Environmental Engineering
Environmental Engineering
DOCTOR OF PHILOSOPHY
(Environmental Engineering and Management)
2023
Nutrients and Energy Recovery from Ecological Sanitation System
2023
DECLARATION
This thesis is my original work and has not been presented for a degree in any other
university.
Signature…………………………………………………..Date……………………….
Austine Owuor Otieno
This thesis has been submitted for examination with our approval as University
supervisors.
Signature…………………………………………………..Date……………………….
Prof. Patrick G. Home, PhD
JKUAT, Kenya
Signature…………………………………………………..Date……………………….
Prof. (Eng). James M. Raude, PhD
JKUAT, Kenya
Signature…………………………………………………..Date……………………….
Dr. Sylvia I. Murunga, PhD
JKUAT, Kenya
ii
DEDICATION
I dedicate this work to my father Mr. Elias Otieno Ongowo, and mother Mrs. Grace
Akinyi Otieno for motivating me to achieve the best in my academic pursuit. I also
dedicate this work to my family for their encouragement and enduring my absence
during my studies.
iii
ACKNOWLEDGEMENT
iv
TABLE OF CONTENTS
DECLARATION ............................................................................................................. ii
ACKNOWLEDGEMENT ............................................................................................. iv
INTRODUCTION ........................................................................................................... 1
v
1.5 Justification of the Study ......................................................................................... 7
LITERATURE REVIEW............................................................................................. 10
2.2.1 Urine Diversion and Dehydration Toilets and Associated Problems with Urine
Fertilizer Recovery ....................................................................................... 12
2.3 Technologies for Nitrogen and Phosphorous Recovery from Human Urine ....... 13
vi
2.6 Pineapple Peel Wastes and Application of Pineapple Peel Biochar as an Adsorbent
............................................................................................................................. 27
2.12 Higher Heating Value , Proximate and Ultimate Analysis of Briquettes ............ 36
vii
2.12.3 Ultimate Analysis ......................................................................................... 38
3.1 Determining the Behaviour and Capacity of Pineapple Peel Biochar and Lateritic
Soil in Adsorbing Ammonium Nitrogen and Phosphorous from Human Urine . 45
3.2 Optimising the Adsorption of Ammonium Nitrogen and Phosphorous from Human
Urine on Pineapple Peel Biochar and Lateritic Soil ............................................ 49
viii
3.3 Determining the Thermal Degradation and Kinetic Behaviour of Pyrolyzed
Human Faeces, Sawdust Char and their Blend ................................................... 52
4.1 Behaviour and Capacity of Pineapple Peel Biochar and Lateritic Soil in Adsorbing
Ammonium Nitrogen and Phosphorous from Human Urine .............................. 60
4.1.1 Composition and Structures of Lateritic Soil and Pineapple Peel Biochar ... 60
ix
4.1.2 The Effect of Contact Time and the Initial Concentration on Adsorption of
Ammonium Nitrogen from Human Urine on Lateritic Soil and Pineapple
Peel Biochar ................................................................................................. 66
4.1.4 The Effect of Contact Time and the Initial Concentration on Adsorption of
Phosphorous from Human Urine on Lateritic Soil and Pineapple Peel
Biochar ......................................................................................................... 72
4.1.5 Adsorption Isotherms for the Adsorption of Phosphorous from Human Urine
on the Lateritic Soil ...................................................................................... 75
x
4.3.1 Morphological, Physicochemical Properties, and Thermal Behaviors of Faecal
Char, Sawdust Char, and their Blend ........................................................... 86
4.4.1 Proximate Analysis Results and Heating Values of Briquettes and Charcoal 98
4.4.2 Elemental Composition of Human Faecal Char, Sawdust Char, and Molasses
.................................................................................................................... 100
4.4.4 Variation of Flue Gases Temperatures with Combustion Time ................... 104
xi
5.2.1 Recommendations for Application ............................................................... 110
xii
LIST OF TABLES
Table 2.1: Similarities and Differences of CCC, CCI, and CCF .................................... 20
Table 3.1: Parameters and Levels for the Experimental Design ..................................... 51
Table 4.1: Nitrogen Sorption Properties of Lateritic Soil and Pineapple Peel Biochar . 62
Table 4.4: Langmuir, Freundlich, and D-R Isotherm Parameters for the Adsorption of
Phosphorous on Lateritic Soil ....................................................................... 77
Table 4.6: Ammonium Nitrogen Adsorption Results Based on the Response Surface
Methodology and the Central Composite Circumscribed Experimental
Design............................................................................................................ 79
Table 4.7: ANOVA for Response Surface Quadratic Model for Optimised Adsorption
of Ammonium Nitrogen from Human Urine on Pineapple Peel Biochar ..... 80
Table 4.9: ANOVA for Response Surface Quadratic Model for Optimised Adsorption
of Phosphorous from Human Urine on Lateritic Soil ................................... 84
xiii
Table 4.10: Proximate and Ultimate Analysis of Faecal Char and Sawdust Char .......... 87
Table 4.11: Activation Energies (Eα (kJ/mol)) Estimated for Faecal Char, Sawdust
Char, and Blend using KAS Method at Different α ...................................... 95
Table 4.12: Activation Energies (Eα (kJ/mol)) Estimated for Faecal Char, Sawdust Char
and Blend using FWO Method at Different α ............................................... 96
Table 4.13: Proximate Analysis Results and Heating Values of Briquettes and Charcoal
....................................................................................................................... 99
Table 4.14: Elemental Composition of Human Faecal Char, Sawdust Char, and
Molasses ...................................................................................................... 100
Table 4.15: Comparison of Flue Gases Quality During Combustion of Fuels ............. 104
xiv
LIST OF FIGURES
Figure 4.1: XRD Powder Patterns of (a) Lateritic Soil, and (b) Pineapple Peel Biochar
....................................................................................................................... 61
Figure 4.2: SEM Images of (a) Lateritic Soil, and (b) Pineapple Peel Biochar .............. 62
Figure 4.3: FTIR Spectra of Lateritic Soil , and Pineapple Peel Biochar ....................... 64
Figure 4.4: Thermogravimetric Analysis (TGA) Curves (solid lines) and DTG Curves
(dashed lines) of Lateritic Soil and Pineapple Peel Biochar at a Heating Rate
of 5 °C/min Under N2 Flow .......................................................................... 65
Figure 4.5: The Effect of Contact Time and the Initial Concentration on Adsorption of
Ammonium Nitrogen from Human Urine on (a) Lateritic Soil, and (b)
Pineapple Peel Biochar.................................................................................. 66
Figure 4.6: (a) Langmuir, (b) Freundlich, and (c) D-R Isotherm Model Fits of Lateritic
Soil and Pineapple Peel Biochar ................................................................... 69
Figure 4.7: The Effect of Contact Time and the Initial Concentration of Phosphorous
Adsorption on (a) Lateritic Soil and (b) Pineapple Peel Biochar.................. 73
Figure 4.8: (a) Langmuir, (b) Freundlich, and (c) D-R Isotherm Model Fits of Lateritic
Soil ................................................................................................................ 76
Figure 4.9: Physical Properties of Faecal Char, Sawdust Char, and Blend; (a) and (b)
SEM Images of Faecal Char and Sawdust Char, Respectively, (c) FTIR
xv
Spectra, and (d) TGA(solid lines) and DTG (dash lines) Profiles at a
Heating Rate of 5 °C/min from 30 to 800 °C Under Air .............................. 88
Figure 4.10: Thermal Behavior at a Heating Rate of 5 °C/min Under Flowing N2 and
Air from 30 to 800 °C ................................................................................... 91
Figure 4.11: TGA (solid lines) and DTG (dash lines) Profiles of (a) Faecal Char, (b)
Sawdust Char, and (c) Blend Under Air with the Ramp Rates of 5, 10, 20,
and 40 °C/min ............................................................................................... 92
Figure 4.12: Conversion Curve of the Entire Reaction (a) Faecal Char, (b) Sawdust
Char, and (c) Blend at Different Heating Rates (β)....................................... 93
Figure 4.13: KAS (a-c) and FWO (d-f) Plots of Faecal Char, Sawdust Char, and Blend
at Different Values of Conversion (α). .......................................................... 94
Figure 4.14: (a) Emissions from Co-combustion of Charcoal with Faecal Char-Sawdust
Char Briquettes, and (b) Emissions from Combustion of Charcoal ............ 101
Figure 4.15: Variation of Flue Gases Temperatures With Combustion Time .............. 105
Figure 4.16: Notched Box Plots of (a) Flue Gases Temperatures During Co-combustion
of Briquettes with Charcoal and Combustion of Charcoal, (b) Oxygen
Concentrations During Co-combustion of Briquettes with Charcoal and
Combustion of Charcoal.............................................................................. 107
Figure 4.17: Variation of Flue Gases Temperatures with Oxygen Concentrations ...... 108
xvi
LIST OF APPENDICES
Appendix I: Composition and Structures of Lateritic Soil (LS) and Pineapple Peel
Biochar (PPB) ......................................................................................... 153
xvii
ABBREVIATIONS AND ACRONYMS
CO Carbon monoxide
DI Deionized water
D-R Dubinin-Radushkevich
xviii
FC Fixed Carbon
FWO Flynn-Wall-Ozawa
KAS Kissinger-Akahira-Sunose
LS Lateritic Soil
P Phosphorous
xix
qm Maximum monolayer adsorption capacity (mg/g)
SD Standard Deviation
VM Volatile Matter
xx
ABSTRACT
Human excreta are abundant waste streams whose safe disposal to the environment is a
challenge in many developing countries due to inadequate sewerage system and poor
faecal sludge management. The growing practice of resource recovery from human
excreta could minimize the challenge of unsafe disposal thereby contributing to
protection of the environment and human health. The overall aim of the study was
therefore to recover nutrients from human urine and energy from human faeces. On
nutrient recovery, the study focused on adsorption of ammonium nitrogen (NH4+-N) and
phosphorous (P) in human urine using pineapple peel biochar (PPB) and lateritic soil
(LS). Physicochemical properties of PPB, and LS were characterized by Scanning
Electron Microscopy-Energy Dispersive Spectroscopy (SEM-EDS), X-ray powder
Diffraction (XRD), Fourier Transform Infrared Spectroscopy (FTIR) and
Thermogravimetric Analysis (TGA) to investigate the relationship of their properties
with adsorption of NH4+-N and P. Langmuir, Freundlich, and Dubinin-Radushkevich
(D-R) isotherm models were fitted to correlate the experimental equilibrium adsorption
data with the coefficient of correlation (R2) used to determine the model that offered the
best fit. The effect of contact time and initial concentration of NH4+-N and P on
adsorption was evaluated. Factors influencing adsorption process such as the contact
time, agitation speed, and adsorbent loading were optimised using Response Surface
Methodology (RSM) in order to predict the optimum conditions for achieving the
highest adsorption of NH4+-N and P. For adsorption of NH4+-N, the D-R isotherm
model best described the behaviour of its adsorption on both PPB and LS based on the
coefficient of correlation values. This model showed that the adsorption of NH4+-N on
both samples was a physical process with PPB and LS having mean surface adsorption
energies of 1.826×10-2, and 1.622×10-2 kJ/mol, respectively. The PPB exhibited a
slightly higher adsorption capacity for NH4+-N (13.40 mg/g) than LS (10.73 mg/g) with
the difference attributed to its higher contact surface area and porosity. The adsorbed
NH4+-N on PPB at optimal conditions described by RSM of 60 contact time, 80 rpm
agitation speed, and adsorbent loading of 0.1g, yielded experimental adsorption capacity
of 36.42 mg/g, which agreed well with the predicted adsorption capacity of 36.80 mg/g.
These values are good indicators for assessing the effectiveness of the materials for
adsorption of NH4+-N from human urine. For adsorption of P, PPB exhibited a tendency
to adsorb and gradually release P into the human urine thereby not attaining equilibrium.
LS on the other hand adsorbed P upto equilibrium. The P adsorption data on LS fitted
the Langmuir model best (R2 = 0.984) compared to the Freundlich model (R2 = 0.966).
The mean surface adsorption energy of 1.313 × 10-2 kJ/mol obtained from the D-R
model (R2 = 0.858) indicated that the adsorption of P on LS occurred through weak
forces of interaction. The amount of adsorbed P on LS at equilibrium (qe) increased with
the initial concentration of urine as well as with the contact time. The Langmuir
maximum adsorption of P on LS was found to be 45.25 mg/g. The RSM demonstrated
that contact time 30 min, agitation speed 80 rpm, and adsorbent loading 0.1 g in 25 mL
of urine were optimum for highest adsorption of P (112.80 mg/g). The results show that
LS is a promising inexpensive adsorbent for effective removal and recovery of P from
xxi
human urine. On energy recovery, the study examined the kinetic analysis of the
thermal decomposition of faecal char, sawdust char and their blend using
thermogravimetric analysis (TGA) under air to assess their reaction rates and thermal
stability as potential sources of fuel. The kinetic parameters during the degradation were
tested by combining the Kissinger-Akahira-Sunose (KAS), and Flynn-Wall-Ozawa
(FWO) iso-conversional methods. The TGA and deconvoluted DTG curves indicated
moisture release followed by devolatilization of hemicellulose, cellulose, and lignin
between ~300 and 450 °C, before ash formation. Both KAS and FWO methods
described the kinetic processes realistically and yielded similar activation energy values.
The lowest activation energy required to initiate degradation as calculated by KAS were
found to be 80.4, 80.0, and 86.4 kJ/mol for faecal char, sawdust char, and blend,
respectively. On the other hand, the activation energy for the whole conversion range
was 103.7, 108.7, and 104.8 kJ/mol for faecal char, sawdust char, and blend,
respectively. These results suggest that the compositional differences between the
samples translated to variability in the weight loss rates, shapes of decomposition peaks,
and parallel, competitive, and complex reaction schemes resulting to irregular trend in
the activation energy. Furthermore, the study on energy recovery compared the heating
properties and toxic flue gases namely; carbon monoxide (CO), nitric oxide (NO), and
hydrogen sulphide (H2S), emitted during combustion of charcoal and co-combustion
(50:50 wt. %) of charcoal with briquettes densified from human faecal char, sawdust
char, and molasses. The physicochemical properties (fixed carbon, volatile matter,
moisture content, ash content, and gross calorific value) of the briquettes and charcoal
was determined and characterization of the flue gases (concentration, oxygen level,
combustion temperature) conducted using E8500P industrial integrated emissions
system combustion gas analyser. It was observed that combustion of charcoal did not
emit NO, however the concentration of the CO was above the critical short-term limits
of 35 ppm. The level of emission of the CO and H2S was above the short-term exposure
limits of 35 and 0.005 ppm, respectively during co-combustion whereas NO
concentration was below dangerous exposure levels of 100 ppm throughout the
combustion period. Co-combustion resulted into release of higher heat energy as
evidenced by the flue gas temperatures reaching upto 475 °C compared to 222 °C during
charcoal combustion. The gross calorific value for briquettes was 19.8 MJ/kg which was
comparable to those reported for fuel wood although lower than 25.7 MJ/kg reported for
charcoal. These results suggest that co-combustion of charcoal with briquettes densified
from faecal and sawdust char is a promising alternative approach to generate safe and
sufficient heat energy for indoor use, reduce deforestation, and mitigate unsafe faecal
waste disposal issues. In conclusion, this research demonstrates that nutrients can be
captured from human urine to produce enriched biomass that can be used as slow-release
fertilizers. The study findings also demonstrated that human faeces are a potential
biomass for production of safe solid fuels. Hence, nutrients and energy recovery from
human excreta could mitigate the environmental and health impacts associated with their
unsafe disposal into the environment.
xxii
CHAPTER ONE
INTRODUCTION
With on-site sanitation facilities being the most common systems used globally (Strande
et al., 2014; WHO & UNICEF, 2017), the entire faecal sludge management chain from
containment of human excreta, emptying, transport, treatment, safe disposal or reuse is
of paramount importance in order to protect the population from water borne diseases.
However, inadequate financial resources have hampered the emptying and or safe
disposal of the accumulated sludge in these facilities, often resulting to their
abandonment or disposal of the untreated sludge into the environment thereby
contaminating water resources (Bassan et al., 2015). The practice of recovery of
valuable resources from human excreta contained in the on-site sanitation facilities with
the aim of revenue generation is considered as a promising way of sustainably managing
the waste (Diener et al., 2014; Gold et al., 2014). This approach to management of
human excreta could potentially improve overall access to properly managed sanitation
facilities by promoting frequent desludging. Noteworthy is the development of the
ecological sanitation concept also commonly referred to as Ecosan. Ecological sanitation
is a sanitation approach that aims at collecting the different waste streams separately at
1
source and treating them individually with the objective of recovering valuable nutrients,
energy, and clarified water (Langergraber & Muellegger, 2005). Urine diversion and
dehydration toilet (UDDT) is a form of ecological sanitation system which operate on
the principle of separation of human faeces and urine at the point of generation thus
enabling utilization of these waste streams separately (Langergraber & Muellegger,
2005).
The potential of both human urine and faeces for agricultural production have been
widely reported (Harder et al., 2019; Gwara et al., 2021). There have also been few
studies at laboratory scale aimed at recovering energy from human urine (Kuntke et al.,
2014; Salar-García et al., 2017), while for human faeces, a variety of options have
recently been explored at both lab and field scale such as for production of solid fuel,
protein for animal feed, biogas and as a component in building materials (Diener et al.,
2014). For human urine, the World Health Organization (WHO) recommends storage for
6 months to achieve complete sanitization (WHO, 2006), a practice likely not to be ideal
especially in urban and peri-urban areas due to odour problems and large space
requirements. This has prompted a focus on nutrients extraction or capture by many
researchers. Human urine is rich in macro-nutrients such as Nitrogen (N), and
Phosphorous (P) (Udert et al., 2003; Maurer et al., 2006; Etter et al., 2011) which are
readily up taken by plants and thus can contribute to enhanced crop production if
captured and incorporated in the soil. However, ammonium (NH4+-N), and P are the
major cause of eutrophication in water bodies as a result of wastewater inflows, with
human urine being the major contributor of these nutrients in municipal wastewater
flows (Nguyen et al., 2019; Preisner et al., 2021).
Among the technologies already reported aimed at recovering NH4+-N from human
urine are microbial electrolysis cell (Kuntke et al., 2014), microbial fuel cells (Ledezma
et al., 2015), ammonia stripping (Behrendt et al., 2002), and adsorption using adsorbents
such as ion exchange resins (BelerBaykal et al., 2004), geological materials (Kithome et
al., 1998), biochar (Wang et al., 2015), and activated carbon (Pillai et al., 2014).
Similarly, researchers have explored various technologies for recovery of P from
2
wastewater or human urine, with struvite (MgNH4PO4.6H2O) precipitation upon
addition of magnesium being the most widely investigated method (Ban and Dave,
2004; Ganrot et al., 2007; Wang et al., 2018; Li et al., 2019). Other techniques reported
are solar thermal evaporation (Antonini et al., 2012), freeze thaw (Ganrot et al., 2007),
electrochemical (Perera & Englehardt, 2020) and adsorption (Mansing R & Rout, 2013;
Pitawala et al., 2013; Ling Zhang et al., 2014). Of the reported NH4+-N and P recovery
techniques, adsorption is considered the most ideal method for nutrients recovery
especially in developing nations due to its flexibility of design, ease of operation and
low cost (Martin et al., 2020). The use of biochar and geological materials for adsorption
of nutrients from aqueous solutions has been demonstrated to be effective. Thus,
investigating the adsorption of NH4+-N, and P in human urine on biochars derived from
organic wastes, and geological materials that have not been reported on is attractive.
Utilization of biochar derived from pineapple peels, and lateritic soil for nutrients
recovery from human urine is rather scarce and hence investigation of their
physicochemical properties, mechanisms of nutrient removal and their adsorption
capacities would be key in optimising their use to produce slow-release fertilizers or an
enriched medium for crop growth. Biochar is a carbon material formed as a result of
combustion of biomass material under minimal or no oxygen conditions whereas
Lateritic soil (LS) is formed as a result of rapid weathering of rocks and high rate of
leaching in the tropical climate and is normally nutrient deficient, thus majorly used for
construction purposes (Ehujuo et al., 2017; Stoops & Marcelino, 2018; Ng et al., 2019).
