., 2017)
Animal feed additive

29
Higher surface area and porosity are associated with increased Surface area and (Schmidt et al., 2019)
adsorption capacity of biochar and thus potential for toxins and porosity (> 150 m2/g)
pathogen removal.
Higher levels of surface functional groups and surface area is Surface functional (Sun et al., 2017b; Teoh
associated with higher reactivity, and thus enables increased groups and surface area et al., 2019)
feed intake, digestive efficiency, and methane reduction (> 150 m2/g)
Wastewater treatment
Higher surface area and presence of surface functional groups Surface functional (Bolton et al., 2019;
facilitates the higher adsorption of heavy metals from groups and surface area Yang et al., 2020)
contaminated water (> 150 m2/g)

C sequestration and fuel value: carbon content

The high level of recalcitrant carbon in biochar (65 – 90 % C, depending on the feedstock
used and pyrolysis conditions) provides for carbon sequestration applications (Chan & Xu,
2012). Indeed, the half-life of biochar in soil is estimated at between 100 to 1000 years
(Spokas, 2010) differing based on feedstock, pyrolysis conditions and soil conditions (Bai et
al., 2013). Higher pyrolysis temperatures are associated with lower yield of carbon in biochar
(Rehrah et al., 2014) but higher biochar carbon concentration (Cao et al., 2018) and more
recalcitrant material, with longer half life in the environment. Recalcitrant alkyl and aromatic
carbon content is increased with increase in pyrolysis temperature from 250 to 450 ℃, while
above 500 ℃ the very recalcitrant poly-condensed aromatic hydrocarbons dominate, with
decreasing alkyl carbons. Thus higher pyrolysis temperatures are associated with more
stable biochar which is resistant to microbial and abiotic degradation (De Pasquale et al.,
2012; Lehmann & Joseph, 2015b).

Biochar can be used as a ‘renewable’ fuel (Siddiqui et al., 2019; Wardani et al., 2018). The
calorific value of biochar is directly related to the carbon content, which is increased with
increased pyrolysis temperature (Mierzwa-Hersztek et al., 2019; Wardani et al., 2018).

Soil improvement: CEC, zeta potential, mineral content, density, porosity, pH,
and liming potential

Positively charged mineral ions can be retained by the negative charges of biochar, as
characterised by CEC. The cation exchange capacity (CEC) of biochar generally ranges
from 5 to 50 cmol (+) kg-1 (Glaser et al., 2002; Munera et al., 2017). Biochar CEC is generally
decreased at higher pyrolysis temperatures, as functional group content is decreased.

Zeta potential is the measure of electronegative strength and stabilising nature of biochar in
a colloidal solution (Hong et al., 2019). Zeta potential and CEC are correlated, with both
related to the presence of negative charges, and thus to retention of cations (Batista et al.,
2018). All biochars show a negative charge across a wide range of pH values, based on

30
which cations are adsorbed (Batista et al., 2018; Yargicoglu et al., 2015). A higher zeta
potential, i.e., weak electronegativity, is favoured by higher pyrolysis temperature, resulting
in a declining negative charge on the surface (Hong et al., 2019). Zeta potential value is also
decreased with increasing pH, as the surface charge on biochar becomes more negative
(Meng et al., 2014; Peng et al., 2017). Thus, increased pyrolysis temperature is associated
with increased zeta potential value and decreased CEC (Batista et al., 2018).

Biochar also has direct fertilser value due to its water-soluble mineral content. Biochar
mineral content is highly dependent on feedstock, both of biomass and any inorganic
amendments, but also pyrolysis conditions. The concentration of plant essential elements
will tend to increase as total biochar mass decreases and ash content increases with
increasing pyrolysis temperature. It is generally expected that the concentration of Mg, Ca,
K, and P in biochar is increased with increased pyrolysis temperature in the ash content
(Zama et al., 2017). For example, when produced at 300 and 750 ℃, respectively, biochar
produced from peanut shells contained 11 and 31 g K kg-1, 0.14 g and 1.18 g S kg-1, 5 and
13 g Ca kg-1, 1 and 3 g Mg kg-1, 0.10 and 1.20 g Fe kg-1, 1.37 and 3.30 g Mn kg-1 and 1.5
and 6.0 g Zn kg-1 (Rehrah et al., 2014).

Biochar bulk density typically ranges from 0.06 – 0.7 g/cm3. The lower the bulk density of
biochar, the more porous the biochar, the higher the water infiltration rate, and the higher the
water holding capacity (Askeland et al., 2019). The density of soil can be reduced when
biochar with low bulk density (<0.6 g cm-3) is applied (Blanco-Canqui, 2017), leading to the
improvement in root penetration (Atkinson et al., 2010) drainage and soil aeration.

Increased pyrolysis temperature from 250 to 750 ℃ was associated with an increase in the
water holding capacity of biochar from 7 to 16 % (Yu et al., 2013). Biochar application can
ameliorate a low soil water holding capacity, i.e., sandy soils, but high rates of application
are required (Yu et al., 2013). Yu et al. (2013) reported a use case for a biochar (400 ℃ for
3 hr) with a water holding capacity of 300 % in amelioration of a sandy soil of low water
holding capacity of < 20 %.

Biochars are generally alkaline, although biochar pH can range from 3.1 to 12.0 (Singh et al.,
2017b), and thus soil amendment with biochar generally results in soil pH increase (Kloss et
al., 2012). Biochar liming potential values 3.3 to 30.2 % are commonly reported (Jiang et al.,
2017). The lower pH in some biochar is associated with lower ash content and is thus
related to the feedstock material used, e.g., biochar from wood feedstock shows lower pH
than that from crops, manures and grasses (Singh et al., 2017b). Similarly, higher
temperature also positively influences the pH of biochar (Hossain et al., 2011). A desired

31
biochar pH for an intended soil application can therefore be achieved by a suitable selection
of feedstock and pyrolysis temperature (Lehmann et al., 2006).

Absorption of toxins: Porosity and surface area

The related attributes of biochar surface area and porosity are increased with higher
pyrolysis temperatures as pore voids are generated with the release of volatiles (Conte et
al., 2021). The nature of feedstock material is another factor influencing the surface area and
porosity, as woody biomass tend to have higher surface area than manure/waste biomasses
(Leng et al., 2021). The surface area of biochar is commonly reported to be within the range
of 8 – 490.8 m2/g, however, the value can be increased with the optimisation of pyrolysis
conditions as well as with the physical and chemical treatment of biochar (Chen et al., 2008).

The porosity and high surface area of biochar facilitates absorption (Manariotis et al., 2015)
of both inorganic and organic compounds to the biochar surface (Cornelissen et al., 2005;
Karhu et al., 2011). Adsorption of Mn, As, Pb, Cd, Cu and other heavy metals by biochar has
been reported by (Qambrani et al., 2017). In one application, the absorption of heavy metals
from polluted soil has been reported by (Park et al., 2011). Energy Dispersive Xray (EDX)
analysis can be used to assess metal distribution in biochar (Bolan et al., 2021; Liang et al.,
2016).

Biochar surfaces contain functional groups predominantly, aliphatic alkyl and ester groups
degrade with temperature and reveal the aromatic core lignin, which results in higher surface
area (Chen & Chen, 2009). The higher the surface area, the higher the absorption of biochar
for toxins and heavy metals, allowing application of biochar in heavy metal remediation uses
(Qambrani et al., 2017).

Another application area is the use of biochar as an animal feed supplement, to enhance
toxin binding properties for the improvement of animal health and productivity (Schmidt et
al., 2019). Indeed, charcoal has long been used to correct digestive disorders in animals and
humans, however, no international guideline exists for its use (O’Toole et al., 2016; Zellner
et al., 2019). For example, Prasai et al. (2018) reported increased chicken feed conversion
efficiency and production with a 2 % w/w amendment of biochar as a feed additive. Other
examples are summarised in Table 3. The extent of toxin absorption by biochar will differ
depending on the toxin chemistry, the animal's body mechanisms, the feedstock used for
biochar, the rate of inclusion in the feed, and the particle size of the biochar (Schmidt et al.,
2019).

32
Table 3. Examples of the use of biochar as a feed additive.

Biochar Addition to Feed Benefit Reference


5 % w/w oakwood biochar Chicken layer production improved (Kutlu et al., 2001)
5 % (w/w) beechwood Chicken broilers body weight gain by 23 % (Kalus et al., 2019)
biochar and feed intake increase of 17 % on control
with increased feed conversion ratio
1 g kg-1 bamboo biochar Digestibility of dry and organic matter (Mui & Ledin, 2006)
improved; 17 % increase in body mass
increment of young goats
1 % w/w Jarrahwood biochar Increment of cattle body weight over two (Robb & Joseph,
months increased by 10.4 % compared to 2019)
control
4 % w/w woody green waste Average egg weight increment by 3 %, (Prasai et al., 2018)
biochar enhanced feed conversion ratio by 8 % and
lowered feed intake by 2 %

Polycyclic aromatic hydrocarbons and Polychlorinated biphenyls

Polycyclic aromatic hydrocarbons (PAH) are organic compounds with two or more fused
benzene rings in their structure (Kim et al., 2013). Some of these compounds are
carcinogenic, mutagenic and teratogenic (Adeniji et al., 2018). PAHs are formed during the
incomplete combustion of feedstocks (Wang et al., 2017). Similarly, polychlorinated
biphenyls (PCBs) are formed during pyrolysis of feedstocks especially with high amount of
chlorine such as PVC, seawater, salt etc (Hale et al., 2012). Thus, feedstocks containing
chlorine should be avoided to minimise the presence of PCBs. The EBC guideline on
maximum PAH content of biochar is 6 - 20 mg/kg for 16 EPA PAHs, with a level of 4 mg/kg
for a premium product and < 12 mg/kg for basic grade biochar (Schmidt et al., 2016; Wang
et al., 2017). Similarly, the threshold limit for total PCB is <0.2 mg/kg and 0.2 – 0.5 mg/kg as
per the EBC and IBI guidelines respectively (Bucheli et al., 2015).
Biochar PAHs are typically lower than the maximum limit as given in EBC guidelines, and
their concentration decreases markedly within five years of biochar application to soil
(Rombolà et al., 2019). However, reported PAH levels in biochar range from < 0.1 to >
10,000 mg/kg (Wang et al., 2017). High levels are more likely to be associated with fast
pyrolysis, in which there is more chance that PAHs will condense onto the biochar surface,
than with slow pyrolysis, in which PAHs are released in gaseous form (Garcia-Perez &
Metcalf, 2008; Quilliam et al., 2013). Similarly, as compared to PAHs, PCBs are unlikely to
be toxic in soil or animal as they are strongly bound than PAHs (Hale et al., 2012). As the
total PCB is related to the total chlorine content, if the feedstock containing low chlorine is
pyrolysed then the resultant biochar is reported to have very low PCB (Chen & Chen, 2009).
Temperature between 400 to 500 ℃ have been reported to yield higher PAHs than higher or
lower temperatures (Devi & Saroha, 2015; Keiluweit et al., 2012).

33
Conversely, biochar is highly recommended for the adsorption of hydrophobic materials,
including PAHs (Conte et al., 2015). Thus, biochar application in soil, environment, and
animal feed also relies on the amount of PAHs present, thus PAHs should be within the
guidelines of IBI and EBC.

Functional groups

With the pyrolysis of biomass, the initiation of decomposition of cellulose, hemicellulose, and
lignin leads to the formation of functional groups (carboxyl, hydroxy, lactone, anhydride,
phenols, ester, pyridine, and aromatic carbons) (Mia et al., 2017). The hydrogen and
oxygen-containing functional groups degrade owing to the dehydration and deoxygenation of
the biomass material with increasing temperature and heat during pyrolysis (Uchimiya et al.,
2011). The higher the temperature, the higher the decomposition of biopolymers and
dissociation of functional groups, resulting in the aromatic carbon structure (Tomczyk et al.,
2020). The decrease in CEC is attributed to the removal of functional groups (hydroxy and
carboxyl) and the formation of aromatic carbon structures, thus, lower pyrolysis
temperatures biochar are recommended for production of biochar intended for soil CEC
amendment (Joseph et al., 2010).

Apatite

Bed properties

In a fluidised bed pyrolysis system, a finely ground material of high heat capacity should be
used as the bed material. The bed material serves as a heat carrier, with efficient heat
exchange resulting in uniformity in the final pyrolysis product (Patel et al., 2019). A bed
material should therefore have a high heat capacity while not being prone to ash fusion.
Sand, lime and ground limestone are commonly used as bed materials given their ready
availability and low cost (Patel et al., 2019).

Winter (2018) proposed the use of ground apatite (rock phosphate) as a bed material.
Apatite is a naturally occurring calcium phosphate with the molecular formula of
Ca10(PO4)6X2, where X refers to F (fluorapatite), OH (hydroxyapatite) or Cl (chlorapatite)
Mullen (2005). Apatite has a heat capacity of around 700 J.mol-1.K-1 (Drouet, 2015). This
value is comparable to that documented for a range of fluidized bed materials used in heat
transfer exercises, e.g., 750 for glass beads and 830 for sand (Trajano et al., 2013).

34
Fertiliser potential

Apatite or rock phosphate is the primary source of mineral phosphorus used in agriculture,
although its low solubility in water limits direct use as a fertiliser (Mullen, 2005). Indeed while
the N:P:K (w/w) composition of apatite is typically 0:30:0, it is typically described as a 0:3:0
fertiliser in respect to the available P (Barker, 2019). The common production method for
agricultural phosphate fertiliser involves the use of sulphuric acid to dissolve phosphate from
the apatite, producing phosphoric acid (Shaw, 1959). However, this is a chemically harsh
method and results in the co-dissolution of heavy metals and the evolution of the toxic
fluorine gas if fluorapatite is present.
A general description of the effect of heating of apatite is provided by Kaia et al. (2011).
Heating to temperatures up to 200 ℃ results in a reversible loss of adsorbed water without
any effect on lattice parameters, while increasing temperature from 200 to 400 ℃ results in
loss of lattice water, resulting in a contraction of the a-lattice dimensions (Tanaka et al.,
2000). Surface phosphorus readily forms a P-OH bond which are dehydroxylated above 400
℃ to form P-O-P surface groups. Dehydroxylation of hydroxyapatite initiates at around 900
℃. Lattice stability is also impacted by the presence of minerals that can affect substitutions
(Kaia et al., 2011).

Hydroxyapatite heated at between 300 to 500 ℃ in aerobic conditions (muffle furnace) for 1
h has been reported to result in increased phosphorus availability, relative to heating to
lower or higher (900 and 1050 ℃) temperatures (Glæsner et al., 2019). Thus bone char, as
used in ‘blood and bone’ products, is a more effective source of phosphorus than a dried
bone meal (Glæsner et al., 2019). Herzel et al. (2021) report that extractable phosphate from
sewage sludge ash increased around 40 to over 80 % following thermal treatments over 800
℃ (Fig. 1). Working with rock phosphate from three locations, Francisco et al. (2007)
reported an increase in phosphate availability (extraction in a neutral ammonium citrate)
from < 1 % to over 80 % following heating for 2 hours in aerobic conditions (muffle furnace).
The temperature optimum varied from 500 to 700 ℃ with the source of the mineral, i.e., with
the extent of chemical substitutions and lattice structure (Fig. 1).

There is therefore potential that the use of apatite as fluid bed material in the pyrolysis
process may result in improve phosphate availability, creating material of value as a fertiliser
as a by-product of the pyrolysis process.

35
Figure 1. (a) Effect of aerobic thermal treatment on phosphorus solubility in neutral
ammonium citrate solution for apatite deposits from three locations. Data was sourced from
Francisco et al. (2007) to create this figure. b) Phosphorus solubility in neutral ammonium
citrate aerobic thermochemical treatment of sewage sludge ash mixed with Na and K
sulphates. Data was sourced from Herzel et al. (2021) to create this figure
Conclusion

Multiple use cases for biochar have been identified, with high cost of production issues
preventing widespread adoption. Cost effective production is likely to be associated with
short pyrolysis time and a continuous flow design. Further, a pyrolysis system must be
designed to tolerate the high corrosion potential associated with acidic products and high
temperatures. The continuous flow, fluidised bed system of Winter (2018) was designed as
an attempt to address these issues.

Specific biochar properties are required for the various end use applications, including use
for soil amendment, heavy metal remediation and animal feed additive. Biochar can be
produced to optimise its value for a given application by selection of feedstock and pyrolysis
temperature and time. While trends are established, e.g., biochar CEC is decreased with
increasing pyrolysis temperature, the optimal feedstock and pyrolysis temperature/time
combination should be established for a given type of reactor, a given feedstock and a given
property/end use case.

36
CHAPTER 3. MATERIALS and METHODS
Introduction

Biochar is commonly characterised in terms of surface area, porosity, functional groups, bulk
density, pH, electrical conductivity (EC), water holding capacity (WHC), liming potential,
calorific value, cation exchange capacity, zeta potential, using proximate analyses, ultimate
analyses, elemental analyses, scanning electron microscopy (SEM) and Fourier Transform
InfraRed spectroscopy (FTIR) (Intani et al., 2018; Mitchell et al., 2013). The International
Biochar Initiative (IBI) and the European Biochar Certificate (EBC) have recommended
characterisation of biochar in terms of moisture, hydrogen, carbon, nitrogen, ash content,
pH/liming potential, and electrical conductivity, heavy metals, polychlorinated biphenyls, and
polyaromatic hydrocarbons when biochar is used as a soil amendment (International Biochar
Initiative, 2015; Schmidt et al., 2016). EBC also recommends the assessment of these
attributes for animal feed applications (Schmidt et al., 2016).

This chapter provides details of procedures followed to produce biochar and for the
characterisation of its physical and chemical properties. IBI and EBC guidelines and
methods detailed in the book “Biochar: A Guide To Analytical Methods” (Singh et al., 2017a)
have been followed.

Biochar production

Feedstock material

Biochar was produced from three feedstocks: macadamia nutshells, eucalyptus woodchips,
and hemp stems. The macadamia (Macadamia integrifolia) nutshell was obtained from
Macadamia Processing Co. Limited, NSW, the eucalyptus material was obtained from
woodchip of 12-year-old plantings of Eucalyptus dunii harvested in the Mallanganee area
and hemp stem of Cannabis sativa harvested for food grade hemp production from plantings
at Mara Seeds Pty. Ltd (Mallanganee, NSW). Raw apatite for fluidised bed was obtained
from the Goondicum mines near Monto, Queensland. Three subsamples (1 kg) of each
material were dried in a fan-forced oven at 105 ℃ for 24 hours to determine moisture
content. The biochar produced from the three feedstocks (macadamia nutshells, eucalyptus
wood chips and hemp stem) are referred to as nutshell, woodchip and stem, respectively,
throughout the thesis.

37
Pyrolysis plant

A fluidised bed pyrolysis unit operating in a batch mode was used to simulate continuous
pyrolysis plant operating conditions, as patented by Winter (2018) (Australian and NZ patent
2017272314) (Fig. 2, 3). The unit was situated at Mara Seeds, Mallanganee, NSW, Australia.
The fluid bed consists of a metallic cylinder (100 mm internal diameter, 600 mm high) with a
conical base. The fluid bed was externally heated using a naturally aspirated propane burner. A
steel marble (Ø 25 mm) was placed into the cylinder before partly filling with 3 kg of apatite
crushed to pass a 500 µm sieve, and before feedstock was introduced. Nitrogen gas was used to
fluidise the bed of apatite sand, with gases exiting from an aperture at the top of the vessel. A
superficial fluidising velocity of 0.5 m/s was maintained. Nitrogen flow was determined via a
rotameter. A long thermocouple, mounted on top of the vessel into the fluidised bed, was used to
measure and control temperature. The fluid bed vessel was mounted on trunnion bearing to allow
the final cooled bed material to be emptied into a steel bucket.