Production of solid fuels (briquettes) from human faeces for domestic heating is an
approach that is currently practiced in Kenya (Karahalios et al., 2018). It is however
recommended to pyrolyze human faeces at temperatures > 300 oC to eliminate all
pathogens to make it safe for briquetting (Atwijukye et al., 2018). During pyrolysis,
biomass materials are heated under inert conditions to yield a charcoal-like material
called biochar (Mishra & Mohanty, 2018). Pyrolysis however reduces the calorific
values of human faeces due to degradation of aliphatic hydrocarbons (Ward et al.,
2014). On the other hand, the calorific values of wood biomass such as sawdust are
3
known to increase with pyrolysis temperature due to an increase in their fixed carbon
content (Bulmau et al., 2010). Thus, it is recommended that pyrolyzed human faeces
and other carbonized woody waste streams could be blended to improve the resultant
calorific values (Atwijukye et al., 2018). Normally charred materials lack plasticity and
therefore it is common practice to moisten them with binders such as starch and
molasses before densification so as to produce durable briquettes. Blending pyrolyzed
human faeces (hereon referred to as faecal char) with sawdust char and using molasses
as a binder in order to produce briquettes has been the most widely reported approach
(Kabok et al., 2018; Karahalios et al., 2018). Sawdust is a widely abundant waste from
saw milling activities in Kenya while molasses is widely abundant as a by-product of
sugar processing and hence they are sustainable source of raw materials in the
production of briquettes.
Despite the ongoing practice of briquettes production from faecal char and sawdust char,
studies involving thermal degradation characteristics, and reaction kinetics of the
pyrolyzed pathogen free human faeces, sawdust char and their blend using
thermogravimetric analysis (TGA) under ambient conditions are lacking. Studying the
thermal degradation characteristics and reaction kinetics is vital in giving insights on the
practical thermal behaviour of briquettes densified from the charred faeces and sawdust
char. Furthermore, studies on the safety of the briquettes containing faecal char for
indoor heating in terms of their level of emission of toxic gases such as CO, H2S and NO
are scarce.
The WHO/UNICEF Joint Monitoring Programme (JMP) report for Water Supply,
Sanitation and Hygiene published in 2021, revealed that 46% of the global population (
3.6 billion people) lack safely managed sanitation service in which excreta is safely
disposed of in situ or treated off-site (WHO & UNICEF, 2021). This implies that nearly
half of the global human population generates excreta whose safe disposal cannot be
accounted for, a situation that pose significant health risks to humans as a result of
4
environmental pollution. Other studies have reported that inadequate management of
human excreta is among the major contributors of diseases causing death of children
under the age of five (Cheng et al., 2012; Bassan et al., 2015). Coupled with increase in
waterborne diseases due to inadequate management of human excreta, is the rising level
of eutrophication of aquatic systems due to domestic wastewater inflows, a situation
which has resulted into hypoxic conditions that affect aquatic life (Nguyen et al., 2019).
Human urine is the major contributor of macronutrients in wastewater conveyed to
centralized treatment plants with approximately N (80–90%), P (50–65%), and K (50–
80%) (Lienert et al., 2007; Jana et al., 2012). Recovery of valuable resources from
human excreta contained in on-site facilities is currently considered as a sustainable way
of managing both faecal matter and urine in unsewered areas. Production of briquettes
from human faeces has been established to be the most profitable venture compared to
its tradition use as soil amender in sub-Saharan Africa (Muspratt et al., 2014). On the
other hand, application of human urine as fertilizer has been reported to yield higher
produce compared to chemical fertilizers (Viskari et al., 2018), with WHO
recommending its storage for 6 months to achieve complete sanitation for use in
agriculture (WHO, 2006). Nevertheless, studies that have reported on briquettes
production from human faecal char for domestic heating and cooking have majorly
focused on the influence of organic binder types (molasses and starch), binder ratios,
densification pressures, and carbonization temperatures on the combustion behaviour
and heating values of the solid fuels (Ward et al., 2014; Kabok et al., 2018). However,
investigations on the thermal decomposition behaviours and reaction kinetics of human
faeces, using thermogravimetric analysis (TGA) under air conditions to depict their
practical behaviours have not been reported. Furthermore, despite the increasing
attention on briquettes production from human faeces, studies on emission of toxic gases
liberated during combustion of these briquettes are lacking. For human urine, there has
been increasing attention on nutrients adsorption from source separated human urine on
biochars and geological materials to produce an enriched biomass for use as slow-
release fertilizers. Nutrients adsorption from human urine is not only considered low
cost but also environmentally friendly since it eliminates the need for prolonged urine
5
storage for 6 months that often generates offensive odours (Alhashimi & Aktas, 2017).
However, studies aimed at investigating the physicochemical properties, adsorption
mechanisms and nutrients adsorption capacities of pineapple peel biochars and lateritic
soil in human urine are lacking despite their wide availability. Moreover, adsorption
process being influenced by a number of factors, the one at a time factor approach used
in obtaining optimum conditions cannot capture the interactive effects of the various
factors. Hence employing interactive tool such as Response Surface Methodology
(RSM) that captures all interactions (Myers et al., 2004) could be useful in predicting
optimum conditions to attain the highest level of adsorption.
1.3 Objectives
The main objective of this study was to evaluate and optimize nutrients and energy
recovery from ecological sanitation system.
i. Determine the behaviour and capacity of pineapple peel biochar and lateritic soil
in adsorbing ammonium nitrogen and phosphorous from human urine.
ii. Optimise the adsorption of ammonium nitrogen and phosphorous from human
urine on pineapple peel biochar and lateritic soil.
iii. Determine the thermal degradation and kinetic behaviours of pyrolyzed human
faeces, sawdust char and their blend.
iv. Evaluate emission properties of briquettes densified from pyrolyzed human
faeces, sawdust char and molasses.
6
1.4 Research Questions
i. What are the behaviours and capacities of pineapple peel biochar and lateritic
soil in adsorbing ammonium nitrogen and phosphorous from human urine?
ii. What are the optimum conditions for adsorption of ammonium nitrogen and
phosphorous from human urine on pineapple peel biochar and lateritic soil?
iii. Are the thermal degradation and kinetic behaviours of pyrolyzed human faeces,
sawdust char and their blend different?
iv. What are the emission properties of briquettes densified from pyrolyzed human
faeces, sawdust char and molasses?
7
This study therefore investigated the thermal stability and reaction rates of faecal char in
comparison to sawdust char and their blends under air atmospheres, as well as the
heating and emission properties of briquettes densified from faecal char, sawdust char,
and molasses. On nutrients recovery, pineapple peel biochar and lateritic soils were used
to adsorb ammonium nitrogen (NH4+-N) and phosphorous (P) from the human urine to
produce slow-release fertilizer.
The overall aim of the study was to recover nutrients and energy from ecological
sanitation systems. Pineapple peel biochar (PPB) and lateritic soil (LS) were used as bio-
sorbents for recovery of ammonium nitrogen (NH4+-N) and phosphorous (P) from the
human urine. Langmuir, Freundlich, and Dubinin-Radushkevich (D-R) isotherm models
were employed to correlate the experimental equilibrium adsorption data, while
Response Surface Methodology (RSM) was used to optimise the adsorption process. It
was assumed that both PPB and LS could effectively adsorb the cationic NH4+-N, and
anionic PO43- species upto equilibrium to enable adsorption isotherm modelling to be
conducted so as to calculate the NH4+-N and P adsorption capacities of both adsorbents
from human urine. Although LS could adsorb NH4+-N and P upto equilibrium, PPB
8
adsorbed and gradually released P into the human urine solution thereby not attaining
equilibrium conditions. The PPB surface could have been modified with metal cations so
as to attract and retain the anionic PO43- species on its surface. On energy recovery,
kinetic analysis of the thermal decomposition of pyrolyzed human faeces, sawdust char
and their blend using thermogravimetric analysis under air was conducted to assess their
thermal stability and reaction rates as potential as sources of fuel. Furthermore,
briquettes densified from faecal char, sawdust char and molasses were produced and co-
combusted with charcoal to characterize their emission properties to determine their
safety for indoor household heating and cooking. It was assumed that briquettes
densified from faecal char, sawdust char, with molasses (10 wt.%) used as a binder
could easily ignite without the need for co-combustion with other fuels. It was however
observed that upon heating, molasses melts causing disintegration of the densified
briquettes thereby making ignition difficult. Although co-combustion of the briquettes
with charcoal enabled the briquettes to ignite easily, lower amounts of charcoal (< 50
wt.%) could have been used to monitor their effect on the heating and emission
properties during co-combustion.
9
CHAPTER TWO
LITERATURE REVIEW
Human urine is a complex water solution excreted from the body through the urethra.
The biological, physical, and chemical properties of human urine have been extensively
studied (Kirchmann & Pettersson, 1994 Etter et al., 2011; Bischel et al., 2015). The use
of urine for agricultural production has attracted attention of researchers due to its high
concentration of major macronutrients (NPK) required for plant growth (Simha et al.,
2017; Viskari et al., 2018; Pradhan et al., 2019). Moreover, according to WHO,
complete pathogen elimination in human urine can be attained by storing it for a period
of six months making it suitable for direct application in the field for agricultural
production after dilution (WHO, 2006).
Approximately 91–96% of human urine is water (Höglund et al., 2000) with inorganic
salts, urea, organic compounds, and organic ammonium salts constituting the remaining
fraction (Putnam, 1971). Therefore, the amount of urine generated is dependent on the
water intake of individuals. Normally, one person produces 0.6 to 2.6 L of urine daily
with an average value estimated at 1.4 L daily (Rose et al., 2015). Other factors
influencing urine generation are body size and age with adults producing twice as much
urine as children (Karak & Bhattacharyya, 2011). The high generation rates of urine by
adults is attributed to greater physical activity, obesity and higher medication intake
than children (Clark et al., 2011). It can be recommended that urine from children is
safer for use in agriculture since they contain less pharmaceuticals.
10
2.1.2 Composition
Since urine is an excretion from the human body, its composition is influenced largely
by the diet. The most predominant constituent in urine (>50%) is urea [CO(NH2)2]
which is the major source of total nitrogen in fresh urine (Udert et al., 2003; Etter et al.,
2011). Urea is produced through metabolism of protein. The other major solutes
(macronutrients) excreted in urine are Na (2.5-3.4 g/L), P (0.8–2.0 g/L), K (1.8-2.7 g/L)
and Cl (3.7-4.9 g/L) which are largely derived from dietary intake (Maurer et al., 2006).
Ammonium (NH4+) is produced through hydrolysis of urea by urease enzyme with the
process also yielding carbamate that spontaneously decomposes to carbonic acid and
another molecule of ammonia through a process called ureolysis as shown in Equation
2.1 (Udert et al., 2003).
According to Hotta & Funamizu (2008), without urease, urea is a very stable compound
with a half-life time of 3.6 years at 38 ⁰C. Although fresh urine is normally acidic with
pH ranging 5.6 – 7.2 (Udert et al., 2003; Etter et al., 2011), the release of ammonia
during ureolysis increases pH (Udert et al., 2003). Variations in urine pH amongst
individuals has also been reported to be influenced by diet and alcohol intake (Kanbara
et al., 2012). Noteworthy is that urine also contains enzymes, uric acid, vitamins,
chlorides, oxalates, amino acids (Lind et al., 2000; Winker et al., 2009), heavy metals
(Kirchmann & Pettersson, 1994), pharmaceuticals and hormones (Lienert et al., 2007),
viruses and bacteria (Björn Vinnerås et al., 2008). Therefore, pre-treatment of fresh
urine before use for agricultural production is recommended.
The concept of source separation involves the separate collection of urine from faeces in
the toilet consequently allowing each of these waste streams, which have significantly
different properties from one another, to be treated in a specific and appropriate manner.
11
This concept of source separation of human excreta also referred to as ecological
sanitation (Eco-San), was originally developed to achieve the Millennium Development
Goals in the sanitation sector with the target of preventing pollution, and sanitizing
excreta for use in agriculture (Esrey et al., 1998; Langergraber & Muellegger, 2005).
The Eco-San concept includes a number of toilet designs and systems that are
appropriate for various contexts. The most common and widely researched Eco-San
technology is the Urine Diversion and Dehydration Toilets (UDDTs), in which urine is
diverted so that it does not come into contact with faeces (Langergraber & Muellegger,
2005). The whole structure of UDDTs is built above the ground and typically consists of
a bowl divided into two parts. That is, the front bowl collects the urine and the rear
bowls the faeces (Larsen et al., 2009). After separation, both faeces and urine are
collected in containers located beneath the superstructure above ground and treated on-
site or transported further to a centralized treatment unit (Larsen et al., 2009). In many
cases, it is advisable that each user sprinkles a desiccant in form of ash, soil, or dry
leaves to the excrement to minimize foul smells and keep away flies. In Kenya, there has
been a growing trend of sanitation entrepreneurs installing UDDTs to meet sanitation
needs in the slum areas and in market centre’s of major towns with the aim of recovering
resources from the waste streams (www.sanergy.com).
2.2.1 Urine Diversion and Dehydration Toilets and Associated Problems with Urine
Fertilizer Recovery
Diversion of urine for storage and reuse that UDDTs promotes exhibit inherent
challenges. Loss of nitrogen through ammonia volatilization during storage as a result
hydrolyses of urea and subsequent elevation of pH reduces potential reusability of
nitrogen in post storage application (Udert et al., 2003). A further concern of UDDT is
cross faecal contamination of the relatively sterile source separated urine (WHO, 2006).
Inactivation studies with urine point towards significant pathogenic risks over its use due
to persistence of faecal sterols, Escherichia Ecoli, Salmonella Typhimurium, Ascaris
Suum eggs and rhesus rotavirus (Nyberg et al., 2014; Winker et al., 2009). WHO
therefore recommends that urine should be stored for a period of 6 months at a
12
temperature of 20 ⁰C to achieve complete sterilization (WHO, 2006). However this
promotes ureolysis (Udert et al., 2003). Studies on optimization of pathogen die-off in
recovered fertilizers from urine however indicates that the pathogens can be eliminated
within a shorter duration by increasing drying temperature to less than 55 0C and pH of
urine to a level ≥ 10.5 (Randall et al., 2016a; Winker et al., 2009). Alkalinization
therefore is an effective approach to stabilize and preserve nutrients in urine.
2.3 Technologies for Nitrogen and Phosphorous Recovery from Human Urine
In order to produce enough food for the growing world population, farmers will need to
rely increasingly on fertilizer to replenish soil. Nitrogen and phosphorous are the key
components in most fertilizers in use. Currently, the Haber-Bosch process is used to
produce nitrogen for fertilizer but consumes more than 1% of world’s total energy
production (Kitano et al., 2012). Phosphorus is a non-renewable resource that is quickly
being mined to depletion (Neset & Cordell, 2012). Fortunately, a large portion of the
nitrogen and phosphorus used in fertilizers eventually become present in human urine
(Bonvin et al., 2015; Kirchmann & Pettersson, 1994). Recycling these components from
urine to fertilizer would make agriculture more sustainable in the future. Recent research
effort has been devoted towards the development of technologies that can safely harness
nutrients from human excreta to yield usable end products.
Although N exist in different forms in human urine, N in the form of free ammonia
(NH3-N) and ammonium (NH4+-N) constitute the largest proportion (Udert et al., 2003;
Etter et al., 2011). However, NH3-N in urine is easily lost through volatilization, while
NH4+-N increases gradually with time due to ureolysis caused by the urease enzyme
(Udert et al., 2003). Technologies aimed at recovery of NH4+-N from human urine
without the necessity of prolonged storage have drawn the attention of many researchers.
Among them are microbial electrolysis cell (Kuntke et al., 2014), microbial fuel cells
(Ledezma et al., 2015), ammonia stripping (Behrendt et al., 2002), and adsorption using
adsorbents such as ion exchange resins (BelerBaykal et al., 2004), geological materials
(Kithome et al., 1998), biochar (Wang et al., 2015), and activated carbon (Pillai et al.,
13
2014). Despite the possibility of recovering NH4+-N from the aforementioned
technologies, operational costs and the sensitivity of a given approach to slight changes
in operational parameters could determine the choice of one approach over the other. For
instance, NH4+-N recovery using biological processes (microbial electrolysis cell,
microbial fuel cell) are likely to be sensitive to changes such as pH and temperature
since they are microorganism dependent. Ammonia stripping on the other hand requires
energy input, and may be expensive to run for large scale production. According to
Kavvada et al. (2017), large scale recovery of NH4+-N through ion exchange process is
quite expensive but not detrimental to the environment. Currently, adsorption is
considered the most ideal method of NH4+-N recovery since it is less costly, flexible,
efficient and environmentally friendly, and thus can be ideal for the developing countries
(Alhashimi & Aktas, 2017).
Phosphorous is mainly present in human urine in the form of phosphates (PO43-) (Maurer
et al., 2006). Researchers have explored various technologies for P recovery from
wastewater or human urine, with struvite (MgNH4PO4.6H2O) precipitation upon
addition of magnesium being the most widely investigated method (Ban and Dave,
2004; Ganrot et al., 2007; Wang et al., 2018; Li et al., 2019). Other techniques reported
are solar thermal evaporation (Antonini et al., 2012), freeze thaw (Ganrot et al., 2007),
electrochemical (Perera & Englehardt, 2020), and adsorption (Mansing & Rout, 2013;
Pitawala et al., 2013; Zhang et al., 2014). Of the reported P recovery techniques,
adsorption is considered the most ideal especially in developing nations due to its
flexibility of design, ease of operation and low cost (Martin et al., 2020). However,
energy requirements of these processes are a major determent.
14
in the removal of solutes from solution and their accumulation at solid surface. The
adsorption of solutes continues until equilibrium is attained. This process is however
influenced by a number of factors. Öztürk & Bektaş (2004) demonstrated that increase
in contact time, pH and adsorbent dosage increases nutrient removal from aqueous
solutions. Mouni et al. (2018) also showed that adsorption is spontaneous and enhanced
at higher temperatures because as temperature rises, solution viscosity drops, which is
favorable to the subsequent adsorption stages. Increase in the surface area of the
adsorbent which is achieved by decreasing the particle sizes of the adsorbent has also
been found to increase adsorption level (Khamparia & Jaspal, 2016; Matsui et al., 2015).
Agitation speed has also been reported to increase collision resulting to higher
adsorption (Shen and Duvnjak, 2005).
Activated carbons which are processed carbonaceous materials with high porosity and
internal surface area available for adsorption of liquids and gases or chemical reactions
(Saleem et al., 2019), are the most popular and widely used adsorbents in recovery of
nutrients from human urine throughout the world (Ganrot et al., 2007). In spite of large
use, the cost of their production may be a hindrance to their sustainability for field scale
application and hence seeking alternative but efficient adsorbents is ideal. It is worth
noting that although both activated carbon and biochars are produced via heating
biomass under minimal oxygen or anaerobic conditions the significant difference lies in
their production temperature with biochar being produced at lower temperature less than
700 ℃ while activated carbon is produced at higher temperatures more than 700 ℃ (e.g.
800-1000 ℃) resulting to higher porosity (Saleem et al., 2019).
Low-cost sorbents sources can be classified as; agricultural and household by-products,
industrial by-products, sludge, sea materials, soil and ore materials and novel low-cost
adsorbents (Ali et al., 2012). The use of wastes from agricultural and household sources
as well as geological materials (soil, ore, etc.) as adsorbents is an attractive option due to
their ease of availability and low processing cost and hence sustainability. Researchers
have thus reported the use of geological materials such as zeolite (Kithome et al., 1998;
Wu et al., 2006; Huang et al., 2010), and biochar derived from pyrolysis of biomass
15
derived from different materials (Hale et al., 2013; Zhang et al., 2014) for extracting
nutrients from human urine.
1 1 1
(2.2)
qe qm K L qm Ce
Where:
KL = the binding energy between the adsorbate and the adsorbent (L/mg)
16
The plot of 1/qe against 1/Ce gives a linear plot from which the parameters qm (mg/g),
and KL (L/mg) are obtained.
The Langmuir model constant (KL), can further be used to compute the dimensionless
separation factor (RL) according to Equation 2.3 given as:
1
RL (2.3)
1 K L Co
Where:
If RL is in the range between 0 and 1, then the adsorption process is considered favorable
(Hameed and ElKhaiary, 2008).
1
log q e log K f log C e (2.4)
n
Where:
17
Kf = the constant incorporating the factors affecting the adsorption capacity
1/n = the constant incorporating the factors affecting the adsorption intensity
The plot of log qe versus log Ce gives a linear plot, from which the empirical parameters
Kf and 1/n are obtained from the intercept and the gradient, respectively.
Where:
The plot of lnqe versus [RTln(1+1/Ce)] 2 gives a linear plot from which the parameters
KDR (mol2/kJ2), and Qo (mg/g) are calculated from the gradient and the intercept,
respectively.
The mean surface adsorption energy (E) (kJ/ mol) is calculated using Equation 2.6 (Hu
& Zhang, 2019) given as:
18
1
E (2.6)
2 K DR
Adsorption process is considered physical if E < 8 kJ/mol, whereas chemisorption
occurs if E > 18 kJ/mol (Mahmoud, 2015).
In order to apply RSM as an optimization tool, some stages need to be followed (Dutta
et al., 2018). They include:
19
2.5.1 Experimental Scheme and Design
Besides analyzing the independent variables effects, this experimental methodology also
generates a mathematical model. The graphical viewpoint of the mathematical model
has led to the term RSM. The relationship between the responses and the inputs are
represented by Equation 2.7.
Y f ( x1 , x2 , x3 ...............xk ) (2.7)
Where:
Y = the response
20
x1, x2, x3 … xk = the input variables which affects the response
After selection of the design, the model equation and coefficients are defined and
predicted. As a result of sequential model sum of squares, a quadratic model is selected
by Design Expert software. Usually a second-order model is utilized in response surface
methodology as shown in Equation 2.8.
k k k
y o i xi ii xi2 ij xi x j (2.8)
i 1 i 1 i 1
Where:
βo = the value of the fixed response at the center point of the design
βi, βii, and βij = the linear, quadratic and interaction effect regression terms, respectively
The model correlates the dependent variables to the independent variables considered
through second- order polynomial response equation according to Equation 2.9 given as:
The selected independent variables are coded according to Equation 2.10 given as:
21
X i X o
xi (2.10)
X
Where:
Design expert software is used for statistical and diagnostic studies in response surface
modeling. RSM offers adequacy measures such as the Analysis of Variance (ANOVA),
sequential F-test and lack-of-fit test. The sum of square for the deviations in all the trials
is known as the total sum of square (SST). It is computed by adding the sum of square
caused by the residual (SSE) and the sum of square caused by the regression (SSR) as
shown in Equation 2.11.
SST SS E SS R (2.11)
During repetition of experiments for accuracy and confidence in the results, errors and
deviations result. Hence, the SSR is found by adding the sum of square due to the lack
of- fit (SSL) and the sum of square due to the pure error (SSP) as shown in Equation
2.12.
SS R SS L SS P (2.12)
The Fischer test (F-test) of a model evaluates the significance of the model by
calculating the ratio of the mean square of regression (MSR) to the mean square of
residuals (MSE). A small F-value for the model is not desired since it indicates that the
variance is caused by random unexplained disturbances referred to as noise. The p-value
22
(p > F) provides an indication of the significance of a model in relation with the F-value.
It can be defined as the probability that a variable did not affect the response for a given
F-value. If the p > F for the model is less than 0.05, a model is said to be significant,
meaning that there is 5% chance that the F-value is due to noise. If the p > F is above
0.1, the model is insignificant (Trinh & Kang, 2011).
The lack-of-fit test (LOF) determines the inability of a model to fit experimental data
that are not represented in the experimental domain. This is commonly done by
calculating its F-value. A small F-value for the LOF is desired, since the experimenter
wants the model to fit. If the p > F is greater than 0.05 the LOF for the model is
insignificant and the model is able to fit any data that are not specified in the
experimental domain (Trinh & Kang, 2011). A good LOF does not guarantee the
adequacy of a model, the coefficient of determination (R²) as shown in Equation 2.13
must be considered, given the fact that it measures the overall performance of a model
and its value should be close to 1 (Trinh & Kang, 2011).
SS
R 2 1 R (2.13)
SST
SS / df E
Adjusted R 2 E (2.14)
SST / dfT
The PRESS (Predicted residual sum of squares) is used to measure the predicted R 2,
which indicates the level of change in the data using the model (Myers et al., 2004).
During the selection of a model, the aim is to maximize the adjusted and predicted R 2
23
values. For a best fit, the difference between the adjusted and predicted R2 must be less
than 0.2. PRESS is calculated using Equation 2.15 given as:
2
n
e
PRESS i (2.15)
I 1 1 hii
Where:
ei = the residual
hii = the leverage which is the difference between the adjusted and predicted R2
Other parameters in the analysis of a model include the standard deviation (SD)
expressed in Equation 2.16 and Adequate Precision Statistics (AP) as in Equation 2.18
given as:
SD MS E (2.16)
The coefficient of variation (CV) measures the unexplained changes in the data and it is
expressed as shown in Equation 2.17 given as:
SD
CV . 100 (2.17)
Y
Where:
The adequate precision statistics (AP) determines the performance of the model in
predicting the responses. A value of AP greater than four means that the model will give
good predictions (Myers et al., 2004).
24
~ ~
AP Y max Y min (2.18)
Where:
Diagnostic plots such as predicted versus actual plot, normal plot of residual, residuals
versus predicted and residual versus the run predicts the reliability of the developed
model and thus confirms that the ANOVA assumptions are correctly met. The residual is
calculated as shown in Equation 2.19;
~
Re sidual ei Y Yi (2.19)
Where:
ei
IS Re s r (2.20)
S 1 hii
25
Where:
ESRes (t) as shown in Eq. 2.21 makes a comparison between the residual and the
residual variance, excluding the first case of SD (s-1). It represents the number of SD
between the predicted value and the actual response (Myers et al., 2004).
ei
ES Re s t (2.21)
S 1 1 hii
The optimization can be done graphically or numerically using Design Expert Software after
the goals or criteria of the conditions desired for a given process is set by the experimenter.
The RSM optimizes its factors through an objective function called the desirability function
(D) (Equation 2.22) and transforms an estimated response to a scale free vale (di) called
desirability. The desirability is a dimensionless entity that represents the closeness of a
response to its ideal value and usually ranges from 0 to 1 with d = 0 being unacceptable,
and d = 1 indicating that the model response is equal to that of the target value (Myers et
al., 2004; Li et al., 2007).
1/ n
n
D (d1 d 2 d 3 ...... d n )1 / n di 2.22
i 1
Where:
di = desirability
26
2.6 Pineapple Peel Wastes and Application of Pineapple Peel Biochar as an
Adsorbent
In Africa, Nigeria, Kenya, Democratic Republic of Congo, Ivory Coast, Guinea and
South Africa are the major pineapple producing countries (Hossain, 2016). According to
Hossain (2016), in Kenya, large scale producers of pineapples namely; Delmonte (K)
Limited based in Thika, Kakuzi limited based in Murang’a and Ndemo farm based in
Kilgoris contribute close to 90% of all pineapples grown in Kenya. Other medium and
small-scale producers account for about 10% of the total pineapple production. The
increasing demand for pineapples with the ever growing human population consequently
results into an increase in the amount of waste generated. Approximately 80% of
pineapple parts (crown, peels, leaves, core and stems) are discarded during pineapple
processing, transportations and storage and ends as waste (Hikal et al., 2021). The
conventional techniques commonly applied in managing peel residues has been disposal
to landfill, composting, and incineration. However, these traditional approaches are
increasingly raising environmental concerns such as emission of methane gas from
landfilling and discharge of toxic compounds like dioxin from incineration (Sial et al.,
2019). Composting has also been reported to result in release of foul odour thereby
affecting air quality (Bu et al., 2014). Conversion of pineapple peels into biochar for use
as adsorbents is an approach that is currently being explored(Lam et al., 2018; Sial et al.,
2019). Among the adsorption studies commonly reported for biochar is on organic
contaminants (Fu et al., 2016; Zhou et al., 2015), heavy metals (Wang et al., 2016), and
dyes (Foo & Hameed, 2012). Despite this, the application of pineapple peel biochar
(PPB) for nutrient recovery from aqueous solutions is rather scarce with the very recent
report (Hu et al., 2020) on its use for NH4+-N recovery from wastewater exhibiting an
adsorption capacity of 5 mg/g. This value is significantly lower than those reported for
other biochars. This variation could be attributed to factors such as the initial
concentrations of adsorbate, and physicochemical characteristics of the adsorbents.
Consequently, human urine whose pH is unaltered with addition of alkaline solutions
27
such as NaOH becomes an interesting solution to determine the adsorption capacity of
the PPB.
Lateritic soil (LS) are commonly found in the tropical climatic regions and are formed as
a result of chemical weathering of pre-existing soils such as granite, sandstone, shale and
quartzite (Ehujuo et al., 2017). Due to humid conditions in the tropics, these soils are
excessively leached making them nutrient deficient and unsuitable for crop production
and thus majorly used in construction industry (Ehujuo et al., 2017; Ng et al., 2019;
Stoops & Marcelino, 2018). These soils are highly rich in metal oxides of Fe and Al
with the Fe2O3 content responsible for the characteristic reddish brown coloration
(Stoops & Marcelino, 2018). Most of the studies on the use of LS for adsorption have
mainly focused on the removal of fluorides (Osei et al., 2016; Sarkar et al., 2006), and
heavy metals (Glocheux et al., 2013; Mitra et al., 2016) from water or wastewaters.
Exploring its capacity for adsorption of nutrients could provide an insight into its
suitability for producing slow release fertilizer and its surface properties. An
investigation on its application for the removal of PO43- from wastewater is very recent
(Bhattacharjee et al., 2021). The concentrations of PO43- in human urine are orders of
magnitude greater than in the domestic wastewater (Maurer et al., 2006a), and so
determining the LS adsorption capacity in natural human urine would be important to
optimize its use for the removal of phosphorous. Its application for NH4+-N adsorption
have however not been reported.
Human faeces is composed of undigested materials that pass through the intestines,
usually mixed with extracts from the blood stream, mucus and bile from the glands and
the intestines impacting the characteristic brown colour (Rose et al., 2015).
28
2.8.1 Generation Rate
The amount of faeces produced by a person depends on the composition of the food
consumed with higher fiber intake resulting to significant increase in faecal weight
(Chen et al., 1998). The faecal production in the developed countries is approximately
80-140 g/p/d (wet weight) of faeces (Vinnerås et al., 2006). Faecal excretion rate in the
developing countries is on average 350 g/p/d in rural areas and 250 g/p/d in urban areas,
giving a range of 250-350g/p/day (Feachem et al., 1984).
2.8.2 Composition
29
depicting that they have calorific values ranging from 19 – 25 MJ/kg (Onabanjo et al.,
2016; Afolabi et al., 2017; Somorin et al., 2020) which is comparable to that of wood
fuel (Ruiz-Aquino et al., 2019).
In order to render human faeces safe for handling, pyrolyzing the waste at temperatures
> 300 oC can lead to complete elimination of pathogens (Atwijukye et al., 2018).
Pyrolysis, which is the decomposition of a feedstock through heating in the absence of
oxygen, is highly affected by a variety of parameters such as heating rate,
temperature, gas flow rate, biomass composition, residence time, moisture content, and
particle size (Mishra & Mohanty, 2018). During pyrolysis, dried human faeces are
heated under inert conditions to yield a charcoal-like material called biochar. For solid
fuel production, the char is then densified by moistening with binders such as starch, and
molasses. Other advantages of pyrolysis is that it minimizes oxygen containing
functional groups, while also increasing the aromatic carbon content, which in turn
lessens the emission of CO2 and smoke (C. He et al., 2013). However, this process also
lowers the calorific values of human faeces due to the degradation of energy rich
aliphatic hydrocarbons with increasing pyrolysis temperature (Ward et al., 2014).
Despite this feature, the gross calorific values of human faeces pyrolyzed between 300
C and 450 oC are within 18 – 26 MJ/kg (Ward et al., 2014), a range comparable to 19
o
– 25 MJ/kg of wood fuel (Ruiz-Aquino et al., 2019). The calorific values of wood
biomass are known to increase with pyrolysis temperature due to an increase in their
fixed carbon content (Bonelli et al., 2003; Bulmau et al., 2010). Thus, pyrolyzed human
faeces and other waste streams such as sawdust char could be blended to improve the
calorific values (Atwijukye et al., 2018). In Kenya, sanitation entrepreneurs like
Sanivation company limited (www.sanivation.com) have ventured into the large scale
production of solid fuels (briquettes) from human faeces collected in dry toilets for
domestic use.
30
2.9 Application of Sawdust Biochar for Solid Fuel Production
Biomass conversion to different forms of bioenergy has recently attracted the attention
of scientific community, policy makers, and industry advocates as a way of meeting the
escalating energy demands with minimal carbon footprint (Reid et al., 2020). The
sustainability of bioenergy production from biomass is definitely dependent on the
availability of the raw materials. Thus, seeking sustainable biomass sources for energy
generation which simultaneously provide an additional benefit of environmental
protection against degradation could be an attractive approach. Sawdust, a woody
biomass residue generated from sawmilling activities by wood-based industries, is
widely abundant in developing countries and have potential to replace energy sources
such as firewood in meeting domestic energy needs (Elehinafe et al., 2017). However,
only a small proportion of the residues are used as fuel because of their high moisture
and low energy density (Njenga et al., 2013). It has been demonstrated that the calorific
values of wood biomass such as sawdust increases with pyrolysis temperature due to an
increase in their fixed carbon content (He et al., 2018). Carbonization could therefore
overcome the drawbacks of using uncarbonized sawdust as a fuel source. Additionally,
carbonized sawdust hereon referred to as sawdust char, could be a suitable material to
blend with faecal char to improve their calorific values.
31
processed raw material is then mixed with binder before briquetting (Ward et al., 2014).
However, if densification pressure is adequate, no extra binder is added (Ngusale et al.,
2014). Uncarbonized material on the other hand is compressed at low pressure when
mixed with binder whereas high compaction pressure and high temperature are required
in the absence of binder for adequate bonding (Asamoah et al., 2016).
For briquettes to compete effectively with other sources of fuel, then its calorific value,
emission levels and ease of transportation and handling are of vital consideration.
Reviewed literature indicates that among the critical factors that affect the quality of
briquettes during production are geometry, compaction pressure, binders, particle size
and moisture.
The shape of the briquettes depends on the discretion of the producers. However it has
been demonstrated that shape has an effect on the combustion (ignition and burning
time) properties of briquettes. Kabok et al. (2018) demonstrated that ignition time which
is the average time required to ignite the composite briquettes was significantly
influenced by the surface area of the briquettes exposed to the airflow. The study
observed that the spherical briquettes with higher surface area had the shortest ignition
time than cylindrical and triangular briquettes. However, the burning time which is the
duration taken to bring a certain amount of water to boiling was insignificantly
influenced by the shape of briquettes. This could be attributed to the fact that the
calorific value of the fuel was the same since the briquettes were of the same materials.
The compaction pressure is usually exerted using a briquetting machine which can either
be a screw or a piston press (mechanical or hydraulic). Compression not only helps in
preventing the briquettes from crumbling (Ugwu, 2013) but also increases briquette
32
density which consequently decreases porosity making it burn longer (Ufot, 2013).
Studies on the effect of compression pressure on heating values of briquettes shows that
higher compression pressures (50 – 150 Mpa) does not significantly influence the
calorific values (Tanui et al., 2018). Mambo (2016) using lower compression pressures
(2 – 8 MPa) to produce briquettes from carbonized maize cobs, observed that the
briquettes produced could withstand crumbling during handling and transportation,
however the author recommended 8 Mpa as the best compression pressure. The findings
suggest that focus should be on producing quality briquettes at low pressures to
minimize energy consumption during production process.
Binders are sticking or agglomerating materials that are mixed with biomass material
that lacks plasticity and densified to form strong briquettes. The quality and quantity of
binders used is of key consideration in the production process since they affect the
strength, thermal stability, combustion performance and cost of briquette (Altun et al.,
2001). Generally, the desirable properties of briquette binder is that they should form a
strong bond, be pollution free, have no effect on the heat released and combustibility of
the biomass materials, environmentally acceptable and economically available
(Githeko., 2013; Habib et al., 2014; Ngusale et al., 2014; Zhang et al., 2018). According
to Zhang et al. (2018), briquette binders can be divided into three types based on the
difference of their material composition, namely; (a) organic binder, (b) inorganic
binder, and (c) compound binder.
Clay, lime, plaster, cement, and sodium silicate are common types of inorganic
briquette binders (Altun et al., 2001; Njenga et al., 2013). These binders have many
excellent advantages, such as good thermal stability, strong adhesion, low pollution
level, low-cost, good sulphur retention, and good hydrophilicity while the disadvantages
33
exhibited are high ash content, low gross calorific value and poor water repellency/
proofing (Zhang et al., 2018).
Briquettes bonded with organic binders have high crushing strength thus durable, lower
ignition temperature, high heating values and emit less smoke (Ngusale et al., 2014;
Kabok et al., 2018; Aransiola et al., 2019). However, organic binders are reported to
decompose easily at high temperatures, some are costly, in addition to possessing poor
water repellency and poor thermal stability (Altun et al., 2001; Zhang et al., 2018).
Among the commonly used organic binders for briquetting in developing countries are
starch sources (cassava flour, wheat flour, potato starch, rice flour, corn starch, and
maize flour), gum arabica, waste papers, and molasses (Njenga et al., 2013; Ngusale et
al., 2014). However, the challenge of using starch sources is that they could create
competition with food for human consumption and therefore making the entire process
unsustainable.
34
% , and higher heating value of 15 MJ/kg. Despite the potential of molasses in
improving the physical and combustion properties of briquettes, it has been reported that
the fructose content in molasses is highly hygroscopic (Davis, 1995). The hygroscopic
property of molasses could result in high moisture content in briquettes densified with
molasses consequently making their ignition difficult. Hence further research on
alternative organic binders is necessary. Studies using molasses as a binder have
reported that 10% of molasses by weight can produce durable briquettes of higher
heating values than fuel wood (Githeko., 2013; Habib et al., 2014; Kabok et al., 2018;
Tanui et al., 2018).
There are advantages and disadvantages associated with both organic and inorganic
binders. For instance, briquettes produced with inorganic binders are more thermally
stable than those produced with organic binders. However, the briquettes produced with
compound binders have low heating values, low combustion efficiency and high ash
content compared to briquettes produced with organic binders. The concept of
compound binder involves mixing two or more binders to combine all advantages of
different kind of binders (Zhang et al., 2018). Examples of these binders include,
cements (85% – 90%), hydrated lime (5% – 10%), polyvinyl alcohol (5% – 10%), coal
tar pitch phenolic resins, and corn straw (70%), sodium hydroxide (7%), and MgCl/MgO
(23%) (Zhang et al., 2018).
The particle sizes of the raw materials densified to form briquettes affect their physical
and thermal properties. Studies have reported that increase in particle size of constituent
materials increases calorific value, reduces ash content and increases thermal efficiency
(Davies, 2013; Tokan et al., 2014; Mambo, 2016). Thus, medium sized particles should
be used in briquetting.
35
2.11.5 Effect of Moisture
Heating value of fuel can be defined as the amount of heat produced by complete
combustion of fuel and it is measured as a unit of energy per unit mass or volume of
substance (Xu & Yuan, 2015). It is normally expressed by the higher (gross calorific
value) and lower heating values. The higher heating value (gross calorific value) is
measured experimentally using a bomb calorimeter. Uzun et al. (2017) defines higher
heating value of fuel as equal to the amount of heat released when a unit mass of the fuel
is burnt completely, accounting for the enthalpy of condensation of liquid water as a
combustion product under standard conditions. That is, the heat of condensation of the
water is included in the total measured heat.
The proximate analysis serves as a simple means for determining the behaviour of a
solid biomass fuel when it is heated. It determines the contents of moisture, volatile
matter, ash and fixed carbon of the fuel (ASTM, 1984). The proximate parameters
provide the potential efficiency and durability of the briquettes that will be produced as
discussed in the following sub sections.