Flue

Reaction vessel

N2 supply

Figure 2. Fluidised bed batch pyrolyser (pilot plant) at Mara Seeds Soft Agriculture
Company, Mallanganee, NSW, Australia (Winter, 2018)

38
The experimental sketch of the fluidised bed reactor is presented as Figure 3.

Figure 3. Experimental setup sketch of fluidised bed pilot plant pyrolyser (with permission of
John Winter)
In operation, the fluidising compressed air and nitrogen supplies was set at 400 kPa gauge,
with the flow of gas to the bed monitored by a rotameter and regulated by a needle valve.
The typical gas flow was 1.5 Nm3/hour. Liquified Petroleum Gas (LPG) supply was set at 10-
50 kPa pressure. The back pressure at the bottom of the vessel was typically about 40 kPa.
The sound of the marble rattling on the reactor wall and warbling of the thermocouple shaft
was used to signal adequate fluidisation of the apatite bed. The burner flame was lit, and the
extraction fan was started, with the fluidised bed heated to the desired temperature
(approximately 15 min) before adding feedstock. During heat-up the bed was fluidised with
air and then changed over to nitrogen before adding the feedstock material.

For a typical run, the ingress lid was opened, and biomass samples of known weight were
introduced into the pre-heated reaction vessel. The ingress lid was then closed, and
thermocouple was inserted into the apatite bed and the flue pipe lowered onto the reactor
egress point. Proper operation of the unit was signalled by equal temperatures of the
fluidised bed and the flue pipe. Similarly, visual signs of proper operation such as visible
smoke, rattling sound of the steel ball in the bottom of the fluid bed were observed. The unit
was operated for the desired residence time, and then the burner turned off and contents
cooled using nitrogen at a down to 150 ℃. The fluid bed vessel was then inverted, allowing

39
extraction of contents through the egress aperture. Biochar was separated from apatite
using a 500 µm sieve. The biochar was purged with N2 gas for about 30 min to prevent
oxidation before weighing.

The yield (%) of biochar was calculated from the dry weight of feedstock (Wf) and weight of
biochar (Wb) as:

𝑊𝑊𝑊𝑊
𝑌𝑌𝑌𝑌𝑌𝑌𝑌𝑌𝑌𝑌 = × 100 % …….. Equation 1
𝑊𝑊𝑊𝑊

Biochar was produced from triplicate 500 g samples of each of the three feedstocks at each
of three temperatures (350, 450 and 600 ℃) and three residence times (5, 10 and 20 min),
i.e., 27 batch runs.

Biochar characterisation

Sample preparation

Biochar and raw apatite were then ground to pass a 2 mm sieve for analysis of most
characteristics. For proximate analysis and CN analysis the size was further reduced to pass
through a sieve of size 100 µm, while for zeta potential and FTIR, particle size was reduced
to < 75 µm.

The properties of biochar were characterised based on published methods (Table 4).
Analyses were primarily undertaken by the author at CQU, or under guidance at another
location (e.g., SEM) while more specialist analyses were undertaken by specialist
laboratories (Table 4).

40
Table 4. Standard methods for the analysis of physicochemical characteristics
Characteristics Measurement Technique / Test Method Citation Analysis 1

pH The 1:20 biochar: deionised water solution was measured by pH meter with glass- (Rajkovich et al., CQU
calomel electrodes, calibrated using pH 7 and pH 10 buffers (International Biochar 2011)
Initiative, 2015)
Moisture Mass loss in percentage at 105 ℃, (Enders & Lehmann, CQU
2017)
ASTM Proximate Analysis (Biochar: A Guide To Analytical Methods)
Ash Content Weight after combustion at 750 ℃, ASTM Proximate Analysis (Biochar: A Guide To (Aller et al., 2017) CQU
Analytical Methods)
Volatile Matter Mass loss in percentage in between 105 to 950 ℃, ASTM Proximate Analysis (Biochar: (Intani et al., 2018) CQU
A Guide To Analytical Methods)
Fixed Carbon ASTM Proximate Analysis (Biochar: A Guide To Analytical Methods) (Intani et al., 2018) CQU
Electrical Conductivity EC of 1:20 biochar: deionised water solution (International Biochar Initiative, 2015 (Singh et al., 2017b) CQU
Carbon and Nitrogen Ultimate (Elemental) Analysis (International Biochar Initiative, 2015 (Bird et al., 2017) CQU
Biochar Liming 19A1 method of rapid titration (Biochar: A Guide To Analytical Methods) (Rayment & CQU
Potential Higginson, 1992)
Bulk Density VDLUFA-Method A 13.2.1, European Biochar Certificate Guidelines (European Biochar (Schmidt et al., CQU
Certificate, 2019) 2016)
Surface Area CO2 Physisorption method (Sigmund et al., UNSW
2017)
Surface Morphology Scanning Electron Microscopy-Electron dispersive X-ray Spectroscopy (SEM-EDX) (Joseph et al., UNSW
2018a)

41
Calorific Value Using a bomb calorimeter apparatus (International Biochar Initiative, 2015) (Malucelli et al., CQU
2019)
Water Holding Capacity German Standard E DIN ISO 14238-2011 (European Biochar Certificate, 2019) (Schmidt et al., CQU
2016)
Hydrogen and sulphur Ultimate (Elemental) Analysis by Gas chromatography for Hydrogen and LECO TruMac (Bird et al., 2017) EAL
CNS determinator for Sulphur (International Biochar Initiative, 2015)
As, Al, B, Cd, Co, Cr, Inductively Coupled Plasma Mass Spectrometry (ICP-MS) (Das et al., 2018) UNSW /
Cu, Fe, Mn, Mo, Ni, Pb, EAL
Ti, Zn, Zr
Ca, K, Mg, Na, P, S, Si Inductively Coupled Plasma Mass Spectrometry (ICP-MS) / Inductively Coupled Plasma (Siddiqui et al., UNSW /
Optical Emission Spectrometry (ICP-OES) 2019) EAL
Extractable 2 % formic acid followed by spectrometry (International Biochar Initiative, 2015) (Arbestain et al., EAL
Phosphorus 2015)
Polycyclic Aromatic USEPA 8270 (2007) using Soxhlet extraction (USEPA 3540) and 100 % toluene as the (Hale et al., 2012) ALS
Hydrocarbons (PAHs) extracting solvent (International Biochar Initiative, 2015)
Polychlorinated USEPA 8082 (2007) (International Biochar Initiative, 2015) (Schmidt et al., ALS
Biphenyls (PCBs) 2016)
Functional Groups Fourier Transform Infra-Red (FTIR) Analysis (Biochar: A Guide To Analytical Methods) (Siddiqui et al., CQU
2019)
Zeta potential Zeta Potential Analyzer (Fahmi et al., 2018) CQU
Cation Exchange Modified AOAC method 973.09 (Huff & Lee, 2016) CQU
Capacity
1Central Queensland University (CQU), University of New South Wales (UNSW), University of Western Australia (UWA), and Environmental
Analysis Laboratory (EAL), Southern Cross University (SCU)

42
Bulk density

Bulk density (g/cm3 ) was determined from the mass and volume of the unground sample,
direct from the pyrolysis unit, following EBC guidelines (Schmidt et al., 2016). A graduated
cylinder of known volume was filled with biochar, with five taps to settle the sample. The
volume of biochar was then assessed.

pH and EC

For pH and EC measurements, 1 g of air-dried biochar sample (ground to < 2 mm) was
weighed into a 50 mL propylene tube, and 20 mL of deionised water was then added. The
tube was shaken for 90 min at 25 ℃ and allowed to stand for 30 min (Rajkovich et al., 2011).
The pH and EC of the suspension were measured with a PC2700 benchtop meter (Eutech
Instruments) calibrated using pH 7 and 10 buffers for pH measurements > 7, pH 4 and 7
buffers for pH measurements < 7, and a 2760 µS/cm ionic buffer.

Liming potential

A modified version of the 19A1 method was used to determine biochar liming potential (as
calcium carbonate equivalence) (Singh et al., 2017b). The 19A1 method is the rapid titration
method of measuring soil carbonate proposed by Rayment and Higginson (1992).

Air-dried sample ground to < 2 mm (0.5 g) was weighed into a 50 mL propylene tube and 10
ml of 1M HCl solution (41.5 ml of conc. 37 % HCl in 500 ml deionised water) was added,
then shaken at 25 ℃ for 2 hours and allowed to stand for 16 hours. The solution was then
titrated with standardised 0.5 M NaOH solution. NaOH solution was prepared with 20 g of
NaOH in 1000 ml deionised water and standardised with potassium hydrogen phthalate (0.8
g dissolved in deionised water with 2 – 5 drops of phenolphthalein indicator and titrated with
the 0.5 M NaOH). A blank titration was performed using 10 mL of standardised 1 M HCL and
a reference sample of 0.5 g CaCO3 powder (dried at 105 ℃ for 1 hour) was mixed with 10
mL of 1M HCL and titrated with standardised 0.5 M NaOH. Liming potential was calculated
as % CaCO3 equivalent as follows:

M × (b − a) × 10−3 × 100.09 × 100


% CaCO3 equivalent = …… Equation 2
2×W

where M is the standardised molarity of NaOH (mol L-1), b is the volume of NaOH consumed
(mL) by the blank, a is the volume of NaOH being consumed (mL) by the biochar sample
and W is the mass of biochar (around 0.5 g). The factor 10-3 is used to convert mL to L,
100.09 is the molar mass of CaCO3; 100 is a multiplier for obtaining % CaCO3 equivalent
and factor 2 is used as 1 mole of CaCO3 consumes 2 moles of H+.

43
Water holding capacity

Water Holding Capacity (WHC) was assessed using the modified German Standard E DIN
ISO 14238-2011 method which involves soaking and drying the sample (Schmidt et al.,
2016). A dried biochar sample (1 g) was weighed into a small plastic container with fine
mesh at the bottom and placed into a petri-dish of water. The sample was allowed to
saturate overnight and taken out and covered with parafilm to avoid moisture loss through
evaporation while allowed to drain for 8 hours. The biochar was removed from the container
and placed in a pre-weighed container (M1) and the total weight of moist biochar and
moisture container (M2) was recorded. The container was then placed in an oven for 12
hours at 105 ℃ and sample weight (M3) was recorded. The water holding capacity (% w/w)
was calculated as follows:

M2 − M3
WHC = × 100 …… Equation 3
M3 − M1

Proximate analysis

The American Society for Testing and Materials (ASTM) provides thermal decomposition
methods for the proximate analysis (moisture, volatile matter, ash, and fixed carbon
determination) in charcoal-like materials. The ASTM method was followed for the proximate
analysis of biochar as per IBI and EBC guidelines. Moisture, volatile matter, ash content and
fixed carbon were determined using the improved proximate analysis method of the
American Standard for Testing and Materials (ASTM) D1762-84 (Enders & Lehmann, 2017).

Moisture content

For moisture determination, 1 g of biochar was weighed in a crucible (the crucible was
previously dried at 105 ℃) and then heated in an oven for 18 hours at 105 ℃ before
reweighing. Moisture content (% w/w) was determined from sample mass (M) as received
and after drying 105 ℃, as:

𝑀𝑀 𝑎𝑎𝑎𝑎 𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 – 𝑀𝑀 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 𝑎𝑎𝑎𝑎 105℃


Moisture = × 100 ….. Equation 4
𝑀𝑀 𝑎𝑎𝑎𝑎 𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟

Volatile matter content

For volatile matter content, 1g of biochar previously dried at 105 ℃ was weighed into a
crucible. The crucible was covered and placed inside the muffle furnace at a temperature of
950 ℃ for exactly 10 min before extraction, left to cool in a desiccator at room temperature
for exactly 10 min and then reweighed. The crucibles were then transferred to desiccators.
The volatile matter (% w/w) was calculated by the formula:

44
M dried at 105℃ − M devolatilised at 950℃
𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 = × 100 ….. Equation 5
M dried at 105℃

Ash content

For ash content, 1g of biochar dried at 105 ℃ was weighed in a crucible. The crucible was
covered and placed inside the muffle furnace at a temperature of 750 ℃ for 6 hours and then
cooled to 105 ℃. The covered crucibles were then transferred to desiccators and weighed
again after cooling to ambient temperature. The ash content (% w/w) was calculated as:

Weight residue after 750℃


𝐴𝐴𝐴𝐴ℎ 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = × 100 ……. Equation 6
Weight dried at 105℃

Fixed carbon content

The fixed carbon content (% w/w) was obtained after the removal of volatile matter and ash
content from the biochar as presented in equation 5 and 6 calculated as:

𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = 100 – 𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 − 𝐴𝐴𝐴𝐴ℎ 𝑐𝑐𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜 ….. Equation 7

Calorific value

The calorific value of biochar was determined following the ASTM D5865-2010 bomb
calorimeter method using a 5E-C5500 Calorimeter (Changsha Kaiyuan Instruments Co.,
Ltd.). Approximately 1 g of biochar dried at 105 ℃ for 24 hours was weighed into a metal
cup, loaded into the sample holder and connected with the ignition wire. Then, deionised
water (1 mL) was added into the bomb vessel and the loaded sample was assembled into
the bomb, with the rim moistened with water. The bomb pressurised to 3 MPa and allowed to
stabilise for 15 seconds. After leak checking, the sample was loaded into the calorimeter and
the sample was ignited. The sample weight and the calorific value of the wire used were
recorded into the 5E open analysis software for the determination of calorific value.

Carbon and nitrogen content

A TruMac Series Carbon and Nitrogen Analyser (LECO, Sydney, Australia) was used to
determine the carbon and nitrogen content at CQUniversity laboratory, Rockhampton. Pre-
dried biochar (0.15 g) was weighed into ceramic boats for the CN measurement (Intani et al.,
2018). Ethylenediaminetetraacetic acid (EDTA) was used as a reference material.

Total elemental analyses

Total elements from three feedstocks (Macadamia nutshell, eucalyptus woodchip and hemp
stem) and apatite were extracted through nitric acid digestion and analysed using ICP-MS
techniques in Environmental Analysis Laboratory (EAL), Lismore, Australia.

45
Water extractable elemental analyses

Water-soluble elements were extracted from 1 g of ground biochar added to 10 mL of


distilled water, with the solution stirred at 50 ℃ for 24 hours. The solution was then filtered
through a 0.45 µm filter following sonication for 2 – 5 min. After filtration, the clear liquid was
held at 40 ℃ before analysis using ICP-MS and ICP-OES. Calcium (Ca), potassium (K),
magnesium (Mg), sodium (Na), silicon (Si), and phosphorus (P) were determined using ICP-
OES (in units of mg/L). Trace elements such as arsenic (As), aluminium (Al), boron (B),
cadmium (Cd), cobalt (Co), chromium (Cr), iron (Fe), manganese (Mn), molybdenum (Mb),
nickel (Ni), lead (Pb), tin (Ti), zinc (Zn) and zirconium (Zr) were estimated using ICP-MS (in
units of µg/L). ICP-MS and ICP-OES analyses were undertaken by Mark Wainwright
Analytical Centre, UNSW, Australia.

Available phosphorus

Available phosphorus from raw and roasted apatite was determined by extraction with 2 %
formic acid followed by spectrometry (International Biochar Initiative, 2015). The analysis
was performed by Environmental Analysis Laboratory (EAL), Lismore, Australia.

Cation exchange capacity

Cation exchange capacity (CEC) was determined using the Modified AOAC method 973.09
(Huff & Lee, 2016). The modified AOAC 973.09 refers to the modified version for the
measurement of total exchangeable cations from the Journal of Association of Official
Analytical Chemists (AOAC) as mentioned in Rippy and Nelson (2007).

Air-dried biochar (0.5 g) was weighed into a 125 ml Erlenmeyer flask and 0.5 ml of 0.5 M
HCl was added to displace the adsorbed cations and saturate biochar with hydrogen ions
(H+) and shaken for 2 hours at 110 rpm (Junior Orbit Shaker, Lab-line Instruments,
Designers and Manufacturers, Melrose Park, ILL). The sample was then filtered and rinsed
with 100 mL of deionised water to remove displaced cations. The samples were checked
with a few drops of dilute (0.028M) AgNO3 solution and re-rinsed if a cloudy white precipitate
formed. The filtrate was then discarded, and filter paper with the washed biochar was
transferred to a clean 125 mL Erlenmeyer flask and 50 mL of 0.5 M Ba (OAc)2 was added,
displacing H+. The samples were kept in a shaker for an hour, filtered, and washed three
more times with 100 mL of deionised water for the removal of displaced H+. The biochar was
then discarded, and the filtrate was titrated with 0.1 M NaOH with a few drops of
phenolphthalein. CEC was calculated by the equation as follows:

a × b ×100
CEC = …… Equation 8
𝑊𝑊

46
where CEC is cation exchange capacity in cmol/kg, a is the volume of 0.1 M NaOH
consumed, b is the molarity of NaOH (0.1 M) and W is mass of biochar (approx. 0.5 g).

Surface morphology by scanning electron microscopy-electron dispersive x-


ray spectroscopy (SEM-EDX)

Surface morphology and mineral distribution were observed using a Phenom Desktop
Scanning electron microscope (SEM) (Thermo Scientific, Sydney, Australia) fitted with a
Bruker energy dispersive x-ray (EDX) (Bruker, Billerica, MA) analyser following the protocol
of (Joseph & Munroe, 2017). Biochar pieces were cut/cracked with a small blade to create
cross-sections, which were then placed in graphite paint on an aluminium stub (three pieces
per stub). The stub was dried for 3 to 4 min using a heater. To make the sample electrically
conductive, the sample was sputter-coated with graphite. A backscattered electron picture
(operating at 15 kV) was generated to illustrate the distribution of mineral phases and their
interaction with biochar phases, then EDX analysis was carried out in a point-by-point
fashion, with the beam focused on a particular region of the samples and the spectrum
recorded. SEM images were recorded at magnifications ranging from 1350 to 7700 X.

Surface area and pore volume determination

The CO2 physisorption technique was used to determine the surface area of the biochar
samples (Sigmund et al., 2017) using a Tristar II plus analyser (Micromeritics, Acta Scientific
Instruments, Sydney, Australia) at 0 °C. The sample was pre-treated for 12 hours under
vacuum at a temperature of about 30 °C.

Zeta potential

No standard protocols were found for the determination of the zeta potential of biochar. The
procedure for sample preparation was adapted from Qi et al. (2017), Castilla-Caballero et al.
(2020), Hossain et al. (2011) and Hong et al. (2019). The use of a high ionic strength
medium is reported to result in the agglomeration of the particles in the suspension and
shielding of the surface charge of particles. However, when deionised/distilled water was
used as a dispersing medium, while particles did not agglomerate, they also did not stay in
suspension, and readings varied with time. Thus, the results did not meet measurement
quality criteria in terms of standard deviation of three repeated zeta potential measurements
and were considered “poor”. The use of 0.1 M NaCl, following the protocol of Fahmi et al.
(2018), resulted in an acceptable deviation of three repeated zeta potential measurement.
The use of dilute NaCl is suggested to maintain consistent conductivity (Kaszuba et al.,
2010).

47
The zeta potential value of the biochar was determined by weighing biochar oven-dried at 40
℃ (10 mg) into a conical flask containing 50 mL of 0.1 M NaCl solution. The suspension was
kept in a bath-type sonicator (300 W) set at 40 kHz for 120 min at 30 ℃ to disperse the
suspension and then left undisturbed for 24 hours. An aliquot of the suspension was
introduced into the disposable folded capillary cells (while taking care to avoid bubbles inside
the capillary cell when inserting the sample). The disposable folded capillary cells were
inserted inside Zeta Sizer Nano Series for zeta potential measurement. The measurement
was taken three times for each sample and the average value was calculated.