36
2.12.2.1 Fixed Carbon
Fixed carbon is the solid combustible carbon that remains in the pyrolyzed material after
volatile matter is released. Briquettes made of materials with higher fixed carbon content
therefore combust and release heat energy for a longer time and have higher heating
values (Falemara et al., 2018).
Volatile matter refers to the non-water gases formed from the thermal degradation of a
material. Materials with higher volatile content ignite faster at lower temperatures and
release higher amount of heat due to higher oxygen consumption rates (Sha et al., 2021).
Higher moisture content in feedstock may increase the production cost in terms of
energy, due to the fact that more energy is required to reduce the water content during
drying and densification. Biochar however have lower moisture content and only little
moisture is required to assist the bonding process of the feedstock. Higher moisture
content in briquettes reduces their heating value (Yank et al., 2016).
37
2.12.3 Ultimate Analysis
Oxygen (%) = 100 - Carbon (% Dry Basis) - Hydrogen (% Dry Basis) - Nitrogen (% Dry
Basis) -Sulphur (% Dry Basis) (2.23)
A flue gas emission from biomass combustion refers to the gas product resulting from
burning of biomass solid fuel (Pilusa et al., 2013). Combustion of biomass normally
releases gases which can pose health hazard to humans depending on their inherent
toxicity and duration of exposure. A good fuel should therefore emit gases below the
threshold limit value as recommended by the occupational safety and health
administration (OSHA). OSHA (2010) defines threshold limit value as the upper
permissible concentration limit of the gases believed to be safe for humans even with an
exposure of 8 hours per day, 5 days per week over a period of many years. OSHA limits
are based on industrial settings, however, short term exposure limits for various toxic
gases which is applicable for domestic settings are also defined by organizations such as
WHO, Environmental Protection Agencies (EPA), and National Institute for
Occupational Safety and Health (NIOSH) (Schieb, 1976; WHO, 1981; EPA, 2002).
Among the critical emissions of concern to human health are the carbon monoxide (CO),
hydrogen sulphide (H2S), and nitric oxide (NO).
38
CO2 is produced as a result of complete combustion of biomass and therefore regarded
harmless to human health , while CO emission results from incomplete combustion
(Obernberger et al., 2006). Inhalation of CO even for shorter exposure period (< 1 hr)
can cause disastrous health effects such as limited oxygen supply in the blood, fatigue,
headache and sudden death (Townsend & Maynard, 2002). Fuels emitting CO
concentrations above the critical limit of 35 ppm allowed for human exposure for one
hour are normally considered unsafe for indoor heating (EPA, 2002). Proper
ventilation is thus necessary to enhance oxidation of carbon during biomass
combustion.
Hydrogen sulphide (H2S) gas is among the common hazardous substances whose
concentrations at elevated levels (1000 – 2000 ppm) can lead to immediate death (Frame
& Schandl, 2015). It has a characteristic offensive odour (rotten eggs) which makes it to
be easily detected at very low concentrations by sense of smell. Many countries have
short term exposure limits for ambient H2S with the WHO recommending its
concentrations not to exceed 0.005 ppm within 30 minute exposure time in order to
avoid substantial complaints about odour annoyance (WHO, 1981). Very short lived
peak concentrations have also been reported to cause discomfort (Collins & Lewis,
2000). Also, Bhambhani and Singh (1991) reported that exposure of human population
to 2.5 – 5 ppm of H2S after 15 minutes resulted to coughs, throat irritation and impaired
oxygen uptake in the blood. It is worth noting that the corrosive nature of H2S (Malone
Rubright et al., 2017) may also pose challenge by corroding metal wares.
The most important oxides of nitrogen that are considered air pollutants are nitric oxide
(NO) and nitrogen dioxide (NO2) (EPA, 1971). NO is a colourless and odourless gas
which is readily oxidised in the atmosphere to form NO2 which is a reddish-brown gas
with a characteristic pungent smell and is more toxic (EPA, 1971; Ku, 1991). NO is
classified as a respiratory irritant with inhalation being the main route of exposure,
however eye and skin irritation also occurs if one is exposed to NO (Ku, 1991).
Although little scientific data is available regarding exposure to NO compared to NO2,
its concentration above 100 ppm is considered dangerous to life or health according to
39
National Institute for Occupational Safety and Health (NIOSH) (Schieb, 1976).
Monitoring emission of these gases from combustion of the briquettes will give useful
information regarding their safety for indoor household use.
Thermogravimetric analysis (TGA) has often been employed in kinetic modelling of the
thermal degradation of biomass-type wastes. This technique determines the thermal
stability of materials by monitoring the change of weight that occurs as a sample is
heated. For example, the thermal behavior of various biomass sources have been studied
including, rice husks, corn residues, and sawdust (Sittisun et al., 2015; Lim et al., 2016;
Mishra & Mohanty, 2018). The information gained from thermal degradation can also
be used in conducting kinetic analysis to determine the activation energy of the material
(Sittisun et al., 2015; Lim et al., 2016; Mishra & Mohanty, 2018). That is, the
experimental data from TGA is routinely fitted to different reaction models and the one
that offers the best fit chosen to be the best that describes the kinetics of the process.
Two approaches that can be used to adequately describe the kinetics of a given reaction
process under both isothermal and non-isothermal conditions are; model-fitting kinetics
and model-free (iso-conversional) kinetics (Vyazovkin & Wight, 1998). The model-
fitting method is based on a single heating rate and therefore not able to account for
multiple reaction steps. This is a disadvantage because the activation energy varies with
the heating rate, due to mass and heat transfer effects. The model-free or iso-
conversional method on the other hand uses multiple heating rates allowing for a change
in mechanism during reactions hence accounting for multiple reaction steps. The iso-
conversional method is used to estimate the activation energy independent of the
reaction mechanism, with no prior assumptions about the analytical form of f (α).
Model-free methods are thus recommended by the ICTAC Kinetics Committee due to
the mentioned advantages (Vyazovkin et al., 2011). Kinetic data are of major interest in
the technological development of processes related to energy production (pyrolysis,
gasification, and combustion). Most studies on TGA of biomass materials have been
performed under different heating rates, and under different reaction atmospheres (e.g.,
40
nitrogen, oxygen, and air) (Farrow et al., 2013; Lim et al., 2016; Yacob et al., 2018)
leading to different thermal behaviors. For example, Ma et al. (2018) has demonstrated
that palm kernel shell loses more weight under air (~95.22%) than under nitrogen flow
(~72.39%). Zhang et al. (2019) reported that cattle manure exhibited lower activation
energy when combusted under oxy-fuel atmosphere (CO2/O2) than under air conditions
(N2/O2). Similar studies on human faeces (Somorin et al., 2020; Yacob et al., 2018) are
available in the literature albeit with variations, especially on how blending affects their
thermal decomposition properties and kinetic behaviour. For instance, when human
faeces is blended with wood biomass in different ratios, the thermal behaviour of the
blend markedly vary with the wood biomass content as explained by Fidalgo et al.
(2019). Conversely, Somorin et al. (2020), acknowledged that regardless of the wood
biomass content, blended samples exhibited similar thermal decomposition profile to
that of human faeces. Additionally, Somorin et al. (2020), and Yacob et al. (2018) found
activation energy values for human faeces to be between 122 – 382 kJ/mol, and 141 –
409 kJ/mol, respectively, for the same degree of conversion (10 – 90 %). The
combustion properties of these biowastes vary depending on factors such as the
composition, nature, particle size, moisture, and the volatile content. Although the
reported studies have focused on the thermal behaviour and kinetics of dried human
faeces under nitrogen atmospheres, investigating the behaviours of these materials after
pyrolysis is important since the practice of solid fuel production from faeces is based on
densification of charred faeces which is considered pathogen free. Moreover, studying
their combustion characteristics under air conditions depicts the practical properties of
these materials.
The summary of related literature reviewed presented in Table 2.2 highlights the
potential of biochars such as pineapple peel biochars and geological materials such as
lateritic soil in adsorbing nutrients from human urine and the capability of response
surface methodology in accurately predicting the optimum conditions for achieving
maximum adsorption of nutrients from aqueous solutions. The summary also delved into
41
the factors affecting thermal stability of biomass materials and the suitability of model-
free kinetic analysis methods over model-fitting kinetic analysis methods in determining
activation energies of materials. The short-term limits of exposure to carbon monoxide
(CO), nitric oxide (NO) and hydrogen sulphide (H2S) applicable for domestic settings
are also highlighted.
42
2.16 Research Gap
Despite the progress made in adsorption technology for recovery of nutrients from
wastewater using low cost adsorbents, the potential of pineapple peel biochars and
lateritic soils for nutrients recovery from human urine haven’t been reported despite the
wide availability. Intricate studies on how the physicochemical properties of the
pineapple peel biochars and lateritic soils influence adsorption process and how their
surfaces interact with the NH4+-N and P in human urine gave an insight on their
potential applicability for large scale recovery of these nutrients from human urine.
Moreover, estimating their NH4+-N and P adsorption capacities from human urine will
enable their optimal use for producing enriched slow-release fertilizers or medium for
crop growth. Also, although the concept of briquettes production from human faecal
chars blended with sawdust char has been reported (Kabok et al., 2018), investigations
on the thermal decomposition characteristics and reaction rates of faecal char and their
blend with sawdust char under air atmospheres haven’t been reported. Hence this study
reveals the practical thermal stability and reaction rates of faecal char and their blend
with sawdust char when used as energy source for household cooking. Moreover, studies
on emission of toxic gases during combustion of briquettes containing faecal char have
not been reported by previous studies and this study therefore adds vital information that
depicts the safety of the briquettes for indoor household cooking and heating.
Conceptual framework is a model that employs the use of drawing or diagram to explain
the interrelationships between the independent and dependent variables. The conceptual
framework for this study can be represented as shown in Figure 2.1
43
Figure 2.1: Conceptual Framework of the Study Variables
44
CHAPTER THREE
3.1 Determining the Behaviour and Capacity of Pineapple Peel Biochar and
Lateritic Soil in Adsorbing Ammonium Nitrogen and Phosphorous from Human
Urine
Fresh undiluted urine was collected from volunteers and used directly without further
purification. Fresh pineapple peels were collected from vendors, dried in a greenhouse
for 48 hours and then carbonized at 400 °C under vacuum for 4 hours. The resulting char
was crushed and particles with ≤ 300 µm sieved with 300 micron stainless steel sieve.
The powder was then kept in an air tight plastic container for further analysis. Lateritic
soil was excavated 20 cm from the ground surface, and impurities such as leaves and
roots removed. The soil was then dried in an oven at 104 ±1 °C for 24 h under vacuum.
The dried soil was then crushed, sieved to ≤ 300 µm and kept in an airtight plastic
container for further studies.
Thermogravimetric Analysis (TGA) of pineapple peel biochar and lateritic soil was
performed on a TGAQ500 thermogravimetric analyzer (TA instruments, USA) from 30
to 800°C at a heating rate of 5°C/min under flowing N2. FTIR spectra were obtained
using a Spectrum One attenuated total reflection (ATR)-FTIR spectrometer. Spectra for
each run were averaged over 40 scans at a 4 cm-1 resolution in the range 4000 – 650 cm-
1
. X-ray powder diffraction (XRD) patterns were recorded at room temperature on a
Bruker AXS D8 Advance X-ray diffractometer using CuKα radiation (λ = 1.54056 Å)
operating at 45 kV, and 40 mA in the 2θ scan range of 5 – 90 with a step size of 0.009
and a total measuring time of 3 hours. The pH of lateritic soil (LS) and pineapple peel
biochar (PPB) were determined by shaking 500 mg of sample powders with 20 mL
45
deionized (DI) water for 30 min. The samples were let to settle for 1 hour before
recording the pH using a pH electrode Seven Compact pH meter S220 (Mettler Toledo,
Switzerland). The Brunauer-Emmett–Teller (BET) surface area, pore volume, and pore
size distribution of the samples were analyzed using a Micromeritics ASAP 2020
analyzer. Prior to the measurements, samples (200 – 400 mg) were degassed under
vacuum at 150°C overnight to eliminate moisture before N2 adsorption. The surface
areas of the samples were calculated by the BET method in relative pressure (p/p0) range
of 0.05 – 0.30 at 77 K. The morphological and elemental mapping studies were
performed by a LEO-1530 scanning electron microscope (SEM) equipped with an
Oxford AZtec energy dispersive spectrometer (EDS).
Nutrient removal ability of lateritic soil (LS) and pineapple peel biochar (PPB) was
evaluated using parallel experiments. Dilution of urine to obtain six varying
concentrations of the solute (NH4+-N and P) was done using DI water. 0.2 g of LS and
PPB were separately added to 25 mL of urine in a beaker at 21 °C, and the solution
shaken in a thermostatic mechanical shaker (Heidolph unimax orbital shaker 2010,
Germany) at a rate of 100 rpm. Samples were filtered using Whatman filter paper no. 42
to separate the adsorbents from the urine at time intervals of 30, 60, 90, 120 min. The
control samples consisted of human urine with no adsorbent added. The NH4+-N and P
levels in the filtrates were measured for all the treated samples and control samples.
Equation 3.1 was used in calculating the amount of nutrients adsorbed at various time
intervals, qt, until the equilibrium concentration, Ce, was reached.
(Co Ct )V
qt (3.1)
W
46
Where:
The NH4+-N concentration in urine samples was determined using Kjeldahl method ISO
standard. Briefly, 2.5 mL of sample was put in a distillation glass tube, and the pH
adjusted to 9.5 by 0.1M NaOH and 5 mL of borate buffer. 25 mL of 2% boric acid was
then added to a separate beaker for collecting the distillate. As the distillation process
continued, the volume of the distillate was monitored until it reached 125 mL upon
which the distillation unit (Kjeltec 1002 System Distilling Unit, Sweden) was switched
off. The pH of the collected sample was measured and 0.01M H2SO4 added until a pH of
5.1 was reached. The final volume titrated was used in calculating the NH4+-N
concentration in the sample using Equation 3.2. Normally a blank sample (DI water) was
also subjected to the same procedure to verify the accuracy of the method since DI water
is theoretically expected to have no ions. This procedure was followed each day prior to
analyzing urine samples. In case of any level of NH4+-N detected in the blank samples, it
was subtracted from the values read in the urine samples according to Equation 3.2
given as:
Where:
47
a = the consumption of H2SO4 when titrating sample (mL),
b = the consumption of H2SO4 when titrating the blank (DI water) sample (mL),
Phosphorous in the samples was determined according to Finnish standards (SFS 3026)
for determination of total phosphorous in water (SFS, 1986).
Phosphorous stock solution (50 mg/L P) was prepared by dissolving 0.2197 g of dry
potassium di-hydrogen phosphate (KH2PO4) in deionized (DI) water. 10 mL of 4 M
H2SO4 was then added and the solution diluted to a total volume of 1 L. Phosphorous
working solution (1/50 dilution of the stock solution) was prepared by diluting 10 mL of
the stock solution with DI water to a total volume of 500 mL.
The first step in the SFS 3026 method for determining phosphate concentrations was to
develop the standard calibration curve using five working standards. In this method,
calibration standards are prepared by pipetting different amounts (1, 2, 10, 25 and 75
mL) of phosphate working solution into 100 mL volumetric flasks. They are then
acidified with 1.0 mL of 4.0 M H2SO4 and diluted with DI water to the 100 mL mark.
The calibration solutions therefore contained 10, 20, 100, 250 and 750 µg/L of P. Blank
or zero samples were also prepared by pipetting 10 mL of 4.0 M H2SO4 in 1000 mL
48
volumetric flask and then adding DI water to the 1000 mL mark. Each sample
investigated was prepared in three replicates. 5 mL of potassium persulfate (5g K2S2O8
per 100 mL DI water) solution was added into each bottle to transform phosphorous into
phosphates. The bottles were autoclaved at 120 °C for 40 min to expedite the
transformation process and then cooled to room temperature, 21±1 °C. 1 mL of each
ascorbic acid (C6H8O6) and ammonium molybdate (NH4)6Mo7O24) reagents were then
added for color development. UV-Spectrophotometer (880 nm wavelength) was used to
obtain the total P readings. The absorbance and concentration values of blank (zero)
samples were read first followed by absorbance of the standards whose concentration
had been predetermined in the experiment. The plot of absorbance against concentration
of the blanks and the standards gave a calibration curve from which the P concentration
of the samples of unknown concentration was calculated based on the absorbance
readings obtained by the spectrophotometer.
For the human urine whose concentrations of P were unknown, 0.5 mL of urine samples
from the original urine samples (control) and the adsorption experiment (filtrate) were
pipetted into 500 mL volumetric flasks, acidified with 5 mL of 4M H2SO4 then topped
with DI water to the 500 mL mark. 25 mL of each of the prepared samples were
transferred into clean autoclave bottles. Subsequent procedures already described in this
sub-section (addition of persulfate, autoclaving, color development) were then followed,
and the absorbance reading obtained from the spectrophotometer used to calculate the
concentration of the urine samples from the calibration curves.
Response Surface Methodology (RSM) was used to optimize NH4+-N and P adsorption
on PPB and LS, respectively, and investigate the interactive effects of contact time,
adsorbent loading and agitation speed on the adsorption process. The optimisation was
49
conducted using “Design-Expert® software, version 12, Stat-Ease, Inc., Minneapolis,
MN, USA, www.statease.com”. Central composite circumscribed design was used since
it is the most effective in estimation of model parameters in addition to being rotatable.
It consists of three critical points notably; factorial (2k), axial (2k) and center points (six
replicates), where k is the number of independent variables. Repetition of center points
provides better estimate of pure error. The total number of experimental runs is the
summation of factorial points, axial and the six center points. In this study, k = 3
(contact time, agitation speed and adsorbent loading) resulting into 8 factorial points and
6 axial points. The total experimental run (n) is the summation of factorial points, axial
points and the centre points. Hence n = 8+6+6=20. The selected independent variables
are coded according to Equation 3.3 given as:
Xi Xo
xi (3.3)
X
Where:
Each variable was studied at 5 levels coded as -1.682, -1, 0, 1, +1.682, where (-1) is low
level, (+1) is high level, (0) is the center point, ±1.682 are codes for points defined by
the design expert software to be below and above the respective designated low and high
levels in each factor studied. The parameters and levels for the experimental design used
in this study are summarized and presented in Table 3.1.
50
Table 3.1: Parameters and Levels for the Experimental Design
The RSM model correlates the dependent variables to the independent variables
considered through second-order polynomial response equation given as in Equtaion 3.4
Where:
βo = the value of the fixed response at the center point of the design
βi, βii,and βij = the linear, quadratic and interaction effect regression terms,
respectively
The analysis of variance (ANOVA) was conducted to study the significance of the
quadratic model, while the value of the coefficient of determination (R2) was used to
determine the reliability of the quadratic model in fitting the experimental data and
accurately predicting the output.
51
3.2.3 Graphical Plots and Numerical Optimisation
The 3D response surfaces were then drawn to visualize the individual and the interactive
effects of the independent variables on the adsorption of phosphorous. The 3D plots are
provided in the Figure D1, and E1 in Appendix IV, and V, respectively. To perform
numerical optimization for the adsorption of NH4+-N and P on PPB and LS,
respectively, the desirability function approach was followed. The desirability is a
dimensionless entity that represents the closeness of a response to its ideal value and
usually ranges from 0 to 1 with d = 0 being unacceptable, and d = 1 indicating that the
model response is equal to that of the target value (Li et al., 2007; Myers et al., 2004).
The RSM was set-up with the goal being to optimise adsorption so that the model could
determine the values of the input variables that yield the highest adsorption capacity of
NH4+-N and P on PPB and LS, respectively.
Faecal matter was collected from a dry toilet and dried in a greenhouse under air for one
week attaining constant moisture content. Sawdust was obtained from saw millers and
similarly dried in a greenhouse for a week attaining constant moisture content. Both the
dried human faeces and sawdust were then pyrolyzed in an electric muffle furnace
FUW232PB (Toyo Seisakusho Kaisha, Ltd) at 350°C for 2 hours under a vacuum. The
resulting char was crushed and particles with ≤ 300 µm sieved and stored in air tight
plastic containers for further analysis. In addition, the two samples (pyrolyzed human
faeces (hereon denoted as faecal char), and sawdust char) were blended in the weight
ratio of 50:50 wt. %, hereon referred to as blend. Previous studies on production of
briquettes from faecal char and sawdust char reported that equal weights (50:50 wt. %)
produced briquettes of desirable heating values comparable to the fuel wood (Kabok et
al., 2018; Atwijukye et al., 2018; Ward et al., 2014), and thus informed the ratio adopted
52
in this study. The morphological features of faecal char and sawdust char were observed
using a LEO-1530 scanning electron microscope (SEM) operated at 1 kV. Fourier
Transform Infrared (FTIR) Spectra of the samples were recorded on a spectrometer in
the range 4000 – 650 cm-1. An oxygen bomb calorimeter (Yoshinda 1013J) was used to
determine the gross calorific values of the samples, while the proximate parameters
(fixed carbon, volatile matter, moisture content, and ash content) were determined
according to ASTM D1762-84 standard test method for chemical analysis of wood
charcoal (American Society for Testing and Materials, 1984). Elemental analysis (C, H,
N, and O) of faecal char and sawdust char was undertaken using Elemental Analyzer
(AAS iCE 3300).
The combustion characteristics and kinetic behavior of faecal char, sawdust char, and
blend were investigated using a thermogravimetric analyzer (TA Instruments TGA-
Q500) under non-isothermal conditions. For each TGA run, about 10 mg of sample was
placed in an Aluminium crucible and heated from 30 to 800 °C under flowing air (60
mL/min), at heating rates of 5, 10, 20, and 40 °C/min. In addition, the effect of sample
atmosphere on decomposition was also tested by running the three samples under N2 at a
heating rate of 5 °C/min.
Kinetic analysis can give insights into the reaction mechanism of a given process. The
d
rate of a chemical reaction is described by the general kinetic rate equation as
dt
presented in Equation 3.5 given as:
d
k (T ). f ( ) (3.5)
dt
53
Where:
t = time
α = the conversion
The degree of conversion, 0 ≤ α ≤ 1, which is a measure of the extent of the reaction can
be defined as in Equation 3.6
mo mt
(3.6)
mo m f
Where:
The rate constant is generally expressed by the Arrhenius equation (Equation 3.7)
E
k (T ) A . exp a (3.7)
RT
Where:
54
R= the universal gas constant (8.314 J/mol·K).
Substituting the Arrhenius expression into Equation 3.5 gives the standard kinetic
equation under isothermal or non-isothermal conditions shown in Equation 3.8
d d E
A . exp a . f ( ) (3.8)
dt dT RT
Where:
dT
β = the heating rate (K time-1)
dt
d E
T
A
g ( ) exp a dT (3.9)
0
f ( ) T0 RT
Where:
g(α) = the integral kinetic function or integral reaction model when its form is
mathematically defined
In this study, the activation energies, Ea, were estimated using the iso-conversional
methods of Kissinger-Akahira-Sunose (KAS) and Flynn-Wall-Ozawa (FWO).
The approximation of Equation 3.9 by Coats & Redfern (1964) give rise to the Equation
3.10 applied by the Kissinger-Akahira-Sunose method (KAS) (Kissinger, 1957; Akahira,
T .; Sunose, 1971).
A . R E
ln 2 ln (3.10)
T E . g ( ) RT
55
At a constant α, the apparent Eα can be estimated from the slope (-Eα/R) of the straight
line obtained by plotting ln(β/T2) versus 1/T.
AE E
ln ln 5.331 1.052 (3.11)
Rg ( ) RT
At a fixed value of α, plotting lnβ versus 1/T gave a straight line. The activation energy
was determined from the slope -1.052Eα/R over a series of α.
Human faeces was obtained from a dry toilet and dried in a greenhouse for 7 days
attaining constant moisture content, while sawdust was obtained from saw millers and
also dried in a greenhouse for 7 days attaining constant moisture content. The charcoal
used in this study was produced in a tradition kiln as is the practice in African countries.
The charcoal was from acacia tree of species (Acacia nilotica) which is among the
preferred tree species for firewood in dryland areas in Kenya (Oduor et al., 2019).
Molasses used as a binder originated from a sugar processing factory.
Both the dried human faeces and sawdust were pyrolyzed in an electric muffle furnace
FUW232PB (Toyo Seisakusho Kaisha, Ltd) at 350 °C for 2 hours under a vacuum. The
resulting char was crushed and particles with ≤ 300 µm sieved and stored in air tight
56
plastic containers. Analysis of Elements (C, H, N, S, and O) of faecal char and sawdust
char was undertaken using elemental analyzer (AAS iCE 3300).
An electronic weighing machine was used to obtain equal weights (450 g each) of both
faecal char and sawdust char and then mixed homogeneously in 50:50 wt. % ratio. 100 g
of cane molasses representing 10 wt. % of the overall weight was added to the charred
samples and mixed homogeneously to stick the particles together. A cylindrical plastic
mould of internal diameter 23 mm and height of 100 mm and a cylindrical wooden die
of diameter 22 mm were used to compress the materials (15 g) fed into the mould upon
application of 5 MPa pressure from a hydraulic press. With the applied pressure and the
diameter of the dies, cylindrical briquettes of diameter 22mm and height of 60 mm were
ejected from the mould and dried in an oven at 104 ±1°C for 24 hours to a constant
weight.
An oxygen bomb calorimeter (Yoshinda 1013J) was used to determine the gross
calorific values of the produced briquettes, and charcoal, while the proximate parameters
(fixed carbon, volatile matter, moisture content, and ash content) were determined
according to ASTM D1762-84 standard test method for chemical analysis of wood
charcoal (ASTM, 1984) as detailed in the following sections:
A sample of material was weighed and put in the oven at 104 ±1°C for 24 hours after
which it was removed and weighed again. The difference in weight divided by the dry
weight of the material gives the moisture content as per the gravimetric method.
57
3.4.4.2 Ash Content Determination
A specimen of the dry briquettes was ground, weighed and then placed in a crucible and
heated gradually in furnace at controlled temperatures to about 600 °C. The sample was
burned until all the carbon was consumed and the residual ash attained constant weight.
1.0g of oven-dried specimen was weighed and then placed in platinum crucible with a
tight- fitting lid and heated in a furnace at 930 – 970 °C for about 7 minutes while
occasionally stirring with a wire. After cooling in a desiccator the weight of the residue
was taken. The loss in weight was then recorded as the weight of volatile matter in the
sample.
Fixed carbon was determined using the data previously obtained in the proximate
analysis as shown in Equation 3.12:
Where:
FC = fixed carbon
The concentration of CO, CO2, H2S and NO emitted from charcoal was measured by
burning 450 g of the charcoal in a small-sized common energy saving cook stove called
Kenya Ceramic Jiko (KCJ) as practiced by households. Co-combustion was done by
58
burning equal amount (225 g) of charcoal and briquettes in the cook stove which
amounted to a total weight of 450 g of fuel filled in the cook stove. A conical shaped
chimney/stack made of aluminium metal of 1.3 m height was fabricated such that the
larger diameter could fit the top of the cook stove to minimise escape of gases before
sampling, while the top of the chimney was left open to allow gases to escape. A circular
opening of 1.0 cm diameter was made on the chimney to act as a sampling pot for the
gases. The height from the cook stove to the sampling pot was 100 cm. E8500P
industrial integrated emissions system combustion gas analyser was used to monitor the
gas concentrations and gas temperatures with the sensor positioned at the sampling pot.
A schematic diagram showing the experimental set-up is presented in Figure 3.1.
59
CHAPTER FOUR
4.1 Behaviour and Capacity of Pineapple Peel Biochar and Lateritic Soil in
Adsorbing Ammonium Nitrogen and Phosphorous from Human Urine
4.1.1 Composition and Structures of Lateritic Soil and Pineapple Peel Biochar
To identify the phases, X-ray diffraction patterns were collected on sample powders at
ambient conditions. The XRD patterns of LS and PPB samples are shown in Figures
4.1a and 4.1b, respectively. The LS powder consisted of several phases, also seen by
EDS analysis in Figure A4 (a-c). These phases were identified as quartz ((SiO2)-
ICSD16331) with main peaks at 2θ = 20.6, 26.7, 36.5, and 50, kaolinite
((Al2Si2O5(OH)4-ICSD30966) with main peaks at 2θ = 12.2, 21.5, 38.5, 45.8, 55.5, and
62.1, goethite-aluminian ((Fe0.83Al0.17)O(OH)-ICSD109411) with main peaks at 2θ = 18,
60
21.3, 26.5, 33, and 37, and illite ((K(Al4Si2O9(OH)3-ICSD90144) with main peaks at 2θ
= 12.4, 21.6, 24.8, 38, and 55.8.
The XRD pattern of PPB (Figure 4.1b) shows an amorphous structure. In the XRD
pattern, sharp peaks found at 2θ = 29.4, 35.9, 39.4, 43.2, and 47.8, and at 2θ = 14.5,
25.6, 29, 37.9, 39.4, 40.9, and 47.4, are attributable to calcite (CaCO3-ICSD18164), and
scapolite (Na, Ca)4 (Al, Si)12O24(Cl, CO3, SO4)-ICSD7102), respectively, as impurity
phases.
Figure 4.1: XRD Powder Patterns of (a) Lateritic Soil, and (b) Pineapple Peel
Biochar
The morphology and particle sizes of LS and PPB from SEM are shown in Figures 4.2a,
and 4.2b respectively. The PPB consists of microparticles with elongated honeycomb
shapes that are uniform and highly porous. LS on the other hand, consist of
nanoparticles which aggregate into different morphologies attributable to the presence of
different structural phases.
61
Figure 4.2: SEM Images of (a) Lateritic Soil, and (b) Pineapple Peel Biochar
The pore volume, the Brunauer-Emmett-Teller (BET) surface area (SBET), and the
average pore diameter of the two samples are presented in Table 4.1. Fittings of the BET
equation to the adsorption isotherms of nitrogen at 77 K give the estimated surface areas
of 140 m2/g, and 44 m2/g for PPB and LS, respectively. Also, the average pore diameter
of PPB obtained was 23 nm, which indicated that its surface was microporous, in line
with SEM results.
Table 4.1: Nitrogen Sorption Properties of Lateritic Soil and Pineapple Peel
Biochar
Sample BET surface area, Total pore volume, Average pore diameter
SBET (m2/g) Vtot (cm3/g) (nm)
LS 44 0.122 11
PPB 140 0.082 23
The surface area and total pore volume values observed herein for LS are generally
within range of those reported for similar materials. For instance, SBET (m2/g) and Vtot
(cm3/g) values of lateritic soil derived from different geographical regions were reported
as 17.5 – 18.5 m2/g and 0.02 – 0.05 cm3/g, and 71 – 182 m2/g and 0.07 – 0.35 cm3/g
(Maiti et al., 2012, 2013). In the case of biochars, a wide variation in the surface areas
and total pore volume values across the literature has been reported making their
comparisons difficult. The variations could be ascribed to the differences in pyrolysis
62
temperatures, which affects the physicochemical properties of biochars (Kim et al.,
2012).
The FTIR spectra in Figure 4.3 show different functional groups of the samples. For LS,
the sharp absorption band over the range 3698–3617 cm-1 are assignable to –OH group,
and appears due to the presence of strongly bonded water molecules. While the band at
1630 cm-1 is due to H–O–H bending mode in the weakly surface adsorbed or trapped
water molecules in the large cavities of the polymeric framework (Lemougna et al.,
2014). The bands in the region between 1376 and 752 cm-1 are associated with Si–O–Al,
Si–O–Fe, Si–O, Al–OH, and Fe–OH vibrations, and are typical of kaolinites (Kakali et
al., 2001). The band at 687 cm-1 is due to Fe–OH vibration (Rocha et al., 1990). Similar
types of bands have also been observed by Maiti et al.(2012).
For PPB, the broad peaks observed between 3400–3000 cm-1 could be attributed to the
stretching mode of O–H, indicating the presence of water (Fu et al., 2016). The band
around 2915 cm-1 is due to aliphatic C–H stretching vibrations in biopolymers (Cantrell
et al., 2012), and corresponds to the presence of cellulose and hemicellulose (Usman et
al., 2015). This may indicate incomplete carbonization (Shakya & Agarwal, 2019). The
strong band at 1589 cm-1 is due to alkyl and alkene stretching groups (C–H, C=C, and
C–C) (Kim et al., 2012). The weak bands in the region 872–752 cm-1 are credited to the
bending of C–H groups of the aromatic rings (Fu et al., 2016).
63
Figure 4.3: FTIR Spectra of Lateritic Soil, and Pineapple Peel Biochar
64
In the case of PPB, gradual weight loss was observed with increasing pyrolysis
temperature. The first mass loss step ≤ 100 °C corresponds to the release of physisorbed
water molecules (~4 wt.%). The weight loss between 110 °C ≤ T ≤191 °C may be
attributed to the removal of strongly bound water ( Li et al., 2017). The mass decrease in
the range 191 °C ≤ T ≤ 280 °C was assigned to degradation of hemicellulose, while
thermal degradation of cellulose occurs between 281 °C and 380 °C (Lee et al., 2013).
The breakdown of lignin and removal of hydrogen and oxygen (Deng et al., 2014)
corresponds to the significant mass loss between 500 and 650 °C, accompanied by a
DTG peak at 550 °C. The total weight loss was found to be ~23 wt.%. Similar finding
have also been reported by Shakya and Agarwal (2019).
Figure 4.4: Thermogravimetric Analysis (TGA) Curves (solid lines) and DTG
Curves (dashed lines) of Lateritic Soil and Pineapple Peel Biochar at a Heating
Rate of 5 °C/min Under N2 Flow
In summary, PPB could be better suited for NH4+-N and P adsorption due to its large
surface area and microporous nature as revealed by BET and SEM results. Also, this
65
material has a high amount of carbonates and organic matter as seen from TGA data. LS
on the other hand, consist of a mixture of phases, and is more thermally stable according
to EDS, XRD, and TGA results.
4.1.2 The Effect of Contact Time and the Initial Concentration on Adsorption of
Ammonium Nitrogen from Human Urine on Lateritic Soil and Pineapple Peel
Biochar
To evaluate the effect of contact time and the initial concentration of NH4+-N adsorption
on LS and PPB, the experiments were carried out with varying initial NH4+-N
concentrations (156.88, 161.36, 341.77, 398.92, 451.59, and 569.24 mg/L) at 21°C, pH
= ~7.27 ± 0.202, dosage of 0.2 g in 25 mL of urine, and at varying contact time of 30,
60, 90, and 120 minutes, respectively. The effect of contact time of NH4+-N adsorption
on both LS and PPB reveals that the uptake of NH4+-N was rapid initially but slowed
with time gradually towards equilibrium (Figure 4.5).
Figure 4.5: The Effect of Contact Time and the Initial Concentration on
Adsorption of Ammonium Nitrogen from Human Urine on (a) Lateritic Soil, and
(b) Pineapple Peel Biochar
For LS, rapid uptake occurred between 0 – 60 minutes and then slowed gradually
between 60 – 90 minutes beyond which NH4+-N adsorbed was negligible. In the case of
66
PPB, rapid uptake occurred between 0 – 30 minutes and then slowed down gradually
between 30 – 60 minutes beyond which NH4+-N adsorbed was negligible. The rapid
uptake could be explained by the existence of high ionic gradient between the solution
and the active adsorption sites at the beginning of adsorption resulting into rapid mass
transfer of NH4+-N onto the adsorption sites. However, as the adsorption progressed, the
ionic gradient reduced thereby slowing the transfer of ions. Eventually, the adsorption
sites got saturated and could no longer adsorb more of the ions. Huang et al. ( 2010)
using natural Chinese zeolite as adsorbent for removal of NH4+-N from aqueous solution
observed a similar trend whereby the uptake of the NH4+-N ions was rapid within the
first 60 minutes and then slowed down considerably, such that beyond 120 minutes the
adsorption was almost negligible. These authors attributed this behaviour to the quick
utilization of the most readily available active adsorption sites of the zeolite resulting
into fast diffusion and attainment of equilibrium rapidly. Guo et al. (2019) in their study
on removal of cadmium ions (Cd2+) from aqueous solution using activated carbon also
observed an initial rapid uptake of the ions and slowing down of the adsorption towards
equilibrium. The authors attributed this phenomenon to the availability of large number
of empty adsorption sites at the initial stage with the remaining adsorption sites being
difficult to occupy as a result of the repulsive forces from the already occupied sites
nearby. According to Inyinbor et al. (2016), the initial rapid adsorption of solutes may be
attributed to the contacts of the solute molecules with the easily available surface of the
adsorption sites, while subsequent gradual adsorption attributed to the slow uptake of the
molecules into the pores of the adsorbents.
The increasing amount of NH4+-N adsorbed with increase in initial concentration alludes
to more ions being progressively available to attach themselves on the adsorption sites.
This finding corroborates those published in literature. For instance, Kizito et al. (2015)
observed that as the initial NH4+-N concentration in pure NH4Cl solution and manure
slurry increased from 250 – 1400 mg/L, the amount adsorbed at equilibrium by both
wood and rice husks biochar increased significantly. Similarly, Sarkhot et al. (2013)
using hardwood biochar for sorption of ammonium from dairy effluent and pure NH4Cl
67
solution observed that the equilibrium adsorbed amount of ammonium increased
significantly with increasing initial concentration from 0 – 1000 mg/L. Both Kizito et al.
(2015), and Sarkhot et al. (2013) ascribed their observation to the dependence of the
amount of solutes adsorbed on the mass flow of the ions to the active sites, which is in
turn was driven by the solute concentration. Thus, at low concentration, there is equally
low flow of ions resulting into many active sites on the adsorbent being unoccupied.
In all the conditions investigated (varying the initial concentrations and the contact
time), both LS and PPB adsorbed NH4+-N from urine. The capability of biochars to
adsorb NH4+-N has been attributed to the presence of acidic functional groups
containing oxygen and hydrogen, which makes the surface to be negatively charged
thereby electrostatically attracting the NH4+-N (Ahmad et al., 2014; Sumaraj & Padhye,
2017; Hu et al., 2020). Based on the FTIR results, LS contained the functional groups
having oxygen and hydrogen, which could have made its surface to be negatively charge
thereby attracting the NH4+-N. Also, studies on zeolites have shown that the presence of
aluminium and silicon elements makes zeolites to have high cation exchange capacity
and high NH4+-N selective properties (Kithome et al., 1998). The EDS, and XRD results
revealed that LS was dominantly composed of aluminium and silicon and this could
have contributed to its high adsorption capacity of 10.73 mg/g, which is relatively higher
than for most of the biochars reported.
68
capacity, the surface morphology of a material plays a significant role in enhancing its
adsorption capacity.
The graphical plots of Langmuir, Freundlich, and D-R isotherm models of NH4+-N
adsorption on LS and PPB are shown in Figure 4.6.
Figure 4.6: (a) Langmuir, (b) Freundlich, and (c) D-R Isotherm Model Fits of
Lateritic Soil and Pineapple Peel Biochar
69
The equilibrium adsorption data obtained for NH4+-N adsorption on LS was best fitted
by the D-R model (R2 = 0.999) compared to Langmuir (R2 = 0.977), and Freundlich
models (R2 = 0.928) as presented in Table 4.2. This implies that the adsorption was
mainly due to pore volume filling as opposed to multilayer adsorption on the pore walls
(Inglezakis, 2007; Alberti et al., 2012).
The mean surface adsorption energies calculated from the D-R model presented in Table
4.2 also indicate that NH4+-N was bound on the active adsorption sites of LS by weak
electrostatic (Van der Waal) forces since the value was below 8 kJ/mol (Hu & Zhang,
2019). This becomes ideal in agricultural production since the adsorbed ions can easily
dissolve in soil solution and readily be utilized by plants. Based on the XRD and SEM
analyses, LS exhibited a mixture of phases, suggesting the possibility that the active
adsorption sites could have varying adsorption energy potential or affinity, a property
which makes the D-R model the best in describing NH4+-N adsorption on LS. This is
because the D-R model does not solely consider homogeneity or constant adsorption
potential of the adsorption surface (Kaur et al., 2015). The D-R model therefore
displayed a more general applicability in comparison to Langmuir, and Freundlich
models, as it gave the best fit to the data. Although most studies have investigated the
70
use of zeolite as a geological material for adsorption of NH4+-N, there are no reports on
the adsorption capacity of NH4+-N on LS that could be used to compare the value of
10.73 mg/g obtained herein.