Polycyclic aromatic hydrocarbons and Polychlorinated biphenyls

The Polycyclic aromatic Hydrocarbons (PAHs) were determined following US EPA 8270
(2007) using Soxhlet extraction (US EPA 2540) and 100 % toluene as the extracting solvent.
The Polychlorinated Biphenyls (PCBs) were determined following US EPA 8082 (2007).

Surface functional groups by fourier transform infrared (FTIR) spectroscopy

Mid-infrared (MIR) spectra were collected between 4000 and 650 cm-1 using a Perkin Elmer
Spectrum 100 FTIR spectrometer (Perkin Elmer Ltd, United Kingdom) fitted with a composite
zinc selenium-diamond ATR accessory. Before each biochar sample, a background
measurement on air was performed. Biochar samples (ground to < 75 µm) were placed on
the platform and a uniform pressure was applied during collection of spectra. In between
each sample, the platform was cleaned with isopropyl alcohol and dried with laboratory
Kimwipes® (Kimberly Clark; Sydney, Australia).

MIR spectra between 4000 and 650 cm-1 were recorded at a resolution of 4 cm-1 using the
Spectrum IR software version 10.6.2.1159 (Perkin Elmer Ltd, United Kingdom) as the
average of 24 scans. Three replicate spectra were collected for each biochar sample.

Statistical analysis

Statistical analyses were carried out to determine the significance of temperature, time, and
feedstock on measured parameters using R version 4.1.0. (2021 v530). Differences between
means were calculated by analysis of variance (ANOVA) with temperature, time, and
feedstock as independent factors. Analysis of variance (ANOVA) performed by linear mixed
model effect using lme function of R package nlme R version 4.1.0. (2021 530). Significance
difference was determined using a two-way ANOVA test with post hoc Tukey HSD test with R
version 4.1.0. (2021 530). P-values less than 0.05 (p < 0.05) were considered statistically
significant. Correlation analysis between different attributes was determined with a “corrplot”
package of R version 4.1.3 (2022-03-10).

48
MIR spectra were analysed using The Unscrambler X Software Version 10.5 (Camo ASA,
Oslo, Norway). The data were arranged based on different temperatures and times during
pyrolysis in each feedstock biochar. The spectra were transformed using Multiplicative
Scatter Correction (MSC) to compensate for multiplicative effects in the spectral data.
Principal component analysis was performed on the MSC transformed spectral data using
Unscrambler X.

49
CHAPTER 4. RESULTS
Feedstock and apatite characteristics

Physical and chemical characters

The three feedstocks were chosen to represent biomass of a range of woodiness, and
consequently a range of attributes varied between the feedstocks (Fig. 4, Table 5). Density
varied from 0.50 g/cm3 for macadamia nutshells to 0.13 g/cm3 for hemp stem. The pH and
EC varied from 5.7 and 2.5 mS/cm in hemp stem to 4.3 and 0.1 mS/cm in eucalyptus
woodchips. Similarly, liming potential and WHC were highest (76 % and 367 % w/w) in hemp
stem and least (35.8 % and 47.2 % w/w, respectively) in macadamia nutshells. Moisture (9.6
% w/w) and ash content (5.0 % w/w) was highest in hemp stem, while the highest volatile
matter content (86.6 % w/w) was in eucalyptus woodchips and highest fixed carbon (19.0 %
w/w) in macadamia nutshells. The highest calorific value (20.7 MJ/kg) was in macadamia
nutshells and the lowest (18.2 MJ/kg) in hemp stem. Carbon content (52.8 % w/w) was
highest in macadamia nutshells, while nitrogen content was highest (0.86 %) in hemp stem,
while hydrogen content (6.9 %) was highest in eucalyptus woodchips.

Figure 4. Feedstock materials (macadamia nutshell, eucalyptus woodchip and hemp stem)
and fluidised bed material (ground apatite). Scale bar is 10 mm

50
Table 5. Chemical characteristics of three feedstocks. Weight units are based on weight
after oven drying at 105 ℃
Feedstocks Macadamia Eucalyptus Hemp Apatite
nutshells woodchips stem
Moisture (%) 5.31 8.1 9.6 0.5
Density (g/cm3) 0.5 0.3 0.1 1.8
pH 4.7 4.3 5.7 7.7
EC (mS/cm) 0.1 0.1 2.5 0.0
Liming Potential (% w/w) 35.8 37.1 75.5 34.4
Ash Content (% w/w) 1.1 1.0 5.0 99.4
Volatile matter (% w/w) 79.9 86.6 77.7 0.5
Fixed Carbon (% w/w) 19.0 12.4 17.3 0.1
Calorific Value (MJ/kg) 20.7 19.3 18.1 -0.2
Carbon (% w/w) 52.8 49.6 45.6 0.1
Nitrogen (% w/w) 0.3 0.1 0.9 0.0
Water Holding Capacity (% w/w) 47.2 76.4 367.3 22.7
Hydrogen (% w/w) 6.7 6.9 6.6 NA
CEC (cmol/kg) 31.8 39.5 39.9 23.7
Zeta potential (mV) -23.5 -20.9 -14.1 -26.8
Note: NA denotes data not available.

The elemental constitution of the three feedstocks and apatite is presented in Table 6.
Distinctly, apatite was found to have the highest Ca, P, S, Al, Fe and Mn as compared to the
feedstocks. Among the feedstocks, stem had the highest Ca, K, Mg, P, Si, Fe and Zn.

Table 6. Total elemental constituents (acid-digested) the feedstocks and apatite. Units are
presented in terms of weight of oven (40 ℃ for 24 – 48 h) dried material
Total Macadamia Eucalyptus Hemp Apatite
elements nutshells woodchips stem
ICP-MS Ca 424 1112 5324 298225
(mg/kg)
K 1365 100 14758 189

Mg 165 140 1436.4 3362.5

Na 36.5 20.3 798.9 622.2

P 67.3 119 1248 127742

S 478 284 1175 1058

Si 273 339 452.3 587

As <0.1 <0.1 <0.1 9.23

51
Al 13.1 82.4 63 3583

B 3.4 1.7 13.6 0.8

Cd <0.1 <0.1 <0.1 <0.1

Co <0.1 0.2 0.1 2.8

Cr 0.2 1.2 0.5 0.5

Cu 5.4 1.1 3.2 1.8

Fe 44.2 143 160 9061

Mn 50.1 24.8 29.7 362

Mo 0.1 <0.1 0.2 <0.1

Ni 0.3 0.8 0.4 0.7

Pb <0.1 0.2 <0.1 0.54

Ba 1.6 3.8 7.8 26.6

Zn 3.8 6.9 26.0 3.7

Ag <1 <1 <1 <1

Se <0.5 <0.5 <0.5 4.1

Hg <0.1 <0.1 <0.1 <0.1

Apatite was notable for a high pH1:20 and high liming potential (Table 5). As expected for a
mineral, ash content was very high while carbon and nitrogen content and calorific value was
negligible (Table 5).

As seen in SEM images, apatite ground to pass a 500 µm sieve produced particles that were
up to 0.5 mm in length (Fig. 5). The nutshell feedstock was characterised by thick walled,
isodiametric sclereid cells, while secondary cell walls characteristic of vascular tissue
dominated woodchip. A mix of parenchymatous and vascular tissue was observed in stem.

52
Figure 5. Backscattered electron SEM images of feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) and apatite. Scale bar is 100 µm
Apatite was characterised by the presence of P, Ca, Si, Fe, Al, and Mg in the EDX spectrum
(Fig. 6) which consistent with the total elemental analysis as presented in Table 6. EDX of
stem biomass detected Cl, K and Ca, while traces of K were detected in macadamia (data
not shown).

Figure 6. Representative EDX analysis of apatite

53
FTIR

The FTIR absorbance spectra (Fig. 7) of apatite was dominated by a feature, a wavenumber
of 1012 cm -1, while spectra of nutshells, woodchips and stem were dominated by a major
peak at 1030 cm -1 and minor peaks at 3328, 2920, and 1230 cm -1. In addition, the spectra
of stem featured a minor peak at 1614 cm -1.

Figure 7. Mid-infrared spectra of apatite, macadamia nutshell, eucalyptus woodchip and


hemp stem feedstocks
A distinct separation of the feedstocks was achieved in a Principal Component Analysis,
principally along PC1 (Fig. 8) which explained 100 % of the variance in spectra. Hemp
feedstock was well separated from both macadamia and eucalyptus feedstocks in both PC 1
and 2.

54
Figure 8. PCA of macadamia, eucalyptus and hemp feedstocks (Feedstock samples were
oven dried at 105℃ for 24 hours and ground to size <75 µm)

Apatite after pyrolysis

The raw apatite contained 127,742 mg P. kg-1 (12.8 %) on a dry weight basis, with 9.5 % of
this being 2 % formic acid extractable (Table 7). Total P increased to slightly over 150,000
mg P. kg-1 following pyrolysis at any of the nine duration-temperature combinations, a result
partly explained by loss of moisture and volatiles content of the raw apatite. A decrease in
extractable phosphorus was noted with pyrolysis duration (at 350 oC) and temperature
(Table 7).

Table 7. Change in phosphorus availability in 2 % v/v formic acid solution with increasing
pyrolysis temperature. Units are presented in terms of weight of oven (40 ℃ for 24 – 48 h)
dried material. There were two replicates per treatment, with a maximum variation between
replicates of 5% of the average value. The % of total P and Extractable P are calculated for
raw apatite.
Treatment Total % of P Extractable % of Extractable
phosphorus P extractable P as % of
(P) as of total P
mg/kg extractable
(mg/kg) raw apatite
Raw apatite 127742 12131 9.5
350 ℃ for 5 min 153369 120 10207 84 6.7
350 ℃ for 20 151257 118 6846 56 4.5
min
450 ℃ for 5 min 152209 119 2159 18 1.4

55
450 ℃ for 20 154939 121 2221 18 1.4
min
600 ℃ for 5 min 155327 122 2423 20 1.6
600 ℃ for 20 151930 119 2212 18 1.5
min

The concentration of other elements such as Ca, Mg, Na, Si, Al, Fe, Mn, etc also increased
in roasted apatite as compared to the raw apatite (Table 8). However, increasing pyrolysis
temperature did not affect elemental concentrations (Table 8).

Table 8. Total elemental analysis of apatite with temperature and residence time variation.
Units are presented in terms of weight of oven dried material (40 ℃ for 24 – 48 h)
Elemental Roasted apatite
concentration
(mg/kg) 350 ℃ 450 ℃ 600 ℃
5 min 20 min 5 min 20 min 5 min 20 min

Ca 301176 292406 292405 299914 301399 290802


K 186 193 174 187 194 189
Mg 3808 3799 3269 3184 3576 3288
Na 894 939 951 996 1073 1033
P 153369 151258 152209 154940 155327 151931
S 96 94 71 94 81 84
Si 2266 2402 2378 2454 2565 2495
As 11 12 11 11 12 11
Al 4911 5155 4957 5040 5429 5233
B 1 0 0 1 1 1
Cd 0 0 0 0 0 0
Co 4 4 3 3 4 3
Cr 1 1 1 1 1 1
Cu 3 3 2 2 2 2
Fe 10315 10435 8419 8262 9310 8491
Mn 417 412 393 394 440 388
Mo <0.5 <0.5 <0.5 <0.5 <0.5 <0.5
Ni 1 1 1 1 1 1
Pb 1 1 1 1 1 1
Zn 5 5 5 4 5 5

56
Biochar characteristics

Morphology

The biochars produced from the three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) were grossly different in morphological character, with minor
differences in coloration associated with pyrolysis temperatures (350, 450 and 600 ℃).
There was no visible difference between material of different pyrolysis times (5, 10 and 20
min) for a given pyrolysis temperature (Fig. 9).

57
Figure 9. Images biochar produced using pyrolysis duration and temperatures (a)
macadamia nutshell (b) eucalyptus woodchip and (c) hemp stem. Scale bar is 10 mm
Cell wall morphology of the feedstocks was preserved in the pyrolyzed material, as
visualised in SEM secondary electron images of biochar samples (Fig. 10). Nutshell biochar
was characterised by the presence of thick walled sclereid cells with pores of approximately
1 mm diameter. Vessel elements with scalariform thickening and interspersed
parenchymatous ray cells were present in woodchip biochar. Stem biochar differed in being

58
dominated by parenchyma cells with some tracheary elements present. Apatite fragments,
as identified by shape, backscattered electron image intensity and EDX Ca content (data not
shown), were present across biochar surfaces.

59
Figure 10. SEM images of macadamia nutshell, eucalypt woodchip and hemp stem biochar produced at 350, 450 and 600 ℃. Differences in
brightness are ascribed to differences in charging between samples. Red arrows point to apatite fragments. Blue arrows highlight pores on
sclereid cell walls. Scale bar is 50 µm

60
Biochar yield

Feedstock and temperature had a highly significant (p < 0.0001) impact on biochar yield
(Table 9, Fig. 11). For example, yield of woodchip decreased significantly from 36 to 15 %
(w/w) for pyrolysis temperatures of 350 and 600 ℃, respectively.

Table 9. ANOVA significance testing of impact of feedstock, pyrolysis temperature and time
on biochar yields

Nutshell Woodchip Stem


p-value p-value p-value

Feedstock <0.0001*** <0.0001*** <0.0001***


Temperature <0.0001*** <0.0001*** 0.0001**
Time 0.1807ns 0.0019* 0.1976ns
Temperature* Time 0.3041ns 0.0133* 0.4437ns
Feedstock x Temperature <0.0001*** <0.0001*** <0.0001***
Feedstock x Time 0.9872 0.9872 0.9872
Feed x temp x time 0.1995ns 0.1995ns 0.1995ns

Figure 11. Yield (% w/w) of biochar from three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) for three pyrolysis temperatures and three residence times.
Different letters signify statistical difference at P < 0.05 within a feedstock. Bars represent
means (n=3) with associated SE
Biochar yield decreased across the feedstocks of nutshell, woodchip, and stem. Residence
time significantly impacted yield for woodchip but not nutshell and stem feedstocks. For
example, for woodchip biochar at 450 ℃, an increase in residence time from 5 to 20 min
resulted in a yield decrease from 27 to 23 %.

61
Bulk density

Bulk density of biochar varied from as low as 0.09 g/cm3 in stem biochar up to as high as
0.37 g/cm3 in nutshell biochar (Fig. 12). Feedstock density was a highly significant influence
(p < 0.05) to biochar bulk density. Density was significantly higher for nutshell and woodchip
biochar produced at 600 ℃ relative to biochar produced at 350 ℃ but no temperature effect
was observed for stem biochar.

Figure 12. Bulk density of biochar from three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) for three pyrolysis temperatures and three residence times.
Different letters signify statistical difference at P < 0.05 within a feedstock. Bars represent
means (n=3) with associated SE
pH

Biochar pH varied from 5.7 in woodchip biochar to 10.5 in stem biochar (Fig. 13), with a
significant trend (p < 0.05) to increased pH with increasing temperature in all feedstocks.
Residence time had no significant influence (p > 0.05) on the pH of stem and nutshell
biochar, but woodchip biochar pH was significantly increased with increased residence time.

62
Figure 13. pH of biochar from three feedstocks (macadamia nutshell, eucalyptus woodchip
and hemp stem) for three pyrolysis temperatures and three residence times. Different letters
signify statistical difference at p < 0.05 within a feedstock. Bars represent means (n=3) with
associated SE
Electrical conductivity

Electrical conductivity was significantly related to feedstock, varying from 0.01 mS/cm in
nutshell biochar to 4.30 mS/cm in stem biochar (Fig. 14). Biochar EC was positively
correlated (p < 0.05) with temperature. Residence time had a negligible effect (p > 0.05) on
biochar EC.

Figure 14. EC of biochar from three feedstocks (macadamia nutshell, eucalyptus woodchip
and hemp stem) for three pyrolysis temperatures and three residence times. Different letters
signify statistical difference at p < 0.05 within a feedstock. Bars represent means (n=3) with
associated SE

63
Liming potential

Feedstocks had a significant effect on the liming potential percentage of different biochar,
with values ranging from 30 % for woodchip biochar to 53 % for stem biochar (Fig. 15).
Temperature was found to be significant (p < 0.001) in determining the liming potential of
nutshell but was non-significant for woodchip or stem biochar. Residence time had no
significance (p > 0.05) to the liming potential of the biochar of any feedstock.

Figure 15. Liming potential of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Water holding capacity

Water holding capacity varied significantly (p < 0.05) between the biochar of the three
feedstocks, ranging from 14.0 % in nutshell biochar to 58.3 % in stem biochar (Fig. 16).
There was a significant effect (p < 0.05) of pyrolysis temperature on biochar water holding
capacity, but residence time had little effect (p > 0.05).

64
Figure 16. Water holding capacity of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Moisture

Moisture content varied from 1.2 % w/w in nutshell biochar to 3.8 % w/w in stem biochar
(Fig. 17). Feedstock had a significant effect (p < 0.05) on moisture content. Pyrolysis
temperature had no significant effect (p > 0.05) on the moisture content of stem biochar,
however, nutshell and woodchip biochar moisture content increased with the increasing
temperature (p < 0.05). Residence time had no significance (p > 0.05) on the moisture
content in all three biochars.

Figure 17. The moisture content of macadamia nutshell, eucalyptus woodchip and hemp
stem biochar produced at three pyrolysis temperatures and three residence times. Data
presented as means with associated SE (n=3). Different letters within a graph denote
significance difference (P < 0.05)

65
Volatile matter

Volatile matter ranged from 16.2 to 45.9 % w/w, with significant (p < 0.05) impact from
feedstock and temperature, but not residence time (Fig.18).

Figure 18. Volatile matter content of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Ash

The ash content of biochar ranged from 4.2 to 24.5 % w/w (Fig. 19) with significant (p <
0.05) impact from feedstock and temperature, but not residence time.

Figure 19. Ash content of biochar from three feedstocks (macadamia nutshell, eucalyptus
woodchip and hemp stem) for three pyrolysis temperatures and three residence times.
Different letters signify statistical difference at P < 0.05 within a feedstock. Bars represent
means (n=3) with associated SE

66
Fixed Carbon

Fixed carbon ranged from 43.8 % w/w in stem biochar to 68.6 % w/w in woodchip biochar,
with significant (p < 0.05) impact from feedstock and increasing with pyrolysis temperature (p
< 0.05) in woodchip biochar but not nutshell or stem. Residence time had no significant
effect (p > 0.05) on the fixed carbon content (Fig. 20).

Figure 20. The fixed carbon content of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Calorific value

Calorific value varied from 22.6 MJ/kg in stem biochar to 32.3 MJ/kg in nutshell biochar (Fig.
21), with pyrolysis temperature significantly impacting (p < 0.05) values for woodchip and
nutshell biochar, but not (p > 0.05) stem biochar. Residence time significantly impacted (p <
0.05) only woodchip biochar.

67
Figure 21. Calorific values of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Carbon

Carbon content of biochar varied significantly (p < 0.05) with feedstock, ranging from 57.2 %
w/w in stem biochar to 73 % w/w in nutshell biochar (Fig. 22). Carbon content increased
significantly (p < 0.05) with pyrolysis temperature of woodchip biochar but not nutshell and
stem biochar.

Figure 22. Carbon content (%) of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Nitrogen

Biochar nitrogen content varied by feedstock (p < 0.05), ranging from 0.20 % in woodchip
biochar to 1.65 % in stem biochar (Fig. 23). Pyrolysis temperature significantly (p > 0.05)
impacted nitrogen content of stem but not nutshell or woodchip biochar. Residence time had
no significant (p > 0.05) effect for any feedstock or temperature.

68
Figure 23. Nitrogen content (%) of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at P < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Hydrogen

Hydrogen content was highest (5 % w/w) in nutshell biochar and lowest (2 %) in stem
biochar (Fig. 24) and was decreased with pyrolysis temperature for all feedstocks.
Residence time had no significant effect on hydrogen content.

Figure 24. Hydrogen content (%) of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and two residence
times
Sulphur

Sulphur content ranged from 0.03 % w/w in woodchip biochar to 0.26 % w/w in stem biochar,
decreasing with increasing temperature in stem but not nutshell and woodchip (Fig. 25).

69
Figure 25. Sulphur content (%) of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and two residence
times
Total and water extractable elements

Total and water extractable elements were higher in biochar derived from stem than other
feedstocks, except for calcium which was higher in woodchip biochar in soluble form. In
comparison to the feedstock, total elements increased in all the biochar produced at 600 ℃
while, water extractable elements tended to decrease with increasing temperature. Calcium,
potassium, phosphorus, magnesium and sodium were among the highly concentrated
elements in total while potassium, boron, iron, manganese, and zinc were higher water-
soluble elements (Table 10 and 11) in the stem biochar owing to the higher concentration in
the hemp stem.

Table 10. Total elemental constituents (acid-digested) of biochar produced at 600 ℃ for 20
min. Units are presented in terms of weight of oven (40 ℃ for 24 – 48 h) dried material
ICPMS (mg/kg) (600 ℃ for 20 min)
Total Elements Nutshell Woodchip Stem
Ag <1 <1 <1
As <2 <2 <2
Pb <0.5 <0.5 <0.5
Cd <0.1 <0.1 <0.1
Cr 28.2 53.6 34.4
Cu 2.12 1.38 8.44
Mn 6.55 32.5 94.2
Ni 13.3 24.1 16.3

70
Se <2 <2 <2
Zn 1.59 2.26 180
Hg <0.1 <0.1 <0.1
Fe 268 828 2,883
Al 50.5 521 885
B <2 <2 15.2
Si 13.0 146 161
V <1 1.23 2.75
Co 0.40 0.88 1.16
Mo <0.5 <0.5 <0.5
Ba 1.12 3.12 22.3
Ca 1,458 19,264 31,200
Mg 17.9 118 1,990
K 841 79.0 20,408
Na 23.7 81.8 1,109
S 16.9 23.0 421
P 709 8,061 11,955

71
Table 11. Water extractable elemental content of biochar from different feedstock produced at varying pyrolysis time and temperatures
Elemental Woodchip Nutshell Stem
Analyses
350 ℃ 450 ℃ 600 ℃ 350 ℃ 450 ℃ 600 ℃ 350 ℃ 450 ℃ 600 ℃

10 20 10 20 10 20 10 20 10 20 10 20 min 10 20 10 20 10 20
min min min min min min min min min min min min min min min min min

ICP- Ca 0.14 0.17 0.39 0.27 0.22 0.22 0.02 0.01 0.01 0.01 0.00 0.00 0.20 0.13 0.03 0.09 0.10 0.08
OES
(mg/kg) K 0.11 0.15 0.02 0.03 0.18 0.22 0.26 0.27 0.83 0.80 1.31 1.37 18.12 19.06 13.97 18.07 28.37 18.99

Mg 0.04 0.04 0.03 0.02 0.01 0.01 0.01 0.00 0.00 0.00 0.00 0.00 0.89 0.78 0.16 0.15 0.03 0.02

Na 0.04 0.04 0.02 0.02 0.03 0.04 0.02 0.02 0.03 0.03 0.03 0.03 1.24 1.27 0.98 1.16 1.34 0.79

P 0.02 0.03 0.01 0.02 0.00 0.00 0.01 0.01 0.01 0.01 0.01 0.01 0.75 0.95 0.13 0.78 0.09 0.07

S 0.01 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 1.57 1.66 1.25 0.97 0.35 0.20

Si 0.01 0.02 0.02 0.01 0.01 0.01 0.00 0.00 0.00 0.00 0.01 0.01 0.66 0.73 0.29 0.52 0.10 0.10

ICP-MS As 0.23 0.35 0.36 0.35 0.26 0.29 0.29 0.26 0.40 0.26 0.16 0.29 0.56 0.49 0.22 0.29 0.36 0.13
(ug/kg)
Al 0.68 0.45 0.30 0.42 2.24 2.07 1.25 0.65 0.91 0.91 4.40 4.70 0.46 0.41 0.09 0.48 0.38 0.24

B 3.04 2.22 2.39 2.29 2.11 1.90 1.90 1.77 4.08 3.65 4.44 5.78 8.53 13.24 9.30 16.22 8.97 6.57

Cd 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

Co 0.03 0.04 0.01 0.01 0.00 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.04 0.06 0.06 0.03 0.00 0.00

Cr 0.01 0.06 0.01 0.01 0.00 0.00 0.01 0.00 0.00 0.02 0.01 0.03 0.01 0.02 0.01 0.01 0.00 0.01

Cu 0.13 0.04 0.01 0.04 0.02 0.02 0.16 0.05 0.01 0.02 0.00 0.02 0.20 0.28 0.05 0.16 0.02 0.01

Fe 0.23 0.20 0.07 0.06 0.06 0.09 0.61 0.21 0.06 0.15 0.52 0.67 6.24 6.92 0.82 4.48 0.12 0.04

Mn 3.79 4.28 2.62 1.95 0.26 0.27 0.89 0.55 0.26 0.16 0.04 0.05 1.44 1.07 0.19 0.40 0.01 0.00

Mb 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.00 0.00 0.01 0.01 0.01 0.37 0.48 0.34 0.48 0.15 0.08

Ni 0.01 0.03 0.01 0.01 0.01 0.01 0.01 0.01 0.00 0.01 0.00 0.02 0.06 0.04 0.04 0.02 0.00 0.00

72
Pb 0.01 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.01 0.00 0.00

Ti 0.09 0.10 0.04 0.06 0.02 0.01 0.02 0.02 0.03 0.04 0.03 0.05 0.31 0.27 0.21 0.23 0.17 0.14

Zn 2.02 2.06 0.42 0.79 0.31 0.18 2.33 0.69 0.19 0.73 0.02 0.06 3.25 2.30 2.59 0.82 0.05 0.05

Zr 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.02 0.02 0.02 0.02 0.00 0.00

Comparatively, the total elements increased in biochar than feedstocks with exception in K, Mg, Na, S and Si. Similarly, most of the elements
were less available in water for both biochar and feedstock expect for K and Si which were more water soluble in biochar.

Table 12. Evaluation of the water and total elemental analysis of feedstocks with biochar at 600 ℃ and 20 min for nutshell, woodchip and stem
Elements Feedstock water ex/total% Biochar water extract / total Biochar total/ Feedstock Biochar water extract% as a
(mg/kg) biochar % total% % of feedstock water
extract%
Nutshell Woodchip Stem Nutshell Woodchip Stem Nutshell Woodchip Stem Nutshell Woodchip Stem
Ca 0.0 0.0 0.0 0.0 0.0 0.0 91.4 297.3 79.4 4.6 8.6 0.8

K 0.1 0.1 0.1 0.2 0.3 0.1 16.4 13.5 18.7 307 299 112.6
Mg 0.0 0.0 0.1 0.0 0.0 0.0 2.9 14.5 18.8 17.7 18.4 1.2
Na 0.1 0.2 0.1 0.1 0.0 0.1 17.3 69.1 18.8 150 30.7 76.3
P 0.1 0.0 0.1 0.0 0.0 0.0 280 1160 129.9 2.2 0.1 0.4
S 0.0 0.0 0.0 0.0 0.1 0.0 0.9 1.4 4.9 272.8 715 116.2
Si 0.0 0.0 0.1 0.1 0.0 0.1 1.3 7.4 4.8 4593 180 99.1
As 0.3 0.6 1.0 0.0 0.0 0.0 531 343 271.1 4.2 2.4 0.7
Al 0.0 0.0 0.0 0.0 0.0 0.0 102 108 190 28.2 3.8 1.1
B 0.1 0.2 0.1 0.3 0.1 0.0 15.5 20.4 15.2 240 45.1 47.3
Cd 0.0 0.0 0.0 0.0 0.0 0.0 26.6 17.2 13.6 90.6 12.4 12.5

73
Co 0.0 0.0 0.1 0.0 0.0 0.0 105 62.4 125.1 8.1 1.7 0.1
Cr 0.0 0.0 0.0 0.0 0.0 0.0 3435 789 876 2.1 0.2 0.2
Cu 0.0 0.0 0.0 0.0 0.0 0.0 10.4 22.5 35.5 7.0 5.7 0.5
Fe 0.0 0.0 0.0 0.0 0.0 0.0 161.3 99.3 243.7 30.2 0.4 0.0
Mn 0.0 0.0 0.1 0.0 0.0 0.0 3.5 22.5 43.0 9.2 2.5 0.0
Mo 0.0 0.0 0.0 0.0 0.0 0.0 121 85.9 27.5 13.4 15.3 84.9
Ni 0.0 0.0 0.1 0.0 0.0 0.0 1289 525 522 1.0 0.6 0.0
Pb 0.0 0.0 0.0 0.0 0.0 0.0 133 34.7 67.8 12.4 2.3 1.4
Zn 0.1 0.0 0.1 0.0 0.0 0.0 11.2 5.6 94.0 5.3 26.2 0.0

74
Cation exchange capacity

The cation exchange capacity of biochar varied significantly (p < 0.05) with feedstocks,
ranging from 12.0 cmol/kg in nutshell biochar to 59.5 cmol/kg in woodchip biochar (Fig. 26).
CEC of biochar derived from woodchip, but not stem or nutshell, decreased significantly (p <
0.05) with pyrolysis temperature. Residence time had no significant effect (p > 0.05) on
biochar CEC.

Figure 26. The cation exchange capacity of biochar from three feedstocks (macadamia
nutshell, eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three
residence times. Different letters signify statistical difference at P < 0.05 within a feedstock.
Bars represent means (n=3) with associated SE
Zeta potential

Zeta potential values varied significantly (p < 0.05) among feedstocks, ranging from - 37.0
mv in stem to - 17.9 mv in nutshell (Fig. 27). Zeta potential significantly (p < 0.01) decreased
with pyrolysis temperature for nutshell, however zeta potential tended to increase with
temperature for stem. Residence time had no significant effect (p > 0.05) on the zeta
potential of biochar created of any feedstock-temperature combination.

75
Figure 27. Zeta potential values of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and three residence
times. Different letters signify statistical difference at p < 0.05 within a feedstock. Bars
represent means (n=3) with associated SE
Surface area

Surface area varied with feedstock and increased with pyrolysis temperature, with a minor
influence from duration. The highest surface area was found in nutshell biochar (151 - 535
m2/g), followed by woodchip biochar (137 - 368 m2/g) and stem biochar (141 - 306 m2/g)
(Fig. 28).

Figure 28. The surface area of biochar from three feedstocks (macadamia nutshell,
eucalyptus woodchip and hemp stem) for three pyrolysis temperatures and two residence
times (n = 1)

76
Total polycyclic aromatic hydrocarbons (PAHs) and polychlorinated biphenyls
(PCBs)

Total PAHs varied from < 1.5 to < 7.5 mg/kg, being highest in nutshell and woodchip biochar
at 450 and 600 ℃, while stem biochar had the lowest level, at <1.5 mg/kg (Table 13).
Similarly total PCBs varied from <0.4 to <0.9 mg/kg being highest in nutshell biochar at
450℃ for 20 minutes and 600℃, while the stem and woodchip biochar had the lowest level
of <0.4 mg/kg. Total PAHs increased with increasing temperature in nutshell and woodchip,
but not stem biochar while total PCBs increased with increasing temperature in nutshell
biochar only. Residence time did not impact total PAH and PCB levels.

Table 13. PAHs and PCBs of biochar from three feedstocks produced at three pyrolysis
temperatures and three durations
Feedstock Temperature Time Total Total
(℃) Polycyclic Polychlorinated
(minutes) aromatic biphenyls
hydrocarbons (mg/kg)
(mg/kg)
Macadamia nutshell 350 5 <1.5 <0.4
Macadamia nutshell 350 10 <1.5 <0.4
Macadamia nutshell 350 20 <1.5 <0.4
Macadamia nutshell 450 5 <7.5 <0.4
Macadamia nutshell 450 10 <7.5 <0.4
Macadamia nutshell 450 20 <7.5 <0.9
Macadamia nutshell 600 5 <7.5 <0.9
Macadamia nutshell 600 10 <7.5 <0.9
Macadamia nutshell 600 20 <1.5 <0.9
Eucalyptus woodchip 350 5 <1.5 <0.4
Eucalyptus woodchip 350 10 <1.5 <0.4
Eucalyptus woodchip 350 20 <1.5 <0.4
Eucalyptus woodchip 450 5 <7.5 <0.4
Eucalyptus woodchip 450 10 <7.5 <0.4
Eucalyptus woodchip 450 20 <7.5 <0.4
Eucalyptus woodchip 600 5 <7.5 <0.4
Eucalyptus woodchip 600 10 <7.5 <0.4
Eucalyptus woodchip 600 20 <7.5 <0.4
Hemp stem 350 5 <1.5 <0.4
Hemp stem 350 10 <1.5 <0.4
Hemp stem 350 20 <1.5 <0.9
Hemp stem 450 5 <1.5 <0.9
Hemp stem 450 10 <1.5 <0.4
Hemp stem 450 20 <1.5 <0.4
Hemp stem 600 5 <1.5 <0.4
Hemp stem 600 10 <1.5 <0.4
Hemp stem 600 20 <1.5 <0.4

77
FTIR spectroscopy

Absorbance spectra

The FTIR spectra of biochar from the three feedstocks contained consistent features (Table
14, Fig. 29). However, the overall absorbance spectrum was notably different for each
feedstock (Fig. 29). Similar trends were present in spectra of biochar from all feedstocks as
pyrolysis temperature increased, while no change was evident with change in duration of
pyrolysis between 5 and 20 min (Table 14, Fig. 29). The spectral features were consistent
with vibrational absorbances associated with the O-H, C-H, COO-, C=C, and C=O bonds
(Johnson et al., 2020) and the decrease in most of these features with increasing pyrolysis
temperature was consistent with the loss of functional groups. The exception to this
generalisation was an increase in absorbance peaks around 2980 – 2970 cm-1 and 950 –
745 cm-1, in biochar of all feedstocks.

Table 14. Features of mid-infrared spectra of biochar of three feedstocks with indication of
the trend in size of each feature with increasing pyrolysis temperature. Peak wavenumber
for each peak is given for biochar for each feedstock in the order nutshell, woodchip and
stem. ‘-‘ indicates that the peak was not present
Wavenumber Nutshell Woodchip Stem Assignment Origin
3460/3333/3336 decrease decrease decrease O-H stretching cellulosic materials and water
2975/2980/2980 increase increase increase C-H stretching asymmetric aliphatic bonds
2928/2928/- decrease decrease - C-H stretching asymmetric aliphatic bonds
1720/1699/1699 decrease decrease - C=O stretching carboxylic acid, amides, esters,
and ketones
1584/1584/1583 decrease decrease decrease COO- lignin components
1450/1450/1395 decrease decrease increase C=C stretching carboxylate anions and amides
1201/1208/- decrease, decrease - C-OH stretch, carboxylic acid, ester/amide
shift region
O-H deformation
-/-/1127 - - decrease C-O carbohydrate region
1030/1090/950 decrease decrease decrease Si-O, O-H minerals associated with biochar
stretching
870/870/873 increase increase increase C-H deformation aromatic carbon core
801/810/816 increase increase increase C-H deformation aromatic carbon core
745/745/- increase increase - C-H deformation aromatic carbon core

78
Figure 29. Mid-infrared spectra of biochar produced using three feedstocks at all
combinations of three pyrolysis temperatures and three pyrolysis durations. Each line is an
average of three spectra per condition

79
PCA

Principal component analyses on spectra of biochar of each feedstock demonstrated


separation of the biochar on pyrolysis temperature (350, 450 and 600 ℃) in scores of the
first principal component (PC1), with no separation based on pyrolysis duration (Fig. 30).
The stem biochar demonstrated the most variation between samples for a given pyrolysis
temperature.

80
Figure 30. Scores biplot for Principal Component Analyses of FTIR spectra of biochar for
each of three feedstocks: (a) macadamia nutshell, (b) eucalypt woodchip and (c) hemp stem.
For each feedstock, biochar was produced at three pyrolysis temperatures and three
durations (n = 27)
The first principal component (PC1) explained 80, 80 and 89 % of the variance in spectra of
nutshell, woodchip, and stem biochar, respectively. The loadings of PC1 for the nutshell and
woodchip PCA models were very similar, while the PC1 loadings of the stem PCA contained
similar peaks, but with different weightings (Fig. 31). In particular, the PC1 loadings for the
stem PCA was dominated by the 1720 cm-1 feature.

81
Figure 31. Loadings plot for PC1 of Principal Component Analyses of FTIR spectra of
biochar produced from each of (a) nutshell, (b) woodchip and (c) stem. For each feedstock,
biochar was produced at three pyrolysis temperatures and three durations (n = 27)
In a PCA of the combined set of all samples, PC1 achieved separation of biochars based on
stem from those based on woody feedstock, i.e., woodchip or nutshell (Fig. 32). In this
model, the loadings of PC1 weighted features associated with the stem spectra (Figs. 29,
32). PC2 of the combined sample PCA achieved separation of samples based on pyrolysis
temperature (Fig. 33), although the loadings of PC2 differed from the PC1 loadings which
separated samples based on pyrolsis temperature in the single feedstock PCAs. A PCA of
the woodchip and nutshell samples only also failed to separate samples based on feedstock,
although good separation in PC1 was achieved on pyrolysis temperature (Fig. 32).

82
a

Figure 32. Scores biplot for Principal Component Analyses of FTIR spectra for biochar
produced from (a) nutshell (Mac), woodchip (Euc), and stem (Hem), and (b) nutshell and
woodchip. For each feedstock, biochar was produced at three pyrolysis temperatures and
three durations (n = 27). In (a), red ovals outline the samples derived from woody materials
(nutshell and woodchip) and from stem material

83
Figure 33. Loadings plot for (a) PC1 and (b) PC2 of a Principal Component Analysis of FTIR
spectra of biochar produced from nutshell and woodchip. For each feedstock, biochar was
produced at three pyrolysis temperatures and three durations (n = 27)

84
CHAPTER 5. DISCUSSION and
INTERPRETATION
Apatite

The total phosphorus content increased when raw apatite was roasted for 350 ℃. As
moisture content and the volatile content of the raw apatite were each only 0.5 % w/w, this
increase is greater than expected and is ascribed to sampling variation. No significant
variation in total phosphorus was observed with increasing temperature from 350 to 600 ℃.
A decrease in extractable phosphorus was noted with roasting. This result was unexpected,
being inconsistent with the report Francisco et al. (2007) who reported an increase in
phosphorus availability from <1 % to over 80% when temperature increased from 500 to 700
℃. The observed decrease in apatite phosphorus availability following co-pyrolysis with
biochar may be associated with formation of insoluble organic compounds. However, Winter
(2018) (unpublished data) reported an increase in extractable P (in 2 % citric acid) from rock
apatite from the same source as used in the current study following pyrolysis for 10 min at
500 ℃ in the presence of organics, with increase from 9.5 % of total P in raw product to 17
%.