PPB also exhibited an adsorption trend similar to LS. The correlation coefficients (R2)
were 0.999, 0.977, and 0.928 for the D-R, Langmuir, and Freundlich models,
respectively. Therefore, the description of the behaviour of NH4+-N adsorption on PPB
is similar to that of LS. The PPB exhibited phase mixtures although it was dominantly
composed of carbon as revealed by XRD data. The phase impurities could have resulted
into its heterogeneity, a property that made D-R model best in describing NH4+-N
adsorption on its active sites. These finding agree well with the study by Halim et al.
(2017). These authors observed that the D-R model best described the behaviour of
NH4+-N adsorption on spent mushroom substrate biochar, and that the NH4+-N ions were
attracted onto the surface of the adsorbent by weak electrostatic forces. Zhu et al. (2012)
reported in their study of NH4+-N removal from aqueous solution using activated carbon
from rice husks that D-R model best described the adsorption process although ion
exchange was the major mechanism of adsorption. A similar study on NH4+-N
adsorption on biochar derived from different feedstocks (maize stover, sugarcane pith,
pine wood and grape pip) concluded that NH4+-N adsorption occurred on heterogeneous
surfaces through weak electrostatic attraction (Aghoghovwia et al. 2020). A comparison
of the maximum adsorption capacity of NH4+-N (13.40 mg/g) on PPB obtained herein
with those reported for biochars from different sources are presented in Table 4.3.
71
Table 4.3: Capacities of Various Biochars in Adsorbing Ammonium Nitrogen
The differences in adsorption capacities presented in Table 4.2 could have resulted from
different experimental conditions such as initial concentrations, biochar modification,
and different pyrolysis temperature as well as solution pH. However, most of the studies
indicate that biochar has a great potential for NH4+-N recovery.
4.1.4 The Effect of Contact Time and the Initial Concentration on Adsorption of
Phosphorous from Human Urine on Lateritic Soil and Pineapple Peel Biochar
The effect of contact time and the initial concentration of P in urine on the amount of
adsorbed P on LS and PPB are presented in Figure 4.7a and 4.7b, respectively.
72
Figure 4.7: The Effect of Contact Time and the Initial Concentration of
Phosphorous Adsorption on (a) Lateritic Soil and (b) Pineapple Peel Biochar
73
finally a steady state condition where negligible uptake of the solutes occurred. These
authors attributed the initial rapid mass transfer of the phosphates to the adsorption sites
to a high ionic gradient. Moreover, the slow rate of phosphate sorption after equilibrium
could have also resulted from an increase in the total negative charge on the LS surface
causing the repulsion of the anionic phosphates in the solution. Özacar (2003) reported
that the adsorption of P onto calcined alunite increased with time but remained constant
after 120 minutes. The author attributed the single, smooth, and continuous time profile
curve of P uptake to the possible monolayer adsorption on the surface of the adsorbent.
On the other hand, Figure 4.7b shows that PPB adsorbed P in human urine but released
it into the solution gradually thereby not reaching equilibrium. The phenomenon of non-
modified biochar leaching phosphorous into aqueous solution has been previously
reported. For instance, Yao et al. (2012) reported that non-modified sugarcane bagasse
carbonized at 300 and 600 °C adsorbed and released phosphorous into the aqueous
solution thereby not reaching equilibrium. Also, a study by Wunderlich (2015) observed
that non-modified biochar could not adsorb phosphorous at all from human urine but
instead leached the inherent P it into the solution. Nevertheless, the capacity of PPB to
74
adsorb P from human urine could be attributed to the presence of calcium oxides which
has been reported to have capacity to remove phosphates from human urine through
formation of amorphous calcium phosphates (Randall et al., 2016). However, Randall et
al. (2016) observed that calcium oxides are highly soluble in urine, which could have
resulted into the dissolution of the formed complexes as evidenced in this study making
equilibrium not to be attained. Attempt to increase the capability of the negatively
charged biochars to adsorb the anionic phosphates has resulted in a concept of
modification of biochar surfaces using metal cations with magnesium oxides being the
most widely reported compound for the modification (Li et al., 2017; Xu et al., 2019).
Yao et al. (2013) in their attempt to study the correlation between P removal rates and
metal element (Mg, Ca, K, Fe, Zn, Cu, and Al) contents of 25 biochars, reported a
varying correlation, with Mg having the highest correlation (77.9%) compared to Ca
(36.4%), K (25.1%), Cu (10.6%), Fe (8.5%), Al (1.3%) and Zn (1.1%). Other studies
(Liu et al., 2019; Luo et al., 2019) have similarly reported that the presence of positively
charged MgO nanocomposites in the biochar matrix is the major mechanism of P
removal from aqueous solutions through electrostatic attraction to form magnesium-
phosphate complexes (MgHPO4 and Mg (H2PO4)2) on the surface.
4.1.5 Adsorption Isotherms for the Adsorption of Phosphorous from Human Urine
on the Lateritic Soil
The adsorption isotherms of the Langmuir, Freundlich and D-R models describing the
behaviour of P adsorption on LS are presented in Figure 4.8, while the isotherm
parameters are summarized in Table 4.6. Lateritic soil was chosen for equilibrium
studies due to its potential to hold phosphorous unlike pineapple peel biochar that shows
a tendency to release phosphorous in urine solution.
75
Figure 4.8: (a) Langmuir, (b) Freundlich, and (c) D-R Isotherm Model Fits of
Lateritic Soil
The correlation coefficients (R2) for Langmuir, Freundlich, and D-R models were 0.984,
0.966, and 0.858, respectively as presented in Table 4.6. On the basis of the R2 values,
both Langmuir and Freundlich isotherm models yielded satisfactory fit to the
experimental data but Langmuir model gave the best fit. Therefore, it can be interpreted
that the available adsorption sites for P on LS have the same adsorption energy and that
the process is reversible. In essence, the sorption mechanism is controlled by the surface
area and chemical properties of the soil. Other studies (Bhattacharjee et al., 2021;
Fulazzaky et al., 2014) have also reported the Langmuir model to best describe P
adsorption on LS with Al and Fe cations considered to be responsible for the removal of
P from aqueous phase through electrostatic attraction to form complexes. The calculated
mean surface adsorption energy (E = 1.313 ×10-2 kJ/mol) from the D-R model presented
in Table 4.6 shows that P was bound on the LS surface by weak electrostatic (Van der
Waal) forces since the value was below 8 kJ/mol (Hu & Zhang, 2019).
76
Table 4.4: Langmuir, Freundlich, and D-R Isotherm Parameters for the
Adsorption of Phosphorous on Lateritic Soil
Isotherm parameters
Langmuir
qmax (mg/g) 45.25
KL 0.001
RL 0.6
R2 0.984
Freundlich
Kf 0.119
1/n 0.778
R2 0.966
Dubinin-Radushkevich (D-R)
qo (mg/g) 13.75
KDR 0.0029
E (kJ/mol) 1.313 × 10-2
R2 0.858
The adsorption capacities of different adsorbents are highly dependent on the operating
temperatures, material surface area, and initial solution concentrations (Liang Zhang et
al., 2011). As shown in Table 4.7, it is evident that LS used in this study can adsorb a
maximum of 45.25 mg/g P from human urine, which is in the typical range for many
adsorbents; 9.8 mg/g for bentonites, 4.3 mg/g for wood aspen fibres, 41.97 mg/g for
calcined alunite, and 36 mg/g for Fe-Mn binary oxide among others.
77
4.2 Optimised Adsorption of Ammonium Nitrogen and Phosphorous from Human
Urine on Pineapple Peel Biochar and Lateritic Soil
The response surface methodology (RSM) was carried out using “Design-Expert®
software, version 12, Stat-Ease, Inc., Minneapolis, MN, USA, www.statease.com” to
optimise the adsorption of NH4+-N on pineapple peel biochar (PPB) and to investigate
interactive effect of contact time, agitation speed and adsorbent loading. Based on the
results from the adsorption isotherm studies, pineapple peel biochar was selected for
optimization studies due to its high monolayer maximum adsorption capacity (22.727
mg/g) compared with lateritic soil (19.231 mg/g). The optimisation modeling was
performed at 21 °C, and pH of 7.2. The relationship between the response of adsorbed
NH4+-N and the three influencing variables; contact time, agitation speed, and adsorbent
loadings was analysed based on the central composite circumscribed design with each
variable studied at 5 levels.
The estimated second-order quadratic response model equation in coded form that was
used to explain the adsorption behavior of NH4+-N is presented in Equation 4.1 given as:
Where:
x1, x2, and x3 = the coded terms for contact time, agitation speed, and adsorbent loading,
respectively
The predicted amount of the adsorbed NH4+-N for each run presented in Table 4.4 was
obtained by solving Equation 4.1.
78
Table 4.6: Ammonium Nitrogen Adsorption Results Based on the Response Surface
Methodology and the Central Composite Circumscribed Experimental Design
As shown in Table 4.4, a wide range of adsorbed amount of NH4+-N on PPB from a
minimum of 7.62 mg/g to a maximum of 56.55 mg/g occurred for the predicted
responses, indicating a strong correlation between the adsorption of NH4+-N and the
contact time, speed and loading.
4.2.2 ANOVA Analysis and Reliability of the Response Surface Methodology Model
for Optimised Adsorption of Ammonium Nitrogen from Human Urine on
Pineapple Peel Biochar
An ANOVA was conducted to study the significance of the quadratic model, while the
value of the coefficient of determination (R2) was used to determine the reliability of the
quadratic model in fitting the experimental data and accurately predicting the output.
The quadratic model had R2 value of 0.999 implying that approximately 99.9% of the
variance in the experimental data is attributed to the variables investigated. The
quadratic model was therefore adequate in predicting the amount of adsorbed P under
79
the given experimental conditions since nearly all the variations could be accounted for
by the quadratic model. Moreover, there was no difference between the R2 (0.999) and
adjusted R2adj (0.999) suggesting that the predicted values (response) matched well with
the observed ones, implying that the regression model had an excellent stability (Myers
et al., 2004). To verify the adequacy of the model, the experiment was conducted at the
optimal conditions described of 60 minutes contact time, 80 rpm agitation speed, and
adsorbent loading of 0.1g, yielding an adsorption capacity of NH4+-N of 36.42 mg/g,
which agreed well with the predicted adsorption capacity of 36.84 mg/g. The ANOVA
results for the response surface quadratic model for the adsorption of P are summarized
and presented in Table 4.5.
Table 4.7: ANOVA for Response Surface Quadratic Model for Optimised
Adsorption of Ammonium Nitrogen from Human Urine on Pineapple Peel Biochar
The probability p-value of < 0.0001 indicates that there was less than 0.01% chance that
the output response (adsorbed NH4+-N) is due to noise and therefore the quadratic model
was statistically significant at 95% confidence level. Among the three factors, contact
time, agitation speed, and adsorbent loading optimised for achieving optimal adsorption
of NH4+-N on PPB, increasing the agitation speed did not significantly (P > 0.05)
increase the uptake of NH4+-N. This is further depicted by the 3D surfaces in Figures
80
D1a and D1b presented in Appendix IV which shows no gradual change in the amount
of NH4+-N with increase in agitation speed. Theoretically, it is expected that increasing
speed results into increased collision of active sites with the solutes thereby increasing
the amount of solutes adsorbed. However, from the findings, it can be stated that
increasing the agitation speed beyond the optimum level does not increase the amount of
NH4+-N adsorbed on PPB. This concurs with a study by Zahoor (2011) who observed
that the adsorption of insecticide (imidacloprid) on activated carbon increased with the
agitation speed from 200 upto 300 rpm beyond which negligible adsorption occurred.
The author argued that increasing turbulence decreases the boundary layer thickness
surrounding the adsorbent particles leading to an increase in the accessibility of the
available adsorption site. However, negligible increase in the adsorption beyond 300
rpm implies that for any adsorption process, there is an optimum speed to achieve
optimal adsorption, and thus should be investigated for every adsorbent.
On the other hand, increasing the contact time beyond 30 minutes significantly (P<0.05)
increased the amount of NH4+-N adsorbed on PPB, an observation which indicates that
equilibrium conditions had not been attained at 30 minutes since no further adsorption
occurs when all active adsorption sites are occupied. Since adsorption occurs on
adsorbent surfaces or pores, the removal of solutes in aqueous phase will generally be
affected by the adsorbent dosage. In this study, increasing the adsorbent dosage beyond
0.1g significantly (P<0.05) reduced the amount of NH4+-N adsorbed which is also
depicted by the 3D surfaces presented as Figures D1a and D1c in the Appendix IV. This
is in line with studies by Pillai et al. (2014) and Sica et al. (2014) who observed that
increasing the adsorbent loadings resulted into a decrease in the amount of solutes
adsorbed. These authors explained that increase in adsorbent loadings results in the
agglomeration of particles thereby reducing the amount of surface available for
attachment, which consequently results into low amounts solutes adsorbed at
equilibrium. Based on these observations, it is worth pointing out that to achieve an
optimal removal of solutes in any given fixed volume of a solution, the adsorption
81
experiments should be carried out at optimum contact time, agitation speed and
adsorbent dosage.
Data analysis using the Response Surface Methodology (RSM) was undertaken using
“Design-Expert® software, version 12, Stat-Ease, Inc., Minneapolis, MN,
USA, www.statease.com” to optimise the adsorption of P on LS and to investigate
interactive effect of contact time, agitation speed and adsorbent loading. The
optimisation modelling was performed at 21 °C, pH of 7.83 and initial P concentration
of 470.03 mg/L. The relationship between the response of adsorbed P and the three
influencing variables, contact time, agitation speed, and adsorbent loadings considered
in this study was analysed based on the central composite circumscribed design with
each variable studied at 5 levels.
The estimated second-order quadratic response model equation in coded form was used
to explain the adsorption behavior of P on LS is presented in Equation 4.2 given as:
Where:
x1, x2, and x3 = the coded terms for contact time, agitation speed, and adsorbent
loading, respectively.
The predicted amount of the adsorbed P for each run presented in Table 4.8 was
obtained by solving Equation 4.2.
82
Table 4.8: Phosphorous Adsorption Results Based on the Response Surface
Methodology and the Central Composite Circumscribed Experimental Design
4.2.4 ANOVA Analysis and Reliability of the Response Surface Methodology Model
for Optimised Adsorption of Phosphorous from Human Urine on Lateritic Soil
An ANOVA was conducted to study the significance of the quadratic model, while the
value of the coefficient of determination (R2) was used to determine the reliability of the
quadratic model in fitting the experimental data and accurately predicting the output.
The quadratic model had R2 value of 0.972 implying that approximately 97.2% of the
variance in the experimental data is attributed to the variables investigated. The
quadratic model was therefore adequate in predicting the amount of adsorbed P under
the given experimental conditions since only 2.84 % of the total variations could not be
83
accounted for by the quadratic model. Moreover, the difference between the R 2 (0.972)
and adjusted R2adj (0.956) was < 0.2 suggesting that the predicted values (response)
matched well with the observed ones, implying that the regression model had an
excellent stability (Myers et al., 2004a). To verify the adequacy of the model, the
experiment was conducted at the optimal conditions described, yielding an adsorption
capacity of P of 112.23 mg/g, which agreed well with the predicted adsorption capacity
of P. The ANOVA results for the response surface quadratic model for the adsorption of
P are summarized and presented in Table 4.9.
Table 4.9: ANOVA for Response Surface Quadratic Model for Optimised
Adsorption of Phosphorous from Human Urine on Lateritic Soil
The results show that the quadratic model is statistically significant at 95% confidence
level, with F-value of 37.97 and very low p-value of < 0.0001, i.e., there is less than
0.01% chance that the output response is due to noise. Among the three factors, contact
time, agitation speed, and adsorbent loading optimised for achieving optimal adsorption
of P on LS, increasing the contact time did not significantly (P > 0.05) increase the
uptake of P. This implies that once the equilibrium is reached, no further adsorption can
occur since all the active sites are occupied.
84
On the other hand, increasing the agitation speed beyond 80 rpm showed a gradual
decline in the amount of the adsorbed P as depicted in the 3D graphs presented in
Figures E1a, and E1c in the appendix. We can draw parallels with a study by
Geethakarthi and Phanikumar (2011) where the adsorption of reactive dyes on activated
carbon was seen to decline with increase in the agitation speed beyond 120 rpm. The
desorption or detachment of the adsorbed molecules at higher speed suggest that there is
an optimal agitation speed or turbulence for any adsorption process. However,
increasing the agitation speed gradually upto the optimum generally leads to a gradual
increase in the amount of solutes adsorbed. For instance, Zahoor (2011) observed that
the adsorption of imidacloprid on activated carbon increased with the agitation speed
from 200 upto 300 rpm beyond which no significant increase in the adsorption occurred.
The author argued that increasing turbulence decreases the boundary layer thickness
surrounding the adsorbent particles leading to an increase in the accessibility of the
available adsorption site. However, insignificant increase in the adsorption beyond 300
rpm implies that for any adsorption process, there is an optimum speed to achieve
optimal adsorption, and thus should be investigated for every adsorbent.
The removal of solutes in aqueous phase will generally be affected by the adsorbent
dosage since the adsorbents provide attachment sites. In this study, adsorbent dosage
beyond 0.1g led to a decline in the amount of the adsorbed P as shown by the 3D graphs
in Figure E1a and E1b in the appendix V. This is line with the work by Drenkova-
Tuhtan et al. (2017), which reported a decrease in the removal of phosphates with
increase in the adsorbent dose beyond the optimum dosage. The authors credited this
observation to agglomeration of particles at higher dosage thereby reducing the
accessibility of the active adsorption sites. In contrast, Xiong et al. (2011) reported that
phosphate removal from aqueous solution increased with the amount of powdered
freshwater mussel shells, an observation they attributed to an increase in the available
surface area for attachment of phosphates. Based on the above discussion, it is worth
pointing out that to achieve an optimal removal of solutes in any given fixed volume of a
85
solution, the adsorption experiments should be carried out at optimum contact time,
agitation speed and adsorbent dosage.
The comparative results of proximate analysis (volatile matter, fixed carbon, moisture
content, and ash content), ultimate analysis (C, H, N, and O) and heating values of faecal
char and sawdust char are listed in Table 4.10. The volatile matter in faecal char
compares well to sawdust char hence good for combustion processes, although the latter
biomass has noticeably more ash content, ~8 %, which may affect its burning rate. This
value is however below those reported in the literature, i.e., 13% to 51.2% (Hafford et
al., 2018; Krueger et al., 2020). High ash content may pose several issues such as poor
combustion, excess processing costs, disposal problem, and reduced energy conversion
(Sait et al., 2012). The predicted gross calorific values for faecal char and sawdust char
were 24.5 MJ/kg, and 26.3 MJ/kg, respectively, which are comparable to those of
natural coals (8 – 27 MJ/kg), wood (19 – 25 MJ/kg), and charcoal (29 – 33 MJ/kg) (Patel
et al., 2007; Ruiz-Aquino et al., 2019). The carbon content in faecal char, 55.3%, is
comparable to 58.3% and 56.1% in human faeces pyrolyzed at 200 and 300 °C,
respectively (Afolabi et al., 2017; Ward et al., 2014). Also, the 78.2% carbon content in
sawdust char is in complete agreement with 78.7% in hazelnut shells pyrolyzed at 350
°C (Bonelli et al., 2003).
86
Table 4.10: Proximate and Ultimate Analysis of Faecal Char and Sawdust Char
a
Estimated as a difference
The morphological SEM image shown in Figure 4.9a reveals faecal char surface as
fibrous with very few pores, which may be due to dehydration and decomposition of
volatile matter (Liang et al., 2016). On the other hand, sawdust exhibits highly
macroporous tubular shapes, attributable to the presence of plant cells (Figure 4.9b). The
tubes are perforated, an indication that they were partially oxidised (Krueger et al.,
2020). These observations correlate fairly well with previous reports on faecal char, and
sawdust char (Chowdhury et al., 2016; Krueger et al., 2020).