The current study was based on 2 % formic acid extraction rather than the 2 % citric acid
used by Winter (2018) however, formic acid generally results in higher values of extractable
P than citric acid (Camps-Arbestain et al., 2017; Zapata & Roy, 2004). The current study
was based on extraction of unmodified material, while Winter (2018) lightly ground the
pyrolyzed apatite and Francisco et al. (2007) treated the material in an aerobic environment
for 2 hours. It is possible that the high temperature treatment resulted in a surface sealing of
the apatite particles. The declining P availability could also be due to increasing crystallinity
with temperature (Glæsner et al., 2019).Thus, further work is recommended to characterise
apatite P extractability following repeated oxidation and reduction cycles, as would occur in a
full-scale pyrolysis plant utilising the fluidised bed material over a longer period.

Biochar properties in context of literature


A pyrolysis plant can be run for multiple objectives, with optimisation of outputs based on
market availability and value. Outputs can include electricity generation, CO2 generation
(e.g., for greenhouse production), volatile gases and oils, as well as biochar. For example, a
decrease in biochar yield with increased pyrolysis temperature is generally associated with
an increased yield of volatiles, i.e., syngas and oils (Mohd Hasan et al., 2019). Plant

85
conditions can also be optimised for each of these product classes, e.g., to produce biochar
of desired characteristics. In the current study, attention is given to the effect of feedstock
and pyrolysis conditions (time and temperature) on biochar properties.

Increasing pyrolysis temperature is expected to result in changes such as loss of functional


groups and increased surface area, and feedstock properties such as lignin and mineral
content are also expected to impact biochar characteristics (Tomczyk et al., 2020; Zhao et
al., 2017). The observed yields and biochar properties were broadly similar with literature
reports for pyrolysis of similar feedstock material and similar temperatures, despite use of
different reactor designs (Table 15). For example, a yield of 44 % was obtained for
macadamia nutshell when pyrolyzed at 450 ℃, while Lee et al. (2013) reported a yield of
around 40 % w/w for coconut shells pyrolyzed at 500℃ in a bench-scale fixed bed reactor for
1 h. Macadamia nutshell is reported to contain 41 % lignin, while coconut shell contains 46
% w/w lignin. A yield of 38 and 30 % were obtained for eucalyptus woodchip pyrolyzed at
350 and 450 ℃ in the current study, comparable to yields of 55 and 26 % w/w at pyrolysis
temperatures of 300 and 500 ℃, using a fixed-bed reactor and a residence time of 30 min,
as reported by Abdullah et al. (2010). Yields of 31 and 27 % were obtained for hemp stem
pyrolyzed at 350 and 450 ℃ in the current study, comparable to yields of 37 and 27 % w/w
yield at 400 and 600 ℃ using a microwave batch reactor operated for 60 min, as reported by
(Wallace et al., 2019).

86
Table 15. Similar reports of biochar properties in literature
Attributes This study Literature data Reference
pH 9.45 at 350; 10.36 at 600 ℃ for 7.8 at 300; 10.6 at 700℃ in peanut Tomczyk et al.
stem shell (2020)
7.5 at 300; 11.0 at 700 ℃ for herb Irfan et al. (2016)
Achnatherum
6.93 at 400; 8.82 at 550 ℃ for
5.6 at 350; to 7.8 at 600 ℃ for eucalyptus wood Singh et al. (2010a)
woodchip
EC 0.04 at 350; to 0.11 mS/cm at 0.09 at 400; 0.16 mS/cm at 550 ℃ for Singh et al. (2010a)
600 ℃ for woodchip eucalyptus woodchip

Volatile 46 at 350; to 16 % w/w at 60 at 300 to 15 % w/w at 500 ℃ for Abdullah et al.


matter eucalyptus woodchip (2010)
600 ℃ for woodchip

Fixed 46 at 350; to 68 % w/w at 600 ℃ 35 at 300; to 80 % w/w at 500 ℃ for Abdullah et al.
carbon for woodchip eucalyptus woodchip (2010)

Calorific 27 at 350; to 32 MJ/kg at 600 ℃ 29 at 400; to 33 MJ/kg at 600 ℃ for Gorshkov et al.
value for nutshell macadamia nutshell (2021)

23 at 350; to 33 MJ/kg at 650 ℃ for Jiang et al. (2018)


chestnut shell

Carbon 57 at 450; to 72 % w/w at 600 ℃ 60 at 300; to 85.5 % w/w at 500 ℃ for Abdullah et al.
for woodchip eucalyptus woodchip (2010)

Nitrogen 1.65 at 350; to 0.64 % w/w at 0.69 at 400; to 0.59 % w/w at 600 ℃ Wallace et al.
600 ℃ for stem for hemp stem (2019)

Hydrogen 5 at 350; to 3 % w/w at 600 ℃ 3.8 at 400; to 2.4 % w/w at 600 ℃ for Gorshkov et al.
for nutshell macadamia nutshell (2021)

5 at 350; to 3 % w/w at 600 ℃ 5.3 at 300; to 3 % w/w at 500 ℃ for Abdullah et al.
for woodchip eucalyptus woodchip (2010)

Sulphur 0.26 at 350; to 0.09 % w/w at 0.58 at 300; to 0.32 % w/w at 700 ℃ Zhao et al. (2018)
600 ℃ for stem for corn stem

CEC 59 at 350; to 15 cmol/kg at 600 10.8 at 350; to 1.4 cmol/kg at Domingues et al.
℃ for woodchip (2020)
750 ℃ for eucalyptus sawdust

Zeta -36 at 350; to -27 mV at 600 ℃ -48 at 300; to -37 mV at 600 ℃ for rice Hong et al. (2019)
potential for stem stem

Surface 141 at 350; to 306 m2/g at 600 5 at 400; to 185 m2/g at 600 ℃ for Gai et al. (2014)
area ℃ for stem peanut shell

Feedstocks
As expected from the literature, increasing woodiness of feedstock material (hemp stem <
eucalyptus woodchip < macadamia nutshell) was associated with higher yields, higher
biochar surface area and higher levels of functional groups, and thus higher associated

87
characteristics such as CEC, at a given pyrolysis temperature. The total PAH content of all
biochars was within the maximum threshold limit according to IBI guidelines and EBC
guidelines for basic grade biochar, however, nutshell and woodchip biochar exceeded the
maximum threshold limit for a premium grade biochar according to EBC guidelines (Schmidt
et al., 2016). Biochar mineral content was higher in less woody material, with relatively less
impact from temperature than other attributes, consistent with previous studies (Zhao et al.,
2017).

Pyrolysis time

The completion of the pyrolysis reaction within 5 minutes is attributed to the efficiency of the
fluidized bed reactor, which is based on the heat transfer rate of the bed material. The
choice of minimum time (5 minutes) thus underestimated the efficiency of the system. This
time was based on the report of Morgan et al. (2015) that pyrolysis was complete in 10
minutes in a fluidised bed system using 7.5 g of feedstock in a 500 g sand bed, i.e., a 1:66
ratio. The current system used apatite, which has a slightly lower heat capacity than sand,
and a feedstock to bed weight ratio of 1:6. Both factors should contribute to a slower rate of
reaction in the current system relative to that of Morgan et al. (2015).

Further work should be undertaken to establish a minimum residence time for the system
employed. Shorter residence times offer greater operating efficiency.

Pyrolysis temperature

The pyrolysis temperature significantly impacted structural and physicochemical properties


of biochar, with results consistent with those of previous studies involving similar feedstock
(Table 15).

Proximate analysis (moisture, volatile matter, ash, and fixed carbon content)

High temperatures used in pyrolysis result in removal of all loosely associated water.
However, moisture can be absorbed from the atmosphere during cooling and storage. As
this process was not controlled, it is unclear whether the differences in moisture content of
the biochar are due to variation in hygroscopicity or variation in handling.

Volatile matter decreased with the temperature during pyrolysis. This decrease is attributed
to the sequential decomposition of biopolymers; cellulose, hemicellulose and lignin (Bolan et
al., 2021), and dehydration of hydroxyl groups (Tomczyk et al., 2020), with increasing
temperature. A decrease in volatile matter will reduce the availability of organic matter to soil
microbes; thus, lower temperature biochar is more suited for use as a soil conditioner.
Similarly, in higher temperature biochar, there is nitrogen loss and wide C:N ratios, thus

88
biochar can be a good carbon source for microbes. However, increasing C:N ratio can lead
N immobilisation when applied to soil (Phillips et al., 2022).

Ash content significantly increased with increasing temperature for all biochar feedstocks.
This result is as expected as increasing temperature results in increasing loss of organic
materials (Tsai et al., 2012). Increasing ash content is correlated to water-extractable
mineral content and EC (Gai et al., 2014). Higher ash containing biochar has higher fertiliser
value.

Fixed carbon increased with temperature as expected, as less resistant forms of carbon are
progressively removed as pyrolysis temperature increases (Zimmerman et al., 2011).

Calorific value and CHNS composition

Biochar calorific value increased with increased pyrolysis temperature, consistent with
increasing % carbon content (Wardani et al., 2018), associated with a high polymerisation
reaction forming a more condensed carbon structure in biochar (Domingues et al., 2017;
Lehmann & Joseph, 2009).

Nitrogen content decreased with an increase in temperature, consistent with increasing


gaseous emission of nitrogen (volatilisation) with increasing temperature (Bagreev et al.,
2001). Hydrogen content decreased with temperature, consistent with a decrease in
oxygenated compounds with the destruction of O-H bonds (Suliman et al., 2016). Sulphur
content declined with temperature in stem biochar owing to the volatilisation of sulphur
containing organic compounds (Al-Wabel et al., 2013).

Bulk density, pH, EC and elemental concentration

The bulk density of biochar decreased with pyrolysis temperature, consistent with literature
e.g., Khanmohammadi et al. (2015) and Ronewicz et al. (2017). This trend is explained by
the loss of components such as hemicelluloses and cellulose, leaving a thermally stable
material skeleton based on material such as lignin (Ding et al., 2010).

Biochar pH and EC values tended to increase with increasing temperature. This trend
reflects a concentration of minerals and liming agents, and loss of acidic functional groups at
higher temperatures, coupled with lower yields.

Pyrolysis temperature was correlated with the water holding capacity of biochar. This trend is
consistent with increased porosity, as cellular materials are progressively removed to leave a
skeleton of more resistant materials, thus increasing water holding capacity (Usevičiūtė &
Baltrėnaitė-Gedienė, 2020).

89
Elemental constituents

The mineral element content of biochar is highly dependent on the feedstock used (Ippolito
et al., 2015), with relatively less impact from temperature and residence time (Zhao et al.,
2017). In general, elemental concentrations tended to increase with pyrolysis temperature as
C is lost (Yuan et al., 2011). In contrast, the concentration of water extractable elements
tended to decline with pyrolysis temperature. For example, nutshell and stem contained 91
and 79 % of feedstock Ca, but only 5 and 1 % of water extractable Ca as compared to the
water extractable Ca in feedstock. Presumably, less Ca is water soluble in biochar than
feedstock because of the formation of insoluble Ca compounds. K apparently volatilised,
with biochar K content being 16, 14 and 19 % of that in feedstock in nutshell, woodchip and
stem, respectively. However, more K was in water soluble form, at 308, 300 and 113% of
feedstock in nutshell, woodchip and stem biochar, respectively.

Cation exchange capacity and zeta potential

The values of CEC are very high, especially for 600℃ nutshell and woodchip biochar. Use of
H+ as cation (pH 5) could be a problem. At low pH H+ ion can be retained by aromatic C in π
cloud. Cation exchange capacity decreased with pyrolysis temperature, as expected from
literature e.g., Mukherjee et al. (2011) and Domingues et al. (2020). This trend is explained
by a decrease in oxygen - containing functional groups, such as COOH, with temperature.
Low-temperature biochar, with higher cation exchange capacity can be recommended for
use when the aim is to improve the CEC of a soil a sandy soil (Tomczyk et al., 2020).

Stem biochar showed increased zeta potential, i.e., weaker electronegativity, at higher
temperatures. This increasing zeta potential is highly associated with the declining hydrogen
and oxygen bonds with the increasing temperature which ultimately decreases the
electronegativity (Hong et al., 2019). However, zeta potential decreased with temperature in
nutshell biochar. Qi et al. (2017) reported that zeta potential is impacted by the concentration
of mineral and organic components.

Surface area and morphology

Biochar surface area increased significantly with pyrolysis temperature for all feedstocks.
This result is associated with increasing pore size and volume, with open pores formed as
volatiles are released from the biochar (Conte et al., 2021). The surface area of biochar is
also determined by the feedstock composition (cellulose, hemicellulose, and lignin)
(Shaaban et al., 2014). A higher surface area enables biochar to hold water, nutrients and
absorb contaminants from soil and water (Cisse, 2021).

90
PAHs and PCBs

The total PAHs content of all the biochars was within the maximum threshold limit according
to IBI guidelines and EBC guidelines for basic grade biochar. However, nutshell and
woodchip biochar exceeded the maximum threshold limit for a premium grade biochar
according to EBC guidelines (Schmidt et al., 2016). These feedstocks are woody, containing
more lignin, supporting greater aromatisation during pyrolysis, thus leading to the formation
of phenolic compounds the PAHs such as naphthalene (Wang et al., 2017). Similarly, the
total PCBs content of all biochar exceeded the maximum threshold limit i.e., <0.2 mg/kg
according to EBC guidelines (Schmidt et al., 2016). However, as per the IBI guidelines, total
PCBs for woodchip and stem biochar was still within the threshold limit of 0.2 -0.5 mg/kg
(Bucheli et al., 2015).

Prediction of production conditions using FTIR

The spectral variation between the feedstocks materials was attributed to the O-H, C-O-C,
and C-O stretching associated with cellulose, hemicellulose and lignin (Lazzari et al., 2018).
PC1 of a PCA on spectra of feedstocks separated samples based on feedstock, with loading
that were consistent with a high influence of C=O and C-O-C stretching, likely associated
with carboxylic acids, amides, esters, and ketones (Johnston, 2017). The mid-infrared
spectra varied with pyrolysis temperature, with little impact of residence time. This result is
consistent with loss of functional groups and increase in persistent carbon and the
aromaticity of biochar (C=C stretching) with increasing temperature (Sohi et al., 2010).
(Yuan et al., 2011). A PCA on spectra of all 81 biochar samples (3 replicates of each of 27
biochar conditions of feedstock-temperature-time) demonstrated poor separation of samples
based on replication or pyrolysis time, consistent with measurements of physical and
chemical parameters, with separation of samples occurring on the basis of feedstock and
temperature, with nutshell and woodchip more resembling to each other than stem. This
result was expected from the relative woodiness of the materials.

For a PCA of biochars of a given feedstock, PC1 clearly separated samples based on
temperature. The loadings of PC1 of the two woody material feedstocks, nutshell and
woodchip, were very similar, indicating similar chemical processes were occurring at the
different temperatures. PC1 of the hemp stem, however, was quite different, indicating that
different chemical changes were occurring for this material as temperature increased. PC1
loadings for the stem PCA model showed a marked feature at 1720 cm-1 attributed to the
C=O from carboxylic acids, amides and ketones (Johnston, 2017).

It is therefore possible that FTIR could be used in a quality control program to monitor quality
attributes of feedstock and biochar in relation to the conditions of production (temperature).

91
Designer biochar

Biochar has a range of potential applications, each with different requirements in terms of
biochar properties (see Chapter 2). Based on the observations of the current study,
recommendations on feedstock and pyrolysis conditions to achieve specifications associated
with a number of biochar uses are given in Table 16. For example, for fertility enhancement
of a sandy soil, biochar with a higher cation exchange capacity, mineral and ash content is
desirable. Of the feedstocks and pyrolysis conditions trialled, these attributes are best
achieved using hemp stem as a feedstock, with biochar production at a lower pyrolysis
temperature (350 ℃). For use as an energy fuel source, a higher calorific value is desired
which can be delivered using macadamia nutshell or eucalyptus woodchip and higher
pyrolysis temperature (600 ℃) is recommended. Furthermore, biochar for animal feed
additive requiring higher surface area and absorption capacity can be achieved with higher
temperature (600 ℃), with macadamia nutshell and eucalyptus woodchip biochar are
recommended. For the wastewater treatment, higher absorption capacity and higher surface
area are desired which can be achieved with higher temperature biochar (macadamia
nutshell and eucalyptus woodchip biochar at 600 ℃).

92
Table 16. Optimisation of biochar produced from fluidised bed reactor to their specific uses
Uses Key required References Feedstocks
biochar properties
Macadamia Eucalyptus Hemp
350℃ 450℃ 600℃ 350℃ 450℃ 600℃ 350℃ 450℃ 600℃
Soil amendment
increased water holding capacity in sandy soil high water holding Asai et al. (2009a)
capacity
decreased bulk density, increased porosity of low bulk density Blanco-Canqui (2017)
heavy soil
decreased soil acidity and aluminium toxicity. high pH and liming Shetty and Prakash
potential (2020); Tsai et al. (2012)
enhanced mineral level in sandy soil high mineral content Ding et al. (2016);
and high cation Munera et al. (2018);
exchange capacity Wu et al. (2012)
carbon sequestration high carbon content Chan and Xu (2012);
Rehrah et al. (2014)
adsorption of cationic compounds low zeta potential Jatav et al. (2021); Yang
et al. (2020)
Increased soil microbial activity high volatiles content Tomczyk et al. (2020)
Renewable energy
fuel source high calorific value Dąbrowska et al. (2018);
Suman (2020)
Animal feed additive
adsorption of toxins, heavy metals, and absence or minimal Schmidt et al. (2019)
pathogens. presence of PAHs
and PCBs

93
enhance redox chemical reactions for improving high surface area Sun et al. (2017b); Teoh
the feed efficiency, digestibility, and reducing et al. (2019)
methane emissions.
Wastewater treatment
adsorption of toxins, heavy metals, and high surface area Bolton et al. (2019);
pathogens. and surface Yang et al. (2020)
functional groups

94
Recently, researchers have introduced the concept of pre-treated biochar (feedstocks
treated prior to pyrolysis), where biochar is treated physically, chemically, and biologically to
optimise the properties of biochar for designated use. Different techniques such as ball
milling, steam activation, magnetisation, and irradiation have been practiced in biochar
modification (Li et al., 2020). For example, specific surface area was found to be higher in
ball-milled biochar than the untreated biochar produced from sugarcane bagasse, bamboo,
and woodchips (Lyu et al., 2018). Fahmi et al. (2018) also reported that the physically
modified biochar tends to have a higher surface area and the greater number of exposed
micropores, thus resulting in the enhanced adsorption of lead. Similarly, Rajapaksha et al.
(2015) reported the enhanced adsorption capacity of a steam-activated biochar over the
non-activated biochar from the same feedstock. Thus, steam-activated biochar has also
been reported used in agricultural soil for reduction of methane and nitrous oxide emissions
(Fungo et al., 2014).

In addition, biochar has been treated with strong acids, bases, and oxidizing agents to retain
the oxygen-containing or carboxyl functional groups to enhance nutrient retention and
adsorptive capacity (Li et al., 2020). For example, chemical activation of biochar with zinc
chloride was found to result in increased surface area as reported in Angın et al. (2013).
Similarly, Peng et al. (2017) reported an enhanced adsorption of Cu and Cd with the
phosphoric acid modification of biochar. Acid modification resulted in increased surface area
and surface functional groups responsible for the higher adsorption of metal ions (Peng et
al., 2017). Similarly, composting of biochar has been reported to increase the nutrient
concentration in biochar thus the composted biochar can offer useful soil amendment by
increasing the nutrient interactions for the cropping benefits (Joseph et al., 2018a).

Hence, biochar can be engineered for its desired performance, however, a careful
consideration should be given during the physical and chemical activation for the intended
applications of biochar, whether as an animal feed additive, or soil amendments aimed for
environmental sustainability.

Choice of measurement attributes


There are many attributes that can be measured of biochar, as documented in references
provided in Table 15. For characterisation of biochar from a pyrolysis plant using set
feedstocks, attributes to be measured can be narrowed to those relevant to the intended
market application, as listed in Table 16. For example, the key attribute for biochar use as an
energy source is calorific value.

Some measurements, however, require access to specialist equipment or are relatively


expensive. Therefore, the correlation of attribute values was considered (Fig. 3), for the

95
recommendation of easier to assess attributes. Correlations do not prove causal relations, but
for a set of defined conditions, e.g., pyrolysis conditions and feedstocks, such correlations will
have predictive value. For example, Carbon content was well correlated (R 0.83) to calorific
value. Strong relationships were also evidenced between pH and Liming potential (0.91),
density and yield (R 0.88) and density and zeta potential (R 0.74), and Electrical Conductivity
and Water Holding Capacity (0.94). Zeta potential has been recommended as an alternative
to Cation Exchange Capacity as an index of the value of biochar for cation retention (Fahmi
et al., 2018), however the correlation between zeta potential and CEC was poor.

Figure 34. Correlation coefficients (R) between pairs of measured attributes for the 81
samples (three pyrolysis temperature*three pyrolysis times*3feedstocks*3replicates). Colors
are keyed to R values (-1 to +1)

96
CHAPTER 6. CONCLUSION and
RECOMMENDATIONS
This study was designed to provide recommendations on optimal pyrolysis conditions and
feedstock to achieve desired properties of biochar in respective areas of agriculture, animal
feed and environment. The three feedstocks (macadamia nutshell, eucalyptus woodchips
and hemp stem) ranging in lignification and woodiness, pyrolysis times (5, 10 and 20 min)
and production temperatures (350, 450 and 600 ℃) were selected to produce biochars with
a range of properties for the multiple applications discussed above.

The pH, EC, liming potential, ash content, carbon content, calorific value, water holding
capacity, surface area and aromaticity increased with elevated temperature. In contrast, the
bulk density, volatile matter, cation exchange capacity, zeta potential and functional groups
decreased with temperature. Feedstock material also determined the physicochemical
characteristics as woody materials resulted higher carbon content biochar while stem
content feedstock resulted in a biochar with higher ash and nutrient content.

Thus, appropriate feedstock-temperature combinations can be selected to match needs of


the range of biochar for agriculture, animal feed and environmental applications. For
example, low-temperature biochar which has a higher CEC can be recommended for use
when the aim is to improve the CEC of a sandy soil. Higher temperature biochar which has
higher surface area is recommended for toxin and heavy metal absorption for wastewater
treatment.

Further research work is suggested for greater control of consistent quality attributes end
use specific biochar products:

(i) Pre-treatments and co-amendments of feedstock for enhancing value of biochar


products. For example, extra P might be loaded, or a microbial supplement added

(ii) Optimisation of the fluid bed pyrolyser for harvesting by products of pyrolysis
(volatiles, CO2) to achieve energy efficiency and emission minimisation

(iii) Given the competing uses of the fluidised bed material as a P fertilizer, optimised
conditions to increase phosphorus availability in apatite should be further explored

97
REFERENCES
Abdullah, H., Mediaswanti, K. A., & Wu, H. (2010). Biochar as a fuel: 2. Significant
differences in fuel quality and ash properties of biochars from various biomass
components of Mallee trees. Energy & Fuels, 24(3), 1972-1979.
Adeniji, A. O., Okoh, O. O., & Okoh, A. I. (2018). Analytical methods for polycyclic
aromatic hydrocarbons and their global trend of distribution in water and sediment: a
review. Recent insights in petroleum science and engineering, 10.
Al-Wabel, M. I., Al-Omran, A., El-Naggar, A. H., Nadeem, M., & Usman, A. R. A. (2013).
Pyrolysis temperature induced changes in characteristics and chemical composition of
biochar produced from conocarpus wastes. Bioresource Technology, 131, 374-379.
doi:https://doi.org/10.1016/j.biortech.2012.12.165
Aller, D., Bakshi, S., & Laird, D. A. (2017). Modified method for proximate analysis of
biochars. Journal of Analytical and Applied Pyrolysis, 124, 335-342.
Andrigo, P., Bagatin, R., & Pagani, G. (1999). Fixed bed reactors. Catalysis Today, 52(2),
197-221. doi:https://doi.org/10.1016/S0920-5861(99)00076-0
Angın, D., Altintig, E., & Köse, T. E. (2013). Influence of process parameters on the surface
and chemical properties of activated carbon obtained from biochar by chemical
activation. Bioresource Technology, 148, 542-549.
doi:https://doi.org/10.1016/j.biortech.2013.08.164
Arbestain, M. C., Amonette, J., Singh, B., Wang, T., & Schmidt, H.-P. (2015). A biochar
classification system and associated test methods. In (pp. 165-194).
Asai, H., Samson, B. K., Stephan, H. M., Songyikhangsuthor, K., Homma, K., Kiyono, Y.,
Inoue, Y., Shiraiwa, T., & Horie, T. (2009a). Biochar amendment techniques for
upland rice production in Northern Laos: 1. Soil physical properties, leaf SPAD and
grain yield. Field Crops Research, 111(1-2), 81-84.
Asai, H., Samson, B. K., Stephan, H. M., Songyikhangsuthor, K., Homma, K., Kiyono, Y.,
Inoue, Y., Shiraiwa, T., & Horie, T. (2009b). Biochar amendment techniques for
upland rice production in Northern Laos: 1. Soil physical properties, leaf SPAD and
grain yield. Fields Crop Research, 111(1-2), 81-84.
Ashwath, N., Nam, H., & Capareda, S. C. (2018). Optimising Pyrolysis Conditions for
Thermal Conversion of Beauty Leaf Tree (Calophyllum inophyllum L.) Press Cake.
In Application of Thermo-fluid Processes in Energy Systems (pp. 267-280): Springer.
Askeland, M., Clarke, B., & Paz-Ferreiro, J. (2019). Comparative characterization of biochars
produced at three selected pyrolysis temperatures from common woody and
herbaceous waste streams. PeerJ, 7, e6784. doi:10.7717/peerj.6784
Atkinson, C. J., Fitzgerald, J. D., & Hipps, N. A. (2010). Potential mechanisms for achieving
agricultural benefits from biochar application to temperate soils: a review. Plant soil,
337(1-2), 1-18.
Bagreev, A., Bandosz, T. J., & Locke, D. C. (2001). Pore structure and surface chemistry of
adsorbents obtained by pyrolysis of sewage sludge-derived fertilizer. Carbon, 39(13),
1971-1979. doi:https://doi.org/10.1016/S0008-6223(01)00026-4
98
Bai, M., Wilske, B., Buegger, F., Esperschütz, J., Kammann, C. I., Eckhardt, C., Koestler,
M., Kraft, P., Bach, M., & Frede, H.-G. (2013). Degradation kinetics of biochar from
pyrolysis and hydrothermal carbonization in temperate soils. Plant soil, 372(1-2),
375-387.
Bamido, A. O. (2018). Design of A Fluidized Bed Reactor For Biomass Pyrolysis. University
of Cincinnati,
Barker, A. V. (2019). Fertilizers☆. In P. Worsfold, C. Poole, A. Townshend, & M. Miró
(Eds.), Encyclopedia of Analytical Science (Third Edition) (pp. 134-144). Oxford:
Academic Press.
Batista, E. M. C. C., Shultz, J., Matos, T. T. S., Fornari, M. R., Ferreira, T. M., Szpoganicz,
B., de Freitas, R. A., & Mangrich, A. S. (2018). Effect of surface and porosity of
biochar on water holding capacity aiming indirectly at preservation of the Amazon
biome. Scientific reports, 8(1), 10677. doi:10.1038/s41598-018-28794-z
Bird, M., Keitel, C., & Meredith, W. (2017). Analysis of biochars for C, H, N, O and S by
elemental analyser. In M. C.-A. a. J. L. Balwant Singh (Ed.), Biochar: A Guide to
Analytical Methods (pp. 39).
Blanco-Canqui, H. (2017). Biochar and soil physical properties. Soil Science Society of
America Journal, 81(4), 687-711.
Bolan, N., Hoang, S., Beiyuan, J., Gupta, S., Hou, D., Karakoti, A., Joseph, S., Jung, S., Kim,
K.-H., Kirkham, M., Kua, H., Kumar, M., Kwon, E., Ok, Y., Perera, V., Rinklebe, J.,
Shaheen, S., Sarkar, B., Sarmah, A., & Van Zwieten, L. (2021). Multifunctional
applications of biochar beyond carbon storage. International Materials Reviews.
doi:10.1080/09506608.2021.1922047
Bolton, L., Joseph, S., Greenway, M., Donne, S., Munroe, P., & Marjo, C. (2019).
Phosphorus adsorption onto an enriched biochar substrate in constructed wetlands
treating wastewater. Ecological Engineering: X, 1, 100005.
Bridgwater, A. V. (2012). Review of fast pyrolysis of biomass and product upgrading.
Biomass bioenergy, 38, 68-94.
Bucheli, T. D., Hilber, I., & Schmidt, H.-P. (2015). Polycyclic aromatic hydrocarbons and
polychlorinated aromatic compounds in biochar. In Biochar for environmental
management (pp. 627-656): Routledge.
Camps-Arbestain, M., Shen, Q., Wang, T., van Zwieten, L., & Novak, J. (2017). 10 Available
nutrients in biochar. Biochar: A Guide to Analytical Methods, 109.
Campuzano, F., Brown, R. C., & Martínez, J. D. (2019). Auger reactors for pyrolysis of
biomass and wastes. Renewable and Sustainable Energy Reviews, 102, 372-409.
doi:https://doi.org/10.1016/j.rser.2018.12.014
Cao, T., Chen, F. w., & Meng, J. (2018). Influence of pyrolysis temperature and residence
time on available nutrients for biochars derived from various biomass. Energy
Sources, Part A: Recovery, Utilization, and Environmental Effects, 40(4), 413-419.
doi:10.1080/15567036.2016.1225137
Castilla-Caballero, D., Barraza-Burgos, J., Gunasekaran, S., Roa-Espinosa, A., Colina-
Márquez, J., Machuca-Martínez, F., Hernández-Ramírez, A., & Vázquez-Rodríguez,

99
S. (2020). Experimental data on the production and characterization of biochars
derived from coconut-shell wastes obtained from the Colombian Pacific Coast at low
temperature pyrolysis. Data in Brief, 28, 104855.
doi:https://doi.org/10.1016/j.dib.2019.104855
Cha, J. S., Park, S. H., Jung, S.-C., Ryu, C., Jeon, J.-K., Shin, M.-C., & Park, Y.-K. (2016).
Production and utilization of biochar: A review. Journal of Industrial and
Engineering Chemistry, 40, 1-15. doi:https://doi.org/10.1016/j.jiec.2016.06.002
Chan, K. Y., & Xu, Z. (2012). Biochar: nutrient properties and their enhancement. In Biochar
for environmental management (pp. 99-116): Routledge.
Chen, B., & Chen, Z. (2009). Sorption of naphthalene and 1-naphthol by biochars of orange
peels with different pyrolytic temperatures. Chemosphere, 76(1), 127-133.
doi:https://doi.org/10.1016/j.chemosphere.2009.02.004
Chen, B., Zhou, D., & Zhu, L. (2008). Transitional adsorption and partition of nonpolar and
polar aromatic contaminants by biochars of pine needles with different pyrolytic
temperatures. Environmental Science & Technology, 42(14), 5137-5143.
Cheruvu, N. (2018). Humanitarian Approach to Biochar. Retrieved from
https://pdfs.semanticscholar.org/1edc/614f9eaab7a6ba3a7100c993713e3120e64e.pdf
Cisse, I. (2021). Characterization of biochar produced by pyrolysis of biomass and co-
pyrolysis of biomass and agricultural mulch film.
Conte, P., Bertani, R., Sgarbossa, P., Bambina, P., Schmidt, H.-P., Raga, R., Lo Papa, G.,
Chillura Martino, D. F., & Lo Meo, P. (2021). Recent Developments in
Understanding Biochar’s Physical–Chemistry. Agronomy, 11(4), 615.
Conte, P., Schmidt, H.-P., & cimò, G. (2015). Research and Application of Biochar in
Europe. In.
Cornelissen, G., Gustafsson, Ö., Bucheli, T. D., Jonker, M. T., Koelmans, A. A., & van
Noort, P. C. (2005). Extensive sorption of organic compounds to black carbon, coal,
and kerogen in sediments and soils: mechanisms and consequences for distribution,
bioaccumulation, and biodegradation. Environmental Science and Technology,
39(18), 6881-6895.
Dąbrowska, M., Jaworek, M., Świętochowski, A., & Lisowski, A. (2018). VALUABLE
ENERGY OF BIOCHAR FROM AGRICULTURAL AND FOREST WASTE
STREAMS.
Das, O., & Sarmah, A. K. (2015). The love–hate relationship of pyrolysis biochar and water:
A perspective. Science of The Total Environment, 512-513, 682-685.
doi:https://doi.org/10.1016/j.scitotenv.2015.01.061
Das, S. K., Ghosh, G. K., Avasthe, R., & Kundu, M. C. (2018). Preparation and
Characterization of Biochars for their Application as Soil Amendment. Indian Journal
of Hill Farming, 31(1), 141-145.
De Pasquale, C., Marsala, V., Berns, A. E., Valagussa, M., Pozzi, A., Alonzo, G., & Conte, P.
(2012). Fast field cycling NMR relaxometry characterization of biochars obtained
from an industrial thermochemical process. Journal of Soils and Sediments, 12(8),
1211-1221. doi:10.1007/s11368-012-0489-x

100
Derlet, R. W., & Albertson, T. E. (1986). Activated charcoal--past, present and future. The
Western journal of medicine, 145 4, 493-496.
Devi, P., & Saroha, A. K. (2015). Effect of pyrolysis temperature on polycyclic aromatic
hydrocarbons toxicity and sorption behaviour of biochars prepared by pyrolysis of
paper mill effluent treatment plant sludge. Bioresource Technology, 192, 312-320.
doi:https://doi.org/10.1016/j.biortech.2015.05.084
Dhyani, V., & Bhaskar, T. (2018). A comprehensive review on the pyrolysis of
lignocellulosic biomass. Renewable Energy, 129, 695-716.
doi:https://doi.org/10.1016/j.renene.2017.04.035
Ding, Y., Liu, Y.-X., Wu, W.-X., Shi, D.-Z., Yang, M., & Zhong, Z.-K. (2010). Evaluation of
Biochar Effects on Nitrogen Retention and Leaching in Multi-Layered Soil Columns.
Water, Air, & Soil Pollution, 213(1), 47-55. doi:10.1007/s11270-010-0366-4
Ding, Y., Liu, Y., Liu, S., Li, Z., Tan, X., Huang, X., Zeng, G., Zhou, L., & Zheng, B. (2016).
Biochar to improve soil fertility. A review. Agronomy for Sustainable Development,
36(2), 36. doi:10.1007/s13593-016-0372-z
Domingues, R. R., Sánchez-Monedero, M. A., Spokas, K. A., Melo, L. C. A., Trugilho, P. F.,
Valenciano, M. N., & Silva, C. A. (2020). Enhancing Cation Exchange Capacity of
Weathered Soils Using Biochar: Feedstock, Pyrolysis Conditions and Addition Rate.
Agronomy, 10(6), 824.
Domingues, R. R., Trugilho, P. F., Silva, C. A., de Melo, I. C. N., Melo, L. C., Magriotis, Z.
M., & Sanchez-Monedero, M. (2017). Properties of biochar derived from wood and
high-nutrient biomasses with the aim of agronomic and environmental benefits. Plos
one, 12(5).
Drouet, C. (2015). A comprehensive guide to experimental and predicted thermodynamic
properties of phosphate apatite minerals in view of applicative purposes. The Journal
of Chemical Thermodynamics, 81, 143-159.
doi:https://doi.org/10.1016/j.jct.2014.09.012
Dukua, M. H., Gu, S. G., & Hagan, E. B. (2011). "Biochar Production Potential in Ghana—A
Review,". Renewable and Sustainable Energy Reviews, 15, 3539-3551.
Eigenberger, G. (1992). Fixed bed reactors.
Enders, A., & Lehmann, J. (2017). Proximate analyses for characterising biochar. In M. C.-A.
a. J. L. Balwant Singh (Ed.), Biochar: A Guide to Analytical Methods (pp. 9-22).
Fahmi, A. H., Jol, H., & Singh, D. (2018). Physical modification of biochar to expose the
inner pores and their functional groups to enhance lead adsorption. RSC advances,
8(67), 38270-38280.
Francisco, E. A. B., Prochnow, L. I., Toledo, M. C. M. d., Ferrari, V. C., & Jesus, S. L. d.
(2007). Thermal treatment of aluminous phosphates of the crandallite group and its
effect on phosphorus solubility. Scientia Agricola, 64, 269-274.
Fungo, B., Guerena, D., Thiongo, M., Lehmann, J., Neufeldt, H., & Kalbitz, K. (2014). N2O
and CH4 emission from soil amended with steam‐activated biochar. Journal of
Plant Nutrition and Soil Science, 177(1), 34-38.

101
Gai, X., Wang, H., Liu, J., Zhai, L., Liu, S., Ren, T., & Liu, H. (2014). Effects of Feedstock
and Pyrolysis Temperature on Biochar Adsorption of Ammonium and Nitrate. Plos
one, 9(12), e113888. doi:10.1371/journal.pone.0113888
Galvano, F., Piva, A., Ritieni, A., & Galvano, G. (2001). Dietary strategies to counteract the
effects of mycotoxins: a review. Journal of food protection, 64 1, 120-131.
Garcia-Perez, M., & Metcalf, J. (2008). The formation of polyaromatic hydrocarbons and
dioxins during pyrolysis: A review of the literature with descriptions of biomass
composition, fast pyrolysis technologies and thermochemical reactions.
Gelardi, D. L., & Parikh, S. J. (2021). Soils and Beyond: Optimizing Sustainability
Opportunities for Biochar. Sustainability, 13(18), 10079.
Glæsner, N., Hansen, H. C. B., Hu, Y., Bekiaris, G., & Bruun, S. (2019). Low crystalline
apatite in bone char produced at low temperature ameliorates phosphorus-deficient
soils. Chemosphere, 223, 723-730.
Glaser, B., Balashov, E., Haumaier, L., Guggenberger, G., & Zech, W. (2000). Black carbon
in density fractions of anthropogenic soils of the Brazilian Amazon region. Organic
Geochemistry, 31(7), 669-678. doi:https://doi.org/10.1016/S0146-6380(00)00044-9
Glaser, B., Lehmann, J., & Zech, W. (2002). Ameliorating physical and chemical properties
of highly weathered soils in the tropics with charcoal–a review. Biology Fertility of
Soils, 35(4), 219-230.
Godbey, H. (2016). Biochar, A Brief History. Permaculture news. Retrieved from
https://www.permaculturenews.org/2016/11/18/biochar-brief-history/
Goodwin, S. (2011). BiG future for Biochar. Retrieved from
https://www.farmonline.com.au/story/3621724/big-future-for-biochar/
Gorshkov, A., Berezikov, N., Kaltaev, A., Yankovsky, S., Slyusarsky, K., Tabakaev, R., &
Larionov, K. (2021). Analysis of the Physicochemical Characteristics of Biochar
Obtained by Slow Pyrolysis of Nut Shells in a Nitrogen Atmosphere. Energies,
14(23), 8075.
GRDC. (2013). Understanding Biochar. Biochar fact Sheet.
Hale, S. E., Lehmann, J., Rutherford, D., Zimmerman, A. R., Bachmann, R. T.,
Shitumbanuma, V., O’Toole, A., Sundqvist, K. L., Arp, H. P. H., & Cornelissen, G.
(2012). Quantifying the total and bioavailable polycyclic aromatic hydrocarbons and
dioxins in biochars. Environmental Science and Technology, 46(5), 2830-2838.
Herzel, H., Aydin, Z., & Adam, C. (2021). Crystalline phase analysis and phosphorus
availability after thermochemical treatment of sewage sludge ash with sodium and
potassium sulfates for fertilizer production. Journal of Material Cycles and Waste
Management, 23(6), 2242-2254. doi:10.1007/s10163-021-01288-3
Hong, M., Zhang, L., Tan, Z., & Huang, Q. (2019). Effect mechanism of biochar’s zeta
potential on farmland soil’s cadmium immobilization. Environmental Science and
Pollution Research, 26(19), 19738-19748.
Hossain, M. K., Strezov, V., Chan, K. Y., Ziolkowski, A., & Nelson, P. F. (2011). Influence
of pyrolysis temperature on production and nutrient properties of wastewater sludge
biochar. Journal of Environmental Management, 92(1), 223-228.