87
Figure 4.9: Physical Properties of Faecal Char, Sawdust Char, and Blend; (a) and
(b) SEM Images of Faecal Char and Sawdust Char, Respectively, (c) FTIR Spectra,
and (d) TGA(solid lines) and DTG (dash lines) Profiles at a Heating Rate of 5
°C/min from 30 to 800 °C Under Air
The main characteristic FTIR functional groups of faecal char, and sawdust char are
shown in Figure 4.9c. There are some similarities between the spectra of these samples,
and are mostly composed of hemicellulose, cellulose and lignin albeit with minor
differences. Bands commonly associated with hemicellulose and cellulose are observed
in the region between 3400 and 3000 cm-1 corresponding to –OH stretching due to
moisture, while the peaks between 2922 and 2851 cm-1 may be assigned to aliphatic C–
H stretching (Abdulrazzaq et al., 2014). The bands around 1700 cm-1 are typical of C=O
stretching vibrations of unconjugated hemicellulose (de Figueredo et al., 2017), while
the sharp band around 2981 cm-1 is due to the asymmetric stretching of CH2 and CH of
cellulose (Lun et al., 2017). The absorption bands in the region between 1600 and 1540
88
cm-1 are due to C=C skeletal stretching of the aromatic compounds (Keiluweit et al.,
2010). The bands around 1026 cm-1 are associated with symmetric C–O stretching of
cellulose, hemicellulose, and lignin (Liu et al., 2015). The bands in the region 850 – 755
cm-1 are credited to the out of plane deformation modes of aromatic C–H (Y. Liu et al.,
2015). Some differences in the two samples are also noticeable. Specifically, the relative
intensities of bands around 1600 and 1540 cm-1 are greater for sawdust char than for
faecal char, attributable to high carbon content in the aromatic rings (Bonelli et al.,
2003). A sharp C–O stretching band, and of high intensity is seen for faecal char around
1034 cm-1, and is typical of carbohydrates or hemicellulose and cellulose (Afolabi et al.,
2017). These results share a number of similarities with those observed for biochars
derived from human faeces, grass, rice husk, and wood (Keiluweit et al., 2010;
Abdulrazzaq et al., 2014; Afolabi et al., 2017).
89
and blend is lower, and is attributed to low volatile content and high ash content, in line
with the proximate analysis. A slight weight loss due to moisture release is observed
before reaching 100 °C for sawdust char, but the main thermal decomposition results in
one major peak around 410 °C, corresponding to the simultaneous decomposition of
hemicellulose and cellulose, whereas the existence of multi-component steps seen for
faecal char and blend between 250 and 550 °C is credited to the co-operative release of
hemicellulose, cellulose, and lignin. Raveendran et al. (1996) demonstrated that small
amounts of volatiles from lignin component may also be given off in the second stage. It
is here further observed that, the decomposition patterns of each sample involved multi-
step reaction. This phenomenon can be explained by the complex nature of the samples
since degradation patterns indicate overlapping zones of hemicellulose, cellulose, and
lignin. These results corroborate FTIR data and are also concurrent with degradation of
other biomasses found in the literature (Grønli et al., 2002; Shaaban et al., 2013;
Abdulrazzaq et al., 2014).
Comparison of the TGA curves under air and nitrogen as the purge gases show
noticeable difference between the two gases for each sample (Figure 4.10).
90
Figure 4.10: Thermal Behavior at a Heating Rate of 5 °C/min under Flowing N2
and Air from 30 to 800 °C
When the samples are heated under air, a higher weight loss is seen than when heated
under nitrogen. This could be due to chemical reactions occurring between oxygen and
the volatiles present in the biomass. The reaction would result in more volatiles being
liberated and therefore a larger amount of mass loss when using air as the purge gas.
This lends support to previous works by Ma et al. (2018), and Chandrasekaran & Hopke
(2012) on the thermal degradation of palm oil waste, and switch grass, respectively.
It is apparent that the thermal degradation process is dependent on the heating rate under
air, i.e., when the heating rates were increased from 5 °C/min to 40 °C/min, the thermal
decomposition process was slower (Figure 4.11). This is consistent with the work by
(Heydari et al., 2015) on the thermal decomposition of lignite coal, and further supports
the idea that a higher heating rate has a shorter decomposition time and the temperature
needed for the sample to reach the same conversion will be higher. Thus, TGA and DTG
curves are shifted towards higher temperature. Also, some individual small DTG peaks
91
at lower heating rates overlapped to form broader peaks at higher heating rates thereby
degrading simultaneously. As discussed by Gai et al. (2013), a short reaction time is
required at higher heating rates, therefore the temperature needed for the sample to
decompose is also higher. Other researchers have attributed this phenomenon to less heat
transfer among sample particle with increasing heating rate due to minimal time
available for volatilization to occur (Chen et al., 2017; Zhang et al., 2019). Sittisun et
al., (2015) noted that the reduced residence time at higher heating rate may be
insufficient for heat transfer to permeate into the center of the tested sample thereby
delaying the thermal decomposition process.
Figure 4.11: TGA (solid lines) and DTG (dash lines) Profiles of (a) Faecal Char, (b)
Sawdust Char, and (c) Blend under Air with the Ramp Rates of 5, 10, 20, and 40
°C/min
92
4.3.4 Conversion Rate
The conversion curves shown in Figure 4.12 exhibit a theoretical temperature hysteresis,
which is manifested as a curve shifting to a higher temperature region with increase in
the heating rate without changing the thermal profile. This is an indication that during a
thermal decomposition process, the overall reaction rate is only dependent on
temperature (Lopez-Velazquez et al., 2013). Hence, combustion characteristic
parameters shifted to higher values with elevated heating rates.
Figure 4.12: Conversion Curve of the Entire Reaction (a) Faecal Char, (b) Sawdust
Char, and (c) Blend at Different Heating Rates (β)
93
4.3.5 Decomposition Kinetics
A number of successive and/or parallel reactions occur during the thermal breakdown of
biomass making it a complicated process. Two iso-conversional methods, KAS, and
FWO were employed to determine the decomposition kinetics during active
decomposition step in the range 0.2 ≤ α ≤ 0.9 at four different heating rates (β), 5, 10,
20, and 40 °C/min (Figure 4.13).
Figure 4.13: KAS (a-c) and FWO (d-f) Plots of Faecal Char, Sawdust Char, and
Blend at Different Values of Conversion (α).
The activation energies (Eα) were extracted from the slopes of iso-conversional lines, -
(Eα/R) in the KAS, and -1.052(Eα/R) in the FWO methods. The calculated Eα together
with regression coefficients (R2) for each α are summarized in Tables 4.15 and 4.16. The
94
two models agreed well with the experimental data over 0.2 ≤ α ≤ 0.9, for faecal char,
and blend. However, for sawdust char the models could only fit 76.3% of the data at α =
0.2, and over 91% of the data at α > 0.2. Sawdust char contained a significant amount of
physisorbed moisture (~8 wt. %) as seen from the TGA results. Therefore, the energy
absorbed at α = 0.2 was partly used for water release, whereas only decomposition of
volatile matter occurred at α > 0.2. Similarly, Chandrasekaran and Hopke (2012)
observed in their study of kinetics of thermal decomposition of switch grass pellets, that
iso-conversional model gave a good fit to the experimental data between a conversion
range of 0 – 70%, beyond which the model deviated significantly from the experimental
data. These authors, however, utilized a single step global reaction model to determine
the kinetics rather than a multi-step reaction model.
Table 4.11: Activation Energies (Eα (kJ/mol)) Estimated for Faecal Char, Sawdust
Char, and Blend using KAS Method at Different α
95
Table 4.12: Activation Energies (Eα (kJ/mol)) Estimated for Faecal Char, Sawdust
Char and Blend using FWO Method at Different α
The trend of the lines presented in Figure 4.13, could be divided into two stages: α = 0.2
– 0.6, and α = 0.7 – 0.9 based on their parallelism. Thus, a two-step kinetic mechanism
is applied to describe the thermal degradation of all the samples. Activation energy is
considered as the minimum energy required to start a reaction, therefore higher
activation energy values means that a reaction is more difficult to start. Between α = 0.2
– 0.6, the average activation energies, using KAS method were 114.4, 119.8 and 112.7
kJ/mol for faecal char, sawdust char, and blend, respectively. These values represent
activation energies associated with the decomposition of hemicellulose and cellulose. On
the other hand, the average activation energies during lignin decomposition, i.e.,
between α = 0.7 – 0.9, were 85.9, 90.4 and 91.7 kJ/mol for faecal char, sawdust char and
the blend, respectively. Similar values were also found for average activation energies
estimated using FWO method at the same α (Table 4.16). Such that, the average
activation energies of faecal char were 120.2 kJ/mol between α = 0.2 – 0.6, whereas
sawdust char, and blend had average activation energies of 124.9 and 118.4 kJ/mol,
respectively. The activation energies between α = 0.7 – 0.9, were 94.3, 97.7 and 99.4
kJ/mol for faecal char, sawdust char, and blend, respectively. Evidently, the activation
energy of the samples decreased in the second reaction step (α = 0.7 – 0.9). This
remarkable difference can be rationalized by assuming that thermal decomposition starts
with slower reactions and is independent of the heating rates. Overall, the two iso-
96
conversional methods yielded similar average activation energies and regression
coefficients.
Independent of the method used, and for all the samples, the average Eα values were
determined to be in the range of 105 – 135 kJ/mol, and 80 –105 kJ/mol for stages I (-
α = 0.2 – 0.6), and II (α = 0.7 – 0.9), respectively. These values align well with those in
the literature for similar biomass materials. For instance, Grønli et al. (2002) contended
that for both hard and soft woods, the E values ranged between 98 – 256 kJ/mol for
hemicellulose and cellulose, and 46 kJ/mol for lignin. In their analysis of sewage sludge,
and animal manure, Sanchez et al. (Sanchez et al., 2009) reported activation energy
values of 146.4, and 143.3 kJ/mol, respectively. Yacob et al. (2018) concluded that E
values for hemicellulose and cellulose components in dried human faeces ranged
between 80 – 238 kJ/mol, while Urban and Antal (1982) identified average E values
between 130 – 250 kJ/mol for the two components in undigested sludge. These authors,
however, argued that the activation energies of human faeces are likely to be higher due
to the presence of micro-organisms. Herein, human faeces were pyrolyzed to eliminate
the effect of micro-organisms on the activation energies.
On the basis of the TGA data, the thermal degradation presented a multi-stage intricate
reaction phenomenon in which a series of simultaneous complex reactions occurred. The
decomposition of hemicellulose, cellulose, and lignin components ensued over a wide
temperature range, 200 – 550 °C at 5 °C/min, with corresponding prominent FTIR bands
in the range between 3400 – 755 cm-1. The thermal reactivity of the samples was
influenced by the heating rate. When two iso-conversional methods were used for
kinetic analysis, an alteration in the activation energy with conversion degree was
noticed for all the samples. Concerning the determination of activation energies, the
degradation regions were split into two stages. The average activation energy for stage-I
during active decomposition was found to be in the range of 105 – 135 kJ/mol, while
stage-II had average activation energy between 80 – 105 kJ/mol from both KAS and
FWO models. We accentuate that, the activation energy values varied greatly with α,
particularly in stage-I (α = 0.2 – 0.6) than in stage-II (α = 0.7 – 0.9). This variation is a
97
signature of heterogeneous nature of solid wastes, which leads to a complex mechanism
involving parallel, competitive, and complex reaction schemes. The higher Eα value in
the region 0.2 ≤ α ≤ 0.6 means that on average more energy is required in the
combustion process of hemicellulose and cellulose than for lignin.
4.4.1 Proximate Analysis Results and Heating Values of Briquettes and Charcoal
The major parameters; volatile matter (VM) , fixed carbon (FC), ash content (ASH),
moisture content (MC), and gross calorific value (GCV) used to assess the quality of
solid fuels depicts that charcoal made from acacia tree can liberate more heat, and
produce less ash making it a better fuel compared to the briquettes produced in this
study (Table 4.13).
98
Table 4.13: Proximate Analysis Results and Heating Values of Briquettes and
Charcoal
The higher gross calorific value of charcoal could be due to its higher fixed carbon
content (Bulmau et al., 2010). Moreover, biomass fuels with low ash content have higher
calorific values since the amount of heat absorbed by the inorganic fraction in ash is
reduced (Hafford et al., 2018). Other studies (Carnaje et al., 2018; Saeed et al., 2021)
have reported that solid fuels with lower moisture content have larger voids/pores which
allow easy diffusion of oxygen from air to sustain better combustion in solid fuels
consequently releasing more heat. Nevertheless, despite the briquettes having lower
fixed carbon content, high moisture content, and higher ash content which lowers the
calorific value of solid fuels, the reported value of 19.8 MJ/kg is comparable to 19–25
MJ/kg of wood fuel (Bulmau et al., 2010) and hence it can still provide net benefit
during combustion.
Solid fuels of higher volatile matter content ignites faster (Kabok et al., 2018) and
therefore, they are desirable. However, higher moisture reduces the speed of ignition of
briquettes thereby necessitating pre-drying before combustion or co-combustion of
briquettes with charcoal particles or wooden sticks so as to initiate the combustion
process (Ngusale et al., 2014). Pre-drying may however not be practical during rainy
seasons. The higher moisture content in the briquettes produced in this study could have
been contributed by molasses due to its hygroscopic nature as a result of the fructose
content (Davis, 1995; Palmonari et al., 2020).
99
4.4.2 Elemental Composition of Human Faecal Char, Sawdust Char, and Molasses
Table 4.14 presents the elemental composition of the main materials (faecal char and
sawdust char) used in the briquette production. Properties of cane molasses reported in
previous studies are also presented.
Table 4.14: Elemental Composition of Human Faecal Char, Sawdust Char, and
Molasses
Pyrolysis of biomass materials increases their carbon content (Li et al., 2017) which
consequently increases their calorific values (Bulmau et al., 2010). Faecal char had
lower carbon content than sawdust char and therefore blending the two materials could
produce a solid fuel with higher calorific value than faecal char. The nitrogen content in
faecal char was also observed to be higher compared to sawdust char, a property that can
be attributed to the presence of proteins in faeces that constitute the larger fraction of
nitrogenous compounds in faecal mass (Rose et al., 2015). Nitrogen content in human
faeces can however be significantly reduced by increasing pyrolysis temperature which
consequently results in the degradation of proteins (Spitzer et al., 2018). It was also
noted that sulphur content in faecal char and sawdust char are negligible which makes
them safe in terms of sulphur emission during combustion. Although the study did not
characterize molasses, other researchers (Zhai et al., 2018; Dirbeba et al., 2021) have
reported that molasses contains sulphur and therefore it could be a contributor of sulphur
containing gases during combustion of briquettes produced in this study.
100
4.4.3 Concentration of Flue Gases from Combustion of Charcoal and Co-
combustion of Charcoal with Faecal Char-Sawdust Char Briquettes
The profile of flue gases emitted against combustion time presented in Figure 4.14,
shows a rise in amount of flue gases liberate upto a peak followed by a gradual decline.
The sudden rise in the amount of gases after the declining phase in both set-ups is as a
result of the shaking of the cook stove so as to shed off the incombustible ash fraction
that accumulated on the surface of the solid fuel during combustion, consequently
allowing more heat and gases to be liberated. It was also observed that there was no
emission of NO during combustion of charcoal. In addition, whereas H2S emission
profile showed short lived peaks during charcoal combustion, there was continuous
liberation of the gas throughout the co-combustion of charcoal with briquettes.
Figure 4.14: (a) Emissions from Co-combustion of Charcoal with Faecal Char-
Sawdust Char Briquettes, and (b) Emissions from Combustion of Charcoal
CO2 results from complete combustion of fuel and therefore it is not regarded as a
pollutant while CO which is a product of incomplete combustion, is toxic to human
health (EPA, 2002; Townsend & Maynard, 2002). The amount of CO and CO2 released
during combustion is influenced by a range of factors. For instance, Kim et al. (2021)
in their attempt to compare emission from different briquettes attributed the lower CO
and CO2 emission from combustion of briquettes densified from spent coffee grounds to
their less carbon content (46.1–54.9 wt. %) compared to the anthracite briquettes (75.23
101
wt. %). High moisture content in biomass fuel also reduces oxidation reactions during
combustion hence significantly increasing CO emission while CO2 emission slightly
declines due to the reduced dry biomass per kg of the fuel (Bhattacharya et al., 2002).
Thus, the higher moisture content of the briquettes could have contributed to the higher
CO emission during co-combustion with charcoal. Both combustion of the charcoal and
co-combustion of charcoal with briquettes caused CO concentrations above the critical
limit of 35 ppm allowed for human exposure for one hour (EPA, 2002). It can be
recommended that proper ventilation is necessary when the fuels are used for indoor
heating to enhance oxidation of the carbon content in fuel.
Liberation of H2S during fuel combustion has been reported to be as a result of reduction
of SO2 that is formed due to oxidation of compounds containing sulphur in the fuel. For
instance, Shirai et al. (2013) observed that during coal combustion in a furnace, SO2
released reacts with H2 to form H2S and H2O as shown by the reaction Equation 4.3
given as:
SO2 3H 2 H 2 S 2 H 2 O (4.3)
Study by Ryason & Harkins (1967), also reported that the SO2 in flue gases released
during coal combustion in a furnace reacts with carbon monoxide to form CO2 and
elemental sulphur as depicted in Equation 4.4
1
2CO SO2 2CO2 S (4.4)
2
Based on WHO short term (30 minutes), H2S odour guidelines (WHO, 1981), it can be
stated that combustion of faecal char-sawdust char briquettes could cause more
discomfort throughout the combustion duration by liberating offensive odours since H2S
concentration was > 0.005 ppm. Moreover, the briquettes could cause other health
effects such as coughs, throat irritation and low oxygen uptake as had been reported by
102
Bhambhani & Singh (1991) since the emission was above 2.5 ppm. Apart from health
effects, H2S poses a challenge of corrosion of metal wares (Malone Rubright et al.,
2017).
Nitric oxides (NO) are mainly formed during combustion of solid fuels as a result of
oxidation of the organic nitrogen (fuel-N) that they contain (Aho et al., 1995;
Zevenhoven & Kilpinen, 2001). Prompt NO and thermal NO may also form from
atmospheric molecular nitrogen (N2) during combustion. However according to
Hayhurst & Lawrence (1996), formation of prompt NO which normally result from
reaction between N2 and hydrogen radicals from the fuel is small. The formation of
thermal NO on the other hand occurs at very high temperatures (≥1300 °C), since
thermal dissociation of N2 to N radicals is needed to start NO formation (Saastamoinen
& Leino, 2019). The highest temperature (475 °C) attained during the co-combustion of
briquettes with charcoal in this study could not result in the release of thermal NO. The
briquettes therefore released NO majorly as a result of oxidation of the organic nitrogen
contained in the faecal char. According to NIOSH exposure limits, it can be reported
that no immediate health hazard due to NO can be posed by combusting the briquettes
produced in this study since the peak emission of 30 ppm recorded was < 100 ppm
(Schieb, 1976). It is also worth noting that although oxygen was measured in the flue
gases, oxidation of NO to NO2 gas was not observed. Formation of NO2 requires
reaction between hydrogen peroxide radical (HO2) and NO which can only occur when
in the cooler zones of the flame or when rapid cooling of the flame occurs (Aho et al.,
1995; Glarborg et al., 2018; Zevenhoven & Kilpinen, 2001). Furthermore, NO2 rapidly
decomposes back to NO if it moves to the hot parts of the flame (Aho et al., 1995).
Cooling process does not exist in a furnace or in the experimental set-up adopted in this
study and thus possibly the reason as to why NO2 wasn’t detected. NO2 is therefore
mostly formed under normal ambient conditions in the atmosphere (Saastamoinen &
Leino, 2019)
A summary of the comparison of flue gases quality during combustion of charcoal and
in co-combustion of charcoal with briquettes against the permissible short term exposure
103
limit guidelines set by the Environmental Protection Agency (EPA), World Health
Organization (WHO), and National Institute for Occupational Safety and Health
(NIOSH) is presented in Table 4.15.