102
Huckfeldt, L. (2020). Biochar as Carbon Sink, Summary, Opportunities for Carbon
Sequestration in Agricultural Soils: Technology, Potential Feedstock and Estimation
on Climate Change Mitigation Potential.
Huff, M. D., & Lee, J. W. (2016). Biochar-surface oxygenation with hydrogen peroxide.
Journal of Environmental Management, 165, 17-21.
doi:https://doi.org/10.1016/j.jenvman.2015.08.046
IBI. (2018a). Biochar Feedstocks. Retrieved from https://biochar-international.org/biochar-
feedstocks/
IBI. (2018b). Biochar Is a Valuable Soil Amendment. Biochar. Retrieved from
https://biochar-international.org/about-ibi/
IBI. (2018c). Biochar Production Technologies. Retrieved from https://biochar-
international.org/biochar-production-technologies/
IBI. (2018d). Biochar Production Technologies. Retrieved from https://biochar-
international.org/biochar-production-technologies/
Ibrahim, H. A.-H. (2020). Introductory Chapter: Pyrolysis. Recent Advances in Pyrolysis, 1.
Intani, K., Latif, S., Cao, Z., & Müller, J. (2018). Characterisation of biochar from maize
residues produced in a self-purging pyrolysis reactor. Bioresource Technology, 265,
224-235.
International Biochar Initiative, I. (2015, 23/11/2015). Standardized Product Definition and
Product Testing Guidelines for Biochar That Is Used In Soil. Retrieved from
https://www.biochar-international.org/wp-
content/uploads/2018/04/IBI_Biochar_Standards_V2.1_Final.pdf
Ippolito, J., Spokas, K. A., Novak, J., Lentz, R., Cantrell, K. B., Lehmann, J., & Joseph, S.
(2015). Biochar elemental composition and factors influencing nutrient retention.
Biochar for envrionmental management: science, technology and implementation,
137-161.
Ippolito, J. A., Cui, L., Kammann, C., Wrage-Mönnig, N., Estavillo, J. M., Fuertes-
Mendizabal, T., Cayuela, M. L., Sigua, G., Novak, J., Spokas, K., & Borchard, N.
(2020). Feedstock choice, pyrolysis temperature and type influence biochar
characteristics: a comprehensive meta-data analysis review. Biochar, 2(4), 421-438.
doi:10.1007/s42773-020-00067-x
Irfan, M., Chen, Q., Yue, Y., Pang, R., Lin, Q., Zhao, X., & Chen, H. (2016). Co-production
of biochar, bio-oil and syngas from halophyte grass (Achnatherum splendens L.)
under three different pyrolysis temperatures. Bioresource Technology, 211.
doi:10.1016/j.biortech.2016.03.077
Jatav, H., Rajput, V., Minkina, T., Singh, S., Sukirtee, Gorovtsov, A., Barakhov, A., Bauer,
T., Sushkova, S., Mandzhieva, S., Burachevskaya, M., & Kalinichenko, V. (2021).
Sustainable Approach and Safe Use of Biochar and Its Possible Consequences.
Sustainability, 13. doi:10.3390/su131810362
Jiang, K.-m., Cheng, C.-g., Ran, M., Lu, Y.-g., & Wu, Q.-l. (2018). Preparation of a biochar
with a high calorific value from chestnut shells. New Carbon Materials, 33(2), 183-
187. doi:https://doi.org/10.1016/S1872-5805(18)60333-6

103
Jiang, S., Nguyen, T. A., Rudolph, V., Yang, H., Zhang, D., Ok, Y. S., & Huang, L. (2017).
Characterization of hard-and softwood biochars pyrolyzed at high temperature.
Environmental geochemistry and health, 39(2), 403-415.
Johnson, J., Collins, T., Power, A., Chandra, S., Skylas, D., Portman, D., Panozzo, J.,
Blanchard, C., & Naiker, M. (2020). Antioxidative properties and macrochemical
composition of five commercial mungbean varieties in Australia. Legume Science,
2(1), e27. doi:https://doi.org/10.1002/leg3.27
Johnston, C. T. (2017). 18 Biochar analysis by Fourier-transform infra-red spectroscopy. In
Biochar: A Guide to Analytical Methods (pp. 199).
Joseph, S., Camps-Arbestain, M., Lin, Y., Munroe, P., Chia, C., Hook, J., Van Zwieten, L.,
Kimber, S., Cowie, A., & Singh, B. (2010). An investigation into the reactions of
biochar in soil. Soil Research, 48(7), 501-515.
Joseph, S., Kammann, C. I., Shepherd, J. G., Conte, P., Schmidt, H.-P., Hagemann, N., Rich,
A. M., Marjo, C. E., Allen, J., Munroe, P., Mitchell, D. R. G., Donne, S., Spokas, K.,
& Graber, E. R. (2018a). Microstructural and associated chemical changes during the
composting of a high temperature biochar: Mechanisms for nitrate, phosphate and
other nutrient retention and release. Science of The Total Environment, 618, 1210-
1223. doi:https://doi.org/10.1016/j.scitotenv.2017.09.200
Joseph, S., Taylor, P., & Cowie, A. (2018b). Choosing a biochar reactor to meet your needs.
Biochar For Sustainable Soils. Retrieved from
https://biochar.international/guides/biochar-reactor-to-meet-needs/
Joseph, S. D., & Munroe, P. R. (2017). 23 Application of scanning electron microscopy to the
analysis of biochar-related materials. In Biochar: A Guide to Analytical Methods (pp.
228).
Kaia, T., Gross, K., Plūduma, L., & Veiderma, M. (2011). A Review on the Thermal Stability
of Calcium Apatites. Journal of Thermal Analysis and Calorimetry, 110.
doi:10.1007/s10973-011-1877-y
Kalus, K., Koziel, J. A., & Opaliński, S. (2019). A Review of Biochar Properties and Their
Utilization in Crop Agriculture and Livestock Production. Applied Sciences, 9(17),
3494.
Karhu, K., Mattila, T., Bergström, I., & Regina, K. (2011). Biochar addition to agricultural
soil increased CH4 uptake and water holding capacity–Results from a short-term pilot
field study. Agriculture, Ecosystem and Environment, 140(1-2), 309-313.
Kaszuba, M., Corbett, J., Watson, F. M., & Jones, A. (2010). High-concentration zeta
potential measurements using light-scattering techniques. Philosophical transactions.
Series A, Mathematical, physical, and engineering sciences, 368(1927), 4439-4451.
doi:10.1098/rsta.2010.0175
Keiluweit, M., Kleber, M., Sparrow, M. A., Simoneit, B. R. T., & Prahl, F. G. (2012).
Solvent-Extractable Polycyclic Aromatic Hydrocarbons in Biochar: Influence of
Pyrolysis Temperature and Feedstock. Environmental Science & Technology, 46(17),
9333-9341. doi:10.1021/es302125k
Khanmohammadi, Z., Majid, A., & Mosaddeghi, M. R. (2015). Effect of pyrolysis
temperature on chemical and physical properties of sewage sludge biochar. Waste

104
management & research : the journal of the International Solid Wastes and Public
Cleansing Association, ISWA, 33. doi:10.1177/0734242X14565210
Kim, K.-H., Jahan, S. A., Kabir, E., & Brown, R. J. C. (2013). A review of airborne
polycyclic aromatic hydrocarbons (PAHs) and their human health effects.
Environment International, 60, 71-80.
doi:https://doi.org/10.1016/j.envint.2013.07.019
Kloss, S., Zehetner, F., Dellantonio, A., Hamid, R., Ottner, F., Liedtke, V., Schwanninger,
M., Gerzabek, M. H., & Soja, G. (2012). Characterization of slow pyrolysis biochars:
effects of feedstocks and pyrolysis temperature on biochar properties. Journal of
environmental quality, 41(4), 990-1000.
Kutlu, H. R., Ünsal, I., & Görgülü, M. (2001). Effects of providing dietary wood (oak)
charcoal to broiler chicks and laying hens. Animal Feed Science Technology, 90(3-4),
213-226.
Lao, E., & Mbega, E. (2020). BIOCHAR AS A FEED ADDITIVE FOR IMPROVING THE
PERFORMANCE OF FARM ANIMALS. Malaysian Journal of Sustainable
Agriculture, 4, 86-93. doi:10.26480/mjsa.02.2020.86.93
Lazzari, E., Schena, T., Marcelo, M. C. A., Primaz, C. T., Silva, A. N., Ferrão, M. F., Bjerk,
T., & Caramão, E. B. (2018). Classification of biomass through their pyrolytic bio-oil
composition using FTIR and PCA analysis. Industrial Crops and Products, 111, 856-
864. doi:https://doi.org/10.1016/j.indcrop.2017.11.005
Lee, Y., Park, J., Ryu, C., Gang, K. S., Yang, W., Park, Y.-K., Jung, J., & Hyun, S. (2013).
Comparison of biochar properties from biomass residues produced by slow pyrolysis
at 500°C. Bioresource Technology, 148, 196-201.
doi:https://doi.org/10.1016/j.biortech.2013.08.135
Lehmann, J., Gaunt, J., & Rondon, M. (2006). Bio-char sequestration in terrestrial
ecosystems–a review. Mitigation adaptation strategies for global change, 11(2), 403-
427.
Lehmann, J., & Joseph, S. (2009). Biochar for environmental management: an introduction
(Vol. 1).
Lehmann, J., & Joseph, S. (2015a). Biochar for environmental management: an introduction.
In Biochar for environmental management (pp. 33-46): Routledge.
Lehmann, J., & Joseph, S. (2015b). Biochar for environmental management: science,
technology and implementation: Routledge.
Leng, L., Xiong, Q., Yang, L., Li, H., Zhou, Y., Zhang, W., Jiang, S., Li, H., & Huang, H.
(2021). An overview on engineering the surface area and porosity of biochar. Science
of The Total Environment, 763, 144204.
doi:https://doi.org/10.1016/j.scitotenv.2020.144204
Li, A. M., Li, X. D., Li, S. Q., Ren, Y., Chi, Y., Yan, J. H., & Cen, K. F. (1999). Pyrolysis of
solid waste in a rotary kiln: influence of final pyrolysis temperature on the pyrolysis
products. Journal of Analytical and Applied Pyrolysis, 50(2), 149-162.
doi:https://doi.org/10.1016/S0165-2370(99)00025-X

105
Li, S.-Q., Yao, Q., Chi, Y., Yan, J.-H., & Cen, K.-F. (2004). Pilot-scale pyrolysis of scrap
tires in a continuous rotary kiln reactor. Industrial & Engineering Chemistry
Research, 43(17), 5133-5145.
Li, S., Chan, C. Y., Sharbatmaleki, M., Trejo, H., & Delagah, S. (2020). Engineered Biochar
Production and Its Potential Benefits in a Closed-Loop Water-Reuse Agriculture
System. Water, 12(10), 2847.
Liang, H., Chen, L., & Liu, G. (2016). Surface morphology properties of biochars produced
from different feedstocks. Paper presented at the International Conference on Civil,
Transportation and Environment (ICCTE 2016).
Liu, W.-J., Li, W.-W., Jiang, H., & Yu, H.-Q. (2017). Fates of Chemical Elements in
Biomass during Its Pyrolysis. Chemical Reviews, 117(9), 6367-6398.
doi:10.1021/acs.chemrev.6b00647
Lyu, H., Gao, B., He, F., Zimmerman, A. R., Ding, C., Huang, H., & Tang, J. (2018). Effects
of ball milling on the physicochemical and sorptive properties of biochar:
Experimental observations and governing mechanisms. Environmental Pollution, 233,
54-63.
Makavana, J., Sarsavadia, P., & Chauhan, P. (2020). Effect of Pyrolysis Temperature and
Residence Time on Bio-char Obtained from Pyrolysis of Shredded Cotton Stalk.
International Research Journal of Pure and Applied Chemistry, 10-28.
Maliutina, K., Tahmasebi, A., Yu, J., & Saltykov, S. N. (2017). Comparative study on flash
pyrolysis characteristics of microalgal and lignocellulosic biomass in entrained-flow
reactor. Energy Conversion and Management, 151, 426-438.
doi:https://doi.org/10.1016/j.enconman.2017.09.013
Malucelli, L. C., Silvestre, G. F., Carneiro, J., Vasconcelos, E. C., Guiotoku, M., Maia, C. M.
B. F., & Fil, M. A. S. C. (2019). Biochar higher heating value estimative using
thermogravimetric analysis. In Journal of Thermal Analysis and Calorimetry:
Springer.
Man, K. Y., Chow, K. L., Man, Y. B., Mo, W. Y., & Wong, M. H. (2020). Use of biochar as
feed supplements for animal farming. Critical Reviews in Environmental Science and
Technology, 1-31. doi:10.1080/10643389.2020.1721980
Manariotis, I., Fotopoulou, K., & Karapanagioti, H. (2015). Preparation and Characterization
of Biochar Sorbents Produced from Malt Spent Rootlets. Industrial & Engineering
Chemistry Research, 54, 150916122648008. doi:10.1021/acs.iecr.5b02698
Mangold, E. (1936). Die Verdaulichkeit der Futtermittel in ihrer Abhängigkeit von
verschiedenen Einflüssen. Forschungsdienst—Reichsarbeitsgemeinschaften d.
Landwirtschaftswissenschaft, 1, 862-867.
Marrero, A. (2018). What Happens To Biochar Right AFter It Is Made? Wakefield Biochar.
Retrieved from https://www.wakefieldbiochar.com/biochar-after-pyrolysis/
Mašek, O., Brownsort, P., Cross, A., & Sohi, S. (2013). Influence of production conditions
on the yield and environmental stability of biochar. Fuel, 103, 151-155.
Mazengarb, M. (2021). Australia’s de-facto carbon price at record high, may reach $50 per
tonne. RenewEconomy.

106
Melas, G. B. (2014). Interactions between different types of biochar and soil microbial
activity: the effects on the dynamics of labile organic matter and the behaviour of
some pesticides: Universitat Autònoma de Barcelona.
Meng, J., Feng, X., Dai, Z., Liu, X., Wu, J., & Xu, J. (2014). Adsorption characteristics of
Cu(II) from aqueous solution onto biochar derived from swine manure.
Environmental Science and Pollution Research, 21(11), 7035-7046.
doi:10.1007/s11356-014-2627-z
Mia, S., Singh, B., & Dijkstra, F. A. (2017). Aged biochar affects gross nitrogen
mineralization and recovery: a 15N study in two contrasting soils. GCB Bioenergy,
9(7), 1196-1206.
Mierzwa-Hersztek, M., Gondek, K., Jewiarz, M., & Dziedzic, K. (2019). Assessment of
energy parameters of biomass and biochars, leachability of heavy metals and
phytotoxicity of their ashes. Journal of Material Cycles and Waste Management,
21(4), 786-800.
Mitchell, P., Dalley, T., & Helleur, R. (2013). Preliminary laboratory production and
characterization of biochars from lignocellulosic municipal waste. Journal of
Analytical and Applied Pyrolysis, 99, 71–78. doi:10.1016/j.jaap.2012.10.025
Mohan, D., Pittman Jr, C. U., & Steele, P. H. (2006). Pyrolysis of wood/biomass for bio-oil: a
critical review. Energy & Fuels, 20(3), 848-889.
Mohd Hasan, M. H., Bachmann, R. T., Loh, S. K., Manroshan, S., & Ong, S. K. (2019).
Effect of Pyrolysis Temperature and Time on Properties of Palm Kernel Shell-Based
Biochar. IOP Conference Series: Materials Science and Engineering, 548, 012020.
doi:10.1088/1757-899x/548/1/012020
Morgan, T. J., Turn, S. Q., & Anthe, G. (2015). Fast Pyrolysis Behavior of Banagrass as a
Function of Temperature and Volatiles Residence Time in a Fluidized Bed Reactor.
Plos one, 10(8). doi:http://dx.doi.org/10.1371/journal.pone.0136511
Mui, N. T., & Ledin, I. (2006). Effect of method of processing foliage of Acacia mangium
and inclusion of bamboo charcoal in the diet on performance of growing goats.
Animal Feed Science Technology, 130(3-4), 242-256.
Mukherjee, A., Zimmerman, A., & Harris, W. (2011). Surface chemistry variations among a
series of laboratory-produced biochars. Geoderma, 163(3-4), 247-255.
Mullen, M. D. (2005). Phosphorus in Soils- Biological Interactions. Encyclopedia of Soils in
the Environment, 3, 210-216. doi:10.1016/B0-12-348530-4/00161-2
Munera, E. J. L., Martinsen, V., Strand, L. T., Zivanovic, V., Cornelissen, G., & Mulder, J.
(2018). Cation exchange capacity of biochar: An urgent method modification. Science
of The Total Environment, 642, 190-197.
doi:https://doi.org/10.1016/j.scitotenv.2018.06.017
Munera, J., Martinsen, V., Mulder, J., Tau Strand, L., & Cornelissen, G. (2017). Cation
Exchange Capacity of Biochar: An urgent method modification. Paper presented at
the EGU General Assembly Conference Abstracts.
Nam, H., Capareda, S. C., Ashwath, N., & Kongkasawan, J. (2015). Experimental
investigation of pyrolysis of rice straw using bench-scale auger, batch and fluidized
bed reactors. Energy, 93, 2384-2394.
107
Neves, E., Petersen, J., Bartone, R., & Silva, C. (2003). Historical and socio-cultural origins
of Amazonian dark earth In: Lehmann J, Kern DC, Glaser B, Woods WI, editors.
Amazonian dark earths: origin, properties, management. In: Dordrecht: Kluwer.
O’Toole, A., Andersson, D., Gerlach, A., Glaser, B., Kammann, C., Kern, J., Kuoppamäki,
K., Mumme, J., Schmidt, H.-P., & Schulze, M. (2016). Current and future
applications for biochar. In G. R. Simon Shackley, Kor Zwart, Bruno Glaser (Ed.),
Biochar in European Soils and Agriculture: Science and Practice (pp. 253-280):
Abington: Taylor & Francis.
Oni, B. A., Oziegbe, O., & Olawole, O. O. (2019). Significance of biochar application to the
environment and economy. Annals of Agricultural Sciences, 64(2), 222-236.
doi:https://doi.org/10.1016/j.aoas.2019.12.006
Park, J. H., Choppala, G. K., Bolan, N. S., Chung, J. W., & Chuasavathi, T. (2011). Biochar
reduces the bioavailability and phytotoxicity of heavy metals. Plant soil, 348(1-2),
439.
Patel, S., Kundu, S., Halder, P., Veluswamy, G., Pramanik, B., Paz-Ferreiro, J., Surapaneni,
A., & Shah, K. (2019). Slow pyrolysis of biosolids in a bubbling fluidised bed reactor
using biochar, activated char and lime. Journal of Analytical and Applied Pyrolysis,
144, 104697.
Peacocke, G. V. C., & Bridgwater, A. V. (1994). Ablative plate pyrolysis of biomass for
liquids. Biomass and Bioenergy, 7(1), 147-154. doi:https://doi.org/10.1016/0961-
9534(94)00054-W
Peng, H., Gao, P., Chu, G., Pan, B., Peng, J., & Xing, B. (2017). Enhanced adsorption of
Cu(II) and Cd(II) by phosphoric acid-modified biochars. Environ Pollut, 229, 846-
853. doi:10.1016/j.envpol.2017.07.004
Phillips, C. L., Meyer, K. M., Garcia-Jaramillo, M., Weidman, C. S., Stewart, C. E., Wanzek,
T., Grusak, M. A., Watts, D. W., Novak, J., & Trippe, K. M. (2022). Towards
predicting biochar impacts on plant-available soil nitrogen content. Biochar, 4(1), 9.
doi:10.1007/s42773-022-00137-2
Prasai, T., Walsh, K., Midmore, D., & Bhattarai, S. (2018). Effect of biochar, zeolite and
bentonite feed supplements on egg yield and excreta attributes. Animal Production
Science, 58(9), 1632-1641.
Qambrani, N. A., Rahman, M. M., Won, S., Shim, S., & Ra, C. (2017). Biochar properties
and eco-friendly applications for climate change mitigation, waste management, and
wastewater treatment: A review. Renewable and Sustainable Energy Reviews, 79,
255-273.
Qi, F., Yan, Y., Lamb, D., Naidu, R., Bolan, N. S., Liu, Y., Ok, Y. S., Donne, S. W., &
Semple, K. T. (2017). Thermal stability of biochar and its effects on cadmium
sorption capacity. Bioresource Technology, 246, 48-56.
Quilliam, R. S., Rangecroft, S., Emmett, B. A., Deluca, T. H., & Jones, D. L. (2013). Is
biochar a source or sink for polycyclic aromatic hydrocarbon (PAH) compounds in
agricultural soils? GCB Bioenergy, 5(2), 96-103.
doi:https://doi.org/10.1111/gcbb.12007

108
Rajapaksha, A. U., Vithanage, M., Ahmad, M., Seo, D.-C., Cho, J.-S., Lee, S.-E., Lee, S. S.,
& Ok, Y. S. (2015). Enhanced sulfamethazine removal by steam-activated invasive
plant-derived biochar. Journal of Hazardous Materials, 290, 43-50.
Rajkovich, S., Enders, A., Hanley, K., Hyland, C., Zimmerman, A., & Lehmann, J. (2011).
Corn growth and nitrogen nutrition after additions of biochars with varying properties
to a temperate soil. Biology and Fertility of Soils, 48. doi:10.1007/s00374-011-0624-7
Rayment, G., & Higginson, F. R. (1992). Australian laboratory handbook of soil and water
chemical methods: Inkata Press Pty Ltd.
Raza, M., Inayat, A., Ahmed, A., Jamil, F., Ghenai, C., Naqvi, S. R., Shanableh, A., Ayoub,
M., Waris, A., & Park, Y.-K. (2021). Progress of the Pyrolyzer Reactors and
Advanced Technologies for Biomass Pyrolysis Processing. Sustainability, 13(19),
11061.
Rehrah, D., Reddy, M., Novak, J., Bansode, R., Schimmel, K. A., Yu, J., Watts, D., &
Ahmedna, M. (2014). Production and characterization of biochars from agricultural
by-products for use in soil quality enhancement. Journal of Analytical and Applied
Pyrolysis, 108, 301-309.
Rippy, J. F. M., & Nelson, P. V. (2007). Cation Exchange Capacity and Base Saturation
Variation among Alberta, Canada, Moss Peats. Hortscience, 42, 349-352.
Robb, S., & Joseph, S. (2019). A Report on the Value of Biochar and Wood Vinegar:
Practical Experience of Users in Australia and New Zealand. Retrieved November 24,
2019. In.
Rombolà, A. G., Fabbri, D., Baronti, S., Vaccari, F. P., Genesio, L., & Miglietta, F. (2019).
Changes in the pattern of polycyclic aromatic hydrocarbons in soil treated with
biochar from a multiyear field experiment. Chemosphere, 219, 662-670.
doi:https://doi.org/10.1016/j.chemosphere.2018.11.178
Ronewicz, K., Kluska, J., Heda, Ł., & Kardaś, D. (2017). Chemical and Physical Properties
of Pine Wood during Pyrolysis. Drvna industrija, 68, 29-36.
doi:10.5552/drind.2017.1617
Ronsse, F., Van Hecke, S., Dickinson, D., & Prins, W. (2013). Production and
characterization of slow pyrolysis biochar: influence of feedstock type and pyrolysis
conditions. Global Change Biology Bioenergy, 5(2), 104-115.
Schmidt, H.-P., Hagemann, N., Draper, K., & Kammann, C. (2019). The use of biochar in
animal feeding. Peer J, 7, e7373.
Schmidt, H.-P., & Wilson, K. (2014). 55 uses of biochar. The Biochar Journal, 286-288.
Schmidt, H. P., Bucheli, T., Kammann, C., Glaser, B., Abiven, S., & Leifeld, J. (2016).
European Biochar Certificate - Guidelines for a sustainable production of biochar.
European Biochar Foundation.
Schmidt, H. P., Taylor, P., Eglise, A., & Arbaz, C. (2014). Kon-Tiki flame curtain pyrolysis
for the democratization of biochar production. Biochar J, 1, 14-24.
Shaaban, A., Se, S.-M., Dimin, M. F., Juoi, J. M., Mohd Husin, M. H., & Mitan, N. M. M.
(2014). Influence of heating temperature and holding time on biochars derived from

109
rubber wood sawdust via slow pyrolysis. Journal of Analytical and Applied Pyrolysis,
107, 31-39. doi:https://doi.org/10.1016/j.jaap.2014.01.021
Shaw, V. E. (1959). Extraction of rare-earth elements from bastnaesite concentrate (Vol.
5474): US Department of the Interior, Bureau of Mines.
Shetty, R., & Prakash, N. B. (2020). Effect of different biochars on acid soil and growth
parameters of rice plants under aluminium toxicity. Scientific reports, 10(1), 12249.
doi:10.1038/s41598-020-69262-x
Siddiqui, M. T. H., Nizamuddin, S., Mubarak, N. M., Shirin, K., Aijaz, M., Hussain, M., &
Baloch, H. A. (2019). Characterization and Process Optimization of Biochar Produced
Using Novel Biomass, Waste Pomegranate Peel: A Response Surface Methodology
Approach. Waste Mass Valorization, 10(3), 521-532. doi:10.1007/s12649-017-0091-y
Siedlecki, M., Jong, W., & Verkooijen, A. H. M. (2011). Fluidized Bed Gasification as a
Mature And Reliable Technology for the Production of Bio-Syngas and Applied in
the Production of Liquid Transportation Fuels—A Review. Energies, 4.
doi:10.3390/en4030389
Sigmund, G., Hüffer, T., Hofmann, T., & Kah, M. (2017). Biochar total surface area and total
pore volume determined by N2 and CO2 physisorption are strongly influenced by
degassing temperature. Science of The Total Environment, 580, 770-775.
doi:https://doi.org/10.1016/j.scitotenv.2016.12.023
Singh, B., Camps-Arbestain, M., & Lehmann, J. (2017a). Biochar: a guide to analytical
methods: Csiro Publishing.
Singh, B., Dolk, M. M., Shen, Q., & Camps-Arbestain, M. (2017b). Biochar pH, electrical
conductivity and liming potential. In Biochar: A Guide to Analytical Methods (Vol.
23): Csiro Publishing, Clayton, Australia.
Singh, B., Singh, B. P., & Cowie, A. L. (2010a). Characterisation and evaluation of biochars
for their application as a soil amendment Soil Research, 48(7), 516-525.
doi:https://doi.org/10.1071/SR10058
Singh, B. P., Hatton, B. J., Balwant, S., Cowie, A. L., & Kathuria, A. (2010b). Influence of
biochars on nitrous oxide emission and nitrogen leaching from two contrasting soils.
Journal of environmental quality, 39(4), 1224-1235. doi:10.2134/jeq2009.0138
Smolker, R. (2015). Biochar: Black Gold or Just Another Snake Oil Scheme? . Retrieved
from
http://www.earthisland.org/journal/index.php/elist/eListRead/biochar_black_gold_or_
just_another_snake _oil_scheme/
SOFT, A. (2012). SOFT AGRICULTURE. Retrieved from
https://www.softagriculture.com.au/
Sohi, S. P., Krull, E., Lopez-Capel, E., & Bol, R. (2010). A review of biochar and its use and
function in soil. In Advances in agronomy (Vol. 105, pp. 47-82): Elsevier.
Spokas, K. A. (2010). Review of the stability of biochar in soils: predictability of O:C molar
ratios. Carbon Management, 1(2), 289-303. doi:10.4155/cmt.10.32
Strezov, V., & Evans, T. J. (2014). Biomass processing technologies: CRC Press.

110
Suliman, W., Harsh, J., Abu-Lail, N., Fortuna, A.-M., Dallmeyer, I., & Garcia-Perez, M.
(2016). Influence of feedstock source and pyrolysis temperature on biochar bulk and
surface properties. Biomass and Bioenergy, 84, 37-48.
doi:10.1016/j.biombioe.2015.11.010
Suman, S. (2020). Conversion of Solid Biomass into Biochar: Act as a Green, Eco-Friendly
Energy Source and a Substitute of Fossil Fuel Inputs. Proceedings, 58, 34.
doi:10.3390/WEF-06916
Sun, J., He, F., Pan, Y., & Zhang, Z. (2017a). Effects of pyrolysis temperature and residence
time on physicochemical properties of different biochar types. Acta Agriculturae
Scandinavica, Section B—Soil & Plant Science, 67(1), 12-22.
Sun, T., Levin, B. D. A., Guzman, J. J. L., Enders, A., Muller, D. A., Angenent, L. T., &
Lehmann, J. (2017b). Rapid electron transfer by the carbon matrix in natural
pyrogenic carbon. Nature Communications, 8(1), 14873. doi:10.1038/ncomms14873
Tag, A. T., Duman, G., Ucar, S., & Yanik, J. (2016). Effects of feedstock type and pyrolysis
temperature on potential applications of biochar. Journal of Analytical and Applied
Pyrolysis, 120, 200-206.
Tanaka, H., Chikazawa, M., Kandori, K., & Ishikawa, T. (2000). Influence of thermal
treatment on the structure of calcium hydroxyapatite. Physical Chemistry Chemical
Physics, 2(11), 2647-2650. doi:10.1039/B001877P
Tchapda, A. H., & Pisupati, S. V. (2015). Characterization of an entrained flow reactor for
pyrolysis of coal and biomass at higher temperatures. Fuel, 156, 254-266.
doi:https://doi.org/10.1016/j.fuel.2015.04.015
Teoh, R., Caro, E., Holman, D. B., Joseph, S., Meale, S. J., & Chaves, A. V. (2019). Effects
of Hardwood Biochar on Methane Production, Fermentation Characteristics, and the
Rumen Microbiota Using Rumen Simulation. Frontiers in Microbiology, 10(1534).
doi:10.3389/fmicb.2019.01534
Thomas, J. R., & Fucher, F. (2000). Thermal Modeling of Microwave Heated Packed and
Fluidized Bed Catalytic Reactors. Journal of Microwave Power and Electromagnetic
Energy, 35(3), 165-174. doi:10.1080/08327823.2000.11688433
Tomczyk, A., Sokołowska, Z., & Boguta, P. (2020). Biochar physicochemical properties:
pyrolysis temperature and feedstock kind effects. Reviews in Environmental Science
and Bio/Technology, 1-25.
Totusek, R., & Beeson, W. (1953). The nutritive value of wood charcoal for pigs. Journal of
Animal Science, 12(2), 271-281.
Trajano, H. L., DeMartini, J. D., Studer, M. H., & Wyman, C. E. (2013). Comparison of the
effectiveness of a fluidized sand bath and a steam chamber for reactor heating.
Industrial & Engineering Chemistry Research, 52(13), 4932-4938.
Tripathi, M., Sahu, J. N., & Ganesan, P. (2016). Effect of process parameters on production
of biochar from biomass waste through pyrolysis: A review. Renewable and
Sustainable Energy Reviews, 55, 467-481.
Tsai, W.-T., Liu, S.-C., Chen, H.-R., Chang, Y.-M., & Tsai, Y.-L. (2012). Textural and
chemical properties of swine-manure-derived biochar pertinent to its potential use as a

111
soil amendment. Chemosphere, 89(2), 198-203.
doi:https://doi.org/10.1016/j.chemosphere.2012.05.085
Uchimiya, M., Wartelle, L., Klasson, K., Fortier, C., & Lima, I. (2011). Influence of
Pyrolysis Temperature on Biochar Property and Function as a Heavy Metal Sorbent in
Soil. Journal of Agricultural and Food Chemistry, 59, 2501-2510.
doi:10.1021/jf104206c
Usevičiūtė, L., & Baltrėnaitė-Gedienė, E. (2020). Dependence of pyrolysis temperature and
lignocellulosic physical-chemical properties of biochar on its wettability. Biomass
Conversion and Biorefinery. doi:10.1007/s13399-020-00711-3
Wallace, C. A., Afzal, M. T., & Saha, G. C. (2019). Effect of feedstock and microwave
pyrolysis temperature on physio-chemical and nano-scale mechanical properties of
biochar. Bioresources and Bioprocessing, 6(1), 33. doi:10.1186/s40643-019-0268-2
Wang, C., Wang, Y., & Herath, H. M. S. K. (2017). Polycyclic aromatic hydrocarbons
(PAHs) in biochar – Their formation, occurrence and analysis: A review. Organic
Geochemistry, 114, 1-11. doi:https://doi.org/10.1016/j.orggeochem.2017.09.001
Wang, K., Brown, R. C., Homsy, S., Martinez, L., & Sidhu, S. S. (2013a). Fast pyrolysis of
microalgae remnants in a fluidized bed reactor for bio-oil and biochar production.
Bioresource Technology, 127, 494-499.
Wang, S., Jiang, X., Han, X., & Tong, J. (2013b). EFFECT OF RESIDENCE TIME ON
PRODUCTS YIELD AND CHARACTERISTICS OF SHALE OIL AND GASES
PRODUCED BY LOW-TEMPERATURE RETORTING OF DACHENGZI OIL
SHALE. Oil Shale, 30(4), 501-516.
Wardani, S., Pranoto, & Himawanto, D. A. (2018). Kinetic parameters and calorific value of
biochar from mahogany (Swietenia macrophylla King) wood pyrolysis with heating
rate and final temperature variations. Paper presented at the AIP Conference
Proceedings.
Wayne, E. (2012). Conquistadors, cannibals and climate change: A brief history of biochar.
Pro-Natura International, 11, 2013.
Winter, J. D. (2018). Method and Apparatus for processing Carbonaceous Material.
Wu, W., Yang, M., Feng, Q., McGrouther, K., Wang, H., Lu, H., & Chen, Y. (2012).
Chemical characterization of rice straw-derived biochar for soil amendment. Biomass
and Bioenergy, 47, 268-276. doi:https://doi.org/10.1016/j.biombioe.2012.09.034
Yadav, K., & Jagadevan, S. (2019). Influence of Process Parameters on Synthesis of Biochar
by Pyrolysis of Biomass: An Alternative Source of Energy. In Pyrolysis. London:
IntechOpen.
Yang, W., Shang, J., Li, B., & Flury, M. (2020). Surface and colloid properties of biochar and
implications for transport in porous media. Critical Reviews in Environmental Science
and Technology, 50(23), 2484-2522. doi:10.1080/10643389.2019.1699381
Yang, X., Wang, H., Strong, P. J., Xu, S., Liu, S., Lu, K., Sheng, K., Guo, J., Che, L., & He,
L. (2017). Thermal properties of biochars derived from waste biomass generated by
agricultural and forestry sectors. Energies, 10(4), 469.

112
Yargicoglu, E. N., Sadasivam, B. Y., Reddy, K. R., & Spokas, K. (2015). Physical and
chemical characterization of waste wood derived biochars. Waste Management, 36,
256-268.
Yeo, J. Y., Chin, B. L. F., Tan, J. K., & Loh, Y. S. (2019). Comparative studies on the
pyrolysis of cellulose, hemicellulose, and lignin based on combined kinetics. Journal
of the Energy Institute, 92(1), 27-37. doi:https://doi.org/10.1016/j.joei.2017.12.003
Young, P., Lawrence, J., Batista, R., Jensen-Fellows, A., Richard, B., & Sheridan, T. (2019).
Biochar Market Profile Report.
Yu, O.-Y., Raichle, B., & Sink, S. (2013). Impact of biochar on the water holding capacity of
loamy sand soil. International Journal of Energy and Environmental Engineering,
4(1), 44.
Yuan, H., Lu, T., Huang, H., Zhao, D., Kobayashi, N., & Chen, Y. (2015). Influence of
pyrolysis temperature on physical and chemical properties of biochar made from
sewage sludge. Journal of Analytical and Applied Pyrolysis, 112, 284-289.
doi:https://doi.org/10.1016/j.jaap.2015.01.010
Yuan, J.-H., Xu, R.-K., & Zhang, H. (2011). The forms of alkalis in the biochar produced
from crop residues at different temperatures. Bioresource Technology, 102(3), 3488-
3497.
Zama, E. F., Zhu, Y.-G., Reid, B. J., & Sun, G.-X. (2017). The role of biochar properties in
influencing the sorption and desorption of Pb (II), Cd (II) and As (III) in aqueous
solution. Journal of cleaner production, 148, 127-136.
Zapata, F., & Roy, R. N. (2004). Use of phosphate rocks for sustainable agriculture. FAO
Fertilizer and Plant Nutrition Bulletin, 1-148.
Zellner, T., Prasa, D., Färber, E., Hoffmann-Walbeck, P., Genser, D., & Eyer, F. (2019). The
Use of activated charcoal to Treat intoxications. Deutsches Ärzteblatt International,
116(18), 311.
Zhang, X., Wang, H., He, L., Lu, K., Sarmah, A., Li, J., Bolan, N. S., Pei, J., & Huang, H.
(2013). Using biochar for remediation of soils contaminated with heavy metals and
organic pollutants. Environmental Science and Pollution Research, 20(12), 8472-
8483.
Zhang, Y., Zhao, L., Guo, R., Song, N., Wang, J., Cao, Y., Orndorff, W., & Pan, W.-p.
(2015). Mercury adsorption characteristics of HBr-modified fly ash in an entrained-
flow reactor. Journal of Environmental Sciences, 33, 156-162.
doi:https://doi.org/10.1016/j.jes.2015.01.011
Zhao, B., Xu, H., Zhang, T., Nan, X., & Ma, F. (2018). Effect of pyrolysis temperature on
sulfur content, extractable fraction and release of sulfate in corn straw biochar. RSC
advances, 8(62), 35611-35617. doi:10.1039/C8RA06382F
Zhao, S.-X., Ta, N., & Wang, X.-D. (2017). Effect of temperature on the structural and
physicochemical properties of biochar with apple tree branches as feedstock material.
Energies, 10(9), 1293.
Zhu, L., Lei, H., Zhang, Y., Zhang, X., Bu, Q., & Wei, Y. (2018). A review of biochar
derived from pyrolysis and its application in biofuel production. SF Journal of
Material and Chemical Engineering, 1007.
113
Zimmerman, A. R., Gao, B., & Ahn, M.-Y. (2011). Positive and negative carbon
mineralization priming effects among a variety of biochar-amended soils. Soil Biology
and Biochemistry, 43(6), 1169-1179. doi:https://doi.org/10.1016/j.soilbio.2011.02.005

114

You might also like