104
Figure 4.15: Variation of Flue Gases Temperatures with Combustion Time
It can be seen from Figure 4.15 that after ignition, a sudden rise in temperature of flue
gases occurs for both combustion set ups followed by a sharp decline in temperature and
finally a gradual temperature loss as combustion proceeds towards the end. Al-
Shemmeri et al.( 2015) made similar observation in a study on combustion of various
biomass fuels in a small-scale biomass combustor. The authors attributed the sudden rise
in temperature of flue gases upto a peak to the combustion of volatile gases, which is
immediately followed by a reduction in the calorific content of the fuel resulting to the
sharp decline in the flue gases temperature since they are carriers of the significant
proportion of the heat liberated during combustion. The authors further explained that
the gradual decline in temperature of flue gases after the combustion of volatile gases as
the fuel diminishes is associated with oxidation of char which is a slow process
105
compared to rate of combustion of volatile gases, resulting in longer periods of
combustion before the formation of ash. Other studies (Borowski et al., 2017; Pilusa et
al., 2013) with similar thermal profile obtained during combustion of briquettes ascribed
the variations in the heat content of the flue gases during combustion of the fuels to
heterogeneous complex reactions occurring during combustion of biomass materials. It
was also evident that the peak temperature (475 °C) of flue gases liberated during co-
combustion of briquettes with charcoal was higher compared to 222 °C attained during
combustion of charcoal. This is an indication that synergistic complex thermochemical
interactions occur during biomass co-combustion. Statistical test of the means of the
temperatures of the flue gases liberated using notched box plot (Figure 4.16a) showed
that there existed significant (P<0.05) differences during combustion in both set-ups
which depicts that co-combustion of briquettes with charcoal enhanced the amount of
heat released.
106
Figure 4.16: Notched Box Plots of (a) Flue Gases Temperatures During Co-
combustion of Briquettes with Charcoal and Combustion of Charcoal, (b) Oxygen
Concentrations During Co-combustion of Briquettes with Charcoal and
Combustion of Charcoal.
107
Figure 4.17: Variation of Flue Gases Temperatures with Oxygen Concentrations
From the Figure 4.17, a decline in O2 concentration in the flue gases occurred with
increase in temperature. Oxidation of hydrocarbons to form gases is an exothermic
process (Ren et al., 2019) therefore, higher oxygen consumption resulted in higher heat
energy content liberated and which are consequently carried by the flue gases in the
chimney. Similar observation was reported from combustion of eco-fuel briquettes
(Pilusa et al., 2013). An increase in O2 concentration in flue gases is expected as the fuel
combustion nears completion since most hydrocarbons will have been oxidised.
Statistically, the means of the oxygen concentrations in the flue gases at different
temperatures in both set-ups were not significantly (P>0.05) different as depicted by the
notched box plot (Figure 4.16b). This suggests that co-combustion of briquettes with
charcoal is beneficial in the provision of higher heat energy than combustion of charcoal
without consuming significantly higher amounts of oxygen.
108
CHAPTER FIVE
5.1 Conclusions
Based on the results and findings presented in this thesis, the following conclusions
can be made:
109
dangerous exposure limit of 100 ppm, while it can cause odor discomfort
since the hydrogen sulphide liberated was above exposure limit of 0.005
ppm. Also, the carbon monoxide released was above the exposure limit of 35
ppm an indication that combustion of the briquettes should be done in a well
ventilated room.
v. Emission of nitric oxides (NO) during combustion of briquettes densified
from human faecal char, sawdust char, and molasses is below the dangerous
exposure limit of 100 ppm set by the National Institute for Occupational
Safety and Health, while the carbon monoxide (CO), and hydrogen sulphide
(H2S) is above short term exposure limits of 35 ppm, and 0.005 ppm,
respectively, as set by Environmental Protection Agency, and World Health
Organisation, respectively.
5.2 Recommendations
Based on the results presented in this thesis, the following recommendations can be
made:
i. Pineapple peel biochar and lateritic soils should be enriched with ammonium
nitrogen and phosphorous from human urine to produce slow release fertilizers
or a medium for crop production so as to manage the pineapple peel wastes in the
environment and also to utilize lateritc soils for agricultural production.
ii. Pyrolyzed human faeces should be blended with pyrolyzed sawdust to lower the
input energy require to initiate its thermal decomposition and also lower its ash
content.
iii. Utilization of binders with no sulphur content is recommended for production of
briquettes from human faecal char and sawdust char so as to avoid emission of
sulphur containing gases during combustion.
110
iv. Combustion of solid fuels for domestic cooking and heating should be conducted
in a well-ventilated room to reduce the amount of carbon monoxide released by
enhancing complete oxidation of carbon content.
111
REFERENCES
Abdulrazzaq, H., Jol, H., Husni, A., & Abu-Bakr, R. (2014). Characterization and
Stabilisation of Biochars Obtained from Empty Fruit Bunch, Wood, and Rice Husk.
BioResources, 9(2), 2888–2898. https://doi.org/10.15376/biores.9.2.2888-2898
Afolabi, O. O. D., Sohail, M., & Thomas, C. L. P. (2017). Characterization of solid fuel
chars recovered from microwave hydrothermal carbonization of human biowaste.
Energy, 134, 74–89. https://doi.org/10.1016/j.energy.2017.06.010
Aguko Kabok, P., Nyaanga, D. M., Mbugua, J. M., & Eppinga, R. (2018). Effect of
Shapes, Binders and Densities of Faecal Matter - Sawdust Briquettes on Ignition
and Burning Times. Journal of Petroleum & Environmental Biotechnology, 09(02).
https://doi.org/10.4172/2157-7463.1000370
Ahmad, M., Rajapaksha, A. U., Lim, J. E., Zhang, M., Bolan, N., Mohan, D., Vithanage,
M., Lee, S. S., & Ok, Y. S. (2014). Biochar as a sorbent for contaminant
management in soil and water: A review. Chemosphere, 99, 19–33.
https://doi.org/10.1016/j.chemosphere.2013.10.071
Aho, M. J., Paakkinen, K. M., Pirkonen, P. M., Kilpinen, P., & Hupa, M. (1995). The
effects of pressure, oxygen partial pressure, and temperature on the formation of
N2O, NO, and NO2 from pulverized coal. Combustion and Flame, 102(3), 387–
400. https://doi.org/10.1016/0010-2180(95)00019-3
112
(Science Technology), 16, 22–31.
Al-Shemmeri, T. T., Yedla, R., & Wardle, D. (2015). Thermal characteristics of various
biomass fuels in a small-scale biomass combustor. Applied Thermal Engineering,
85, 243–251. https://doi.org/10.1016/j.applthermaleng.2015.03.055
Alberti, G., Amendola, V., Pesavento, M., & Biesuz, R. (2012). Beyond the synthesis of
novel solid phases: Review on modelling of sorption phenomena. Coordination
Chemistry Reviews, 256(1–2), 28–45. https://doi.org/10.1016/j.ccr.2011.08.022
Alhashimi, H. A., & Aktas, C. B. (2017). Life cycle environmental and economic
performance of biochar compared with activated carbon: A meta-analysis.
Resources, Conservation and Recycling, 118, 13–26.
https://doi.org/10.1016/j.resconrec.2016.11.016
Ali, I., Asim, M., & Khan, T. A. (2012). Low cost adsorbents for the removal of organic
pollutants from wastewater. Journal of Environmental Management, 113, 170–183.
https://doi.org/10.1016/j.jenvman.2012.08.028
Altun, N. E., Hicyilmaz, C., & Kök, M. V. (2001). Effect of different binders on the
combustion properties of lignite - Part I. Effect on thermal properties. Journal of
Thermal Analysis and Calorimetry, 65(3), 787–795.
https://doi.org/10.1023/A:1011915829632
American Society for Testing and Materials. (1984). Standard test method for chemical
analysis of wood charcoal - designation: D1762 - 84 (Reapproved 2007). Annual
Book of ASTM Standards, 84(2), 292–293. https://doi.org/10.1520/D1762-84R07.2
113
Antonini, S., Nguyen, P. T., Arnold, U., Eichert, T., & Clemens, J. (2012). Solar thermal
evaporation of human urine for nitrogen and phosphorus recovery in Vietnam.
Science of the Total Environment, 414, 592–599.
Aransiola, E. F., Oyewusi, T. F., Osunbitan, J. A., & Ogunjimi, L. A. O. (2019). Effect
of binder type, binder concentration and compacting pressure on some physical
properties of carbonized corncob briquette. Energy Reports, 5, 909–918.
https://doi.org/10.1016/j.egyr.2019.07.011
Asamoah, B., Nikiema, J., Gebrezgabher, S., Odonkor, E., & Njenga, M. (2016).
RESOURCE RECOVERY & REUSE SERIES 7 Fuel Briquettes. International Water
Management Institute.
Atwijukye, O., Kulabako, R., Niwagaba, C., & Sugden, S. (2018). Low cost faecal
sludge dewatering and carbonisation for production of fuel briquettes. 41st WEDC
International Conference, Srep 2014, 1–7.
Ayawei, N., Ebelegi, A. N., & Wankasi, D. (2017). Modelling and Interpretation of
Adsorption Isotherms. Journal of Chemistry, 2017.
https://doi.org/10.1155/2017/3039817
Bai, X., Li, Z., Zhang, Y., Ni, J., Wang, X., & Zhou, X. (2018). Recovery of
Ammonium in Urine by Biochar Derived from Faecal Sludge and its Application as
Soil Conditioner. Waste and Biomass Valorization, 9(9), 1619–1628.
https://doi.org/10.1007/s12649-017-9906-0
114
Ban, Z. S., & Dave, G. (2004). Laboratory studies on recovery of n and p from human
urine through struvite crystallisation and zeolite adsorption. Environmental
Technology (United Kingdom), 25(1), 111–121.
https://doi.org/10.1080/09593330409355443b
Bassan, M., Koné, D., Mbéguéré, M., Holliger, C., & Strande, L. (2015). Success and
failure assessment methodology for wastewater and faecal sludge treatment projects
in low-income countries. Journal of Environmental Planning and Management,
58(10), 1690–1710. https://doi.org/10.1080/09640568.2014.943343
Behrendt, J., Arevalo, E., Gulyas, H., Niederste-Hollenberg, J., Niemiec, A., Zhou, J., &
Otterpohl, R. (2002). Production of value added products from separately collected
urine. Water Science and Technology, 46(6–7), 341–346.
https://doi.org/10.2166/wst.2002.0698
Beler-Baykal, B., Bayram, S., Akkaymak, E., & Cinar, S. (2004). Removal of
ammonium from human urine through ion exchange with clinoptilolite and its
recovery for further reuse. Water Science and Technology, 50(6), 149–156.
https://doi.org/10.2166/wst.2004.0371
Bezerra, M. A., Santelli, R. E., Oliveira, E. P., Villar, L. S., & Escaleira, L. A. (2008).
Response surface methodology (RSM) as a tool for optimization in analytical
chemistry. Talanta, 76(5), 965–977. https://doi.org/10.1016/j.talanta.2008.05.019
Bhattacharjee, A., Jana, B. B., Mandal, S. K., Lahiri, S., & Bhakta, J. N. (2021).
Assessing phosphorus removal potential of laterite soil for water treatment and eco-
technological application. Ecological Engineering, 166(April), 106245.
https://doi.org/10.1016/j.ecoleng.2021.106245
115
Bhattacharya, S. C., Albina, D. O., & Khaing, A. M. (2002). Effects of selected
parameters on performance and emission of biomass- red cookstoves. Biomass and
Bioenergy, 23(2002), 387 – 395.
Bischel, H. N., Özel Duygan, B. D., Strande, L., McArdell, C. S., Udert, K. M., & Kohn,
T. (2015). Pathogens and pharmaceuticals in source-separated urine in eThekwini,
South Africa. Water Research, 85, 57–65.
https://doi.org/10.1016/j.watres.2015.08.022
Bonelli, P. R., Cerrella, E. G., & Cukierman, A. L. (2003). Slow pyrolysis of nutshells:
Characterization of derived chars and of process kinetics. Energy Sources, 25(8),
767–778. https://doi.org/10.1080/00908310390207819
Bonvin, C., Etter, B., Udert, K. M., Frossard, E., Nanzer, S., Tamburini, F., & Oberson,
A. (2015). Plant uptake of phosphorus and nitrogen recycled from synthetic source-
separated urine. Ambio, 44(2), 217–227. https://doi.org/10.1007/s13280-014-0616-
6
Borowski, G., Stȩpniewski, W., & Wójcik-Oliveira, K. (2017). Effect of starch binder on
charcoal briquette properties. International Agrophysics, 31(4), 571–574.
https://doi.org/10.1515/intag-2016-0077
Bu, Q., Lei, H., Wang, L., Wei, Y., Zhu, L., Zhang, X., Liu, Y., Yadavalli, G., & Tang,
J. (2014). Bio-based phenols and fuel production from catalytic microwave
pyrolysis of lignin by activated carbons. Bioresource Technology, 162, 142–147.
https://doi.org/10.1016/j.biortech.2014.03.103
Bulmau, C., Mǎrculescu, C., Badea, A., & Apostol, T. (2010). Pyrolysis parameters
influencing the bio-char generation from wooden biomass. UPB Scientific Bulletin,
Series C: Electrical Engineering, 72(1), 29–38.
116
Cantrell, K. B., Hunt, P. G., Uchimiya, M., Novak, J. M., & Ro, K. S. (2012). Impact of
pyrolysis temperature and manure source on physicochemical characteristics of
biochar. Bioresource Technology, 107, 419–428.
https://doi.org/10.1016/j.biortech.2011.11.084
Carnaje, N. P., Talagon, R. B., Peralta, J. P., Shah, K., & Paz-Ferreiro, J. (2018).
Development and characterisation of charcoal briquettes from water hyacinth
(Eichhornia crassipes)-molasses blend. PLoS ONE, 13(11), 1–14.
https://doi.org/10.1371/journal.pone.0207135
Ceylan, Z., & Sungur, B. (2020). Estimation of coal elemental composition from
proximate analysis using machine learning techniques. Energy Sources, Part A:
Recovery, Utilization and Environmental Effects, 42(20), 2576–2592.
https://doi.org/10.1080/15567036.2020.1790696
Chen, H. L., Haack, V. S., Janecky, C. W., Vollendorf, N. W., & Marlett, J. A. (1998).
Mechanisms by which wheat bran and oat bran increase stool weight in humans.
American Journal of Clinical Nutrition, 68(3), 711–719.
https://doi.org/10.1093/ajcn/68.3.711
Chen, J., Liu, J., He, Y., Huang, L., Sun, S., Sun, J., Chang, K. L., Kuo, J., Huang, S., &
Ning, X. (2017). Investigation of co-combustion characteristics of sewage sludge
and coffee grounds mixtures using thermogravimetric analysis coupled to artificial
neural networks modeling. Bioresource Technology, 225, 234–245.
https://doi.org/10.1016/j.biortech.2016.11.069
Cheng, J. J., Schuster-Wallace, C. J., Watt, S., Newbold, B. K., & Mente, A. (2012). An
117
ecological quantification of the relationships between water, sanitation and infant,
child, and maternal mortality. Environmental Health: A Global Access Science
Source, 11(1). https://doi.org/10.1186/1476-069X-11-4
Chowdhury, Z. Z., Ziaul Karim, M., Ashraf, M. A., & Khalid, K. (2016). Influence of
carbonization temperature on physicochemical properties of biochar derived from
slow pyrolysis of durian wood (Durio zibethinus) sawdust. BioResources, 11(2),
3356–3372. https://doi.org/10.15376/biores.11.2.3356-3372
Clark, W. F., Sontrop, J. M., Macnab, J. J., Suri, R. S., Moist, L., Salvadori, M., & Garg,
A. X. (2011). Urine volume and change in estimated GFR in a community-based
cohort study. Clinical Journal of the American Society of Nephrology, 6(11), 2634–
2641. https://doi.org/10.2215/CJN.01990211
Coats, A. W., & Redfern, J. P. (1964). Kinetic parameters from thermogravimetric data.
Nature, 201(4914), 68–69.
Collins, J., & Lewis, D. (2000). Hydrogen Sulfide : Evaluation Of Current California
Air Quality Standards With Respect To Protection Of Children A . Extended
abstract.
Davies, R. (2013). Ignition and Burning Rate of Water Hyacinth Briquettes. Journal of
Scientific Research and Reports, 2(1), 111–120.
https://doi.org/10.9734/jsrr/2013/1964
de Figueredo, N. A., da Costa, L. M., Melo, L. C. A., Siebeneichlerd, E. A., & Tronto, J.
(2017). Characterization of biochars from different sources and evaluation of
release of nutrients and contaminants. Revista Ciencia Agronomica, 48(3), 395–
118
403. https://doi.org/10.5935/1806-6690.20170046
De Gisi, S., Lofrano, G., Grassi, M., & Notarnicola, M. (2016). Characteristics and
adsorption capacities of low-cost sorbents for wastewater treatment: A review.
Sustainable Materials and Technologies, 9, 10–40.
https://doi.org/10.1016/j.susmat.2016.06.002
Deng, H., Yu, H., Chen, M., & Ge, C. (2014). Sorption of atrazine in tropical soil by
biochar prepared from cassava waste. BioResources, 9(4), 6627–6643.
https://doi.org/10.15376/biores.9.4.6627-6643
Diener, S., Semiyaga, S., Niwagaba, C. B., Muspratt, A. M., Gning, J. B., Mbéguéré, M.,
Ennin, J. E., Zurbrugg, C., & Strande, L. (2014). A value proposition: Resource
recovery from faecal sludge - Can it be the driver for improved sanitation?
Resources, Conservation and Recycling, 88, 32–38.
https://doi.org/10.1016/j.resconrec.2014.04.005
Dirbeba, M. J., Brink, A., Lindberg, D., Hupa, M., & Hupa, L. (2021). Thermal
Conversion Characteristics of Molasses. ACS Omega, 6(33), 21631–21645.
https://doi.org/10.1021/acsomega.1c03024
Drenkova-Tuhtan, A., Schneider, M., Franzreb, M., Meyer, C., Gellermann, C., Sextl,
G., Mandel, K., & Steinmetz, H. (2017). Pilot-scale removal and recovery of
dissolved phosphate from secondary wastewater effluents with reusable ZnFeZr
adsorbent @ Fe3O4/SiO2 particles with magnetic harvesting. Water Research, 109,
77–87. https://doi.org/10.1016/j.watres.2016.11.039
Dutta, P., Mandal, S., & Kumar, A. (2018). Comparative study: FPA based response
surface methodology & ANOVA for the parameter optimization in process control.
119
Advances in Modelling and Analysis C, 73(1), 23–27.
https://doi.org/10.18280/ama_c.730104
Eberhardt, T. L., Min, S. H., & Han, J. S. (2006). Phosphate removal by refined aspen
wood fiber treated with carboxymethyl cellulose and ferrous chloride. Bioresource
Technology, 97(18), 2371–2376. https://doi.org/10.1016/j.biortech.2005.10.040
Ehujuo, N., Ehujuo, *, Okeke, N. N., & Akaolisa, O. C. (2017). Geotechnical Properties
of Lateritic Soils Derived from Various Geologic Formations in Okigwe Area,
Southeastern Nigeria. Geotechnical Properties Of… Futo Journal Series
(FUTOJNLS), 3(2), 178–189. www.futojnls.org
Elehinafe, F. B., Okedere, O. B., Fakinle, B. S., & Sonibare, J. A. (2017). Assessment of
sawdust of different wood species in Southwestern Nigeria as source of energy.
Energy Sources, Part A: Recovery, Utilization and Environmental Effects, 39(18),
1901–1905. https://doi.org/10.1080/15567036.2017.1384869
Esrey, S. A., Gough, J., Rapaport, D., Sawyer, R., Simpson-Hébert, M., Vargas, J., &
Winblad, U. (1998). Ecological Sanitataion. Swedish International Development
Cooperation Agency, 1–100.
Etter, B., Tilley, E., Khadka, R., & Udert, K. M. (2011). Low-cost struvite production
using source-separated urine in Nepal. Water Research, 45(2), 852–862.
https://doi.org/10.1016/j.watres.2010.10.007
Falemara, B. C., Joshua, V. I., Aina, O. O., & Nuhu, R. D. (2018). Performance
evaluation of the physical and combustion properties of briquettes produced from
agro-wastes and wood residues. Recycling, 3(3), 1–13.
120
https://doi.org/10.3390/recycling3030037
Farrow, T. S., Sun, C., & Snape, C. E. (2013). Impact of biomass char on coal char burn-
out under air and oxy-fuel conditions. Fuel, 114, 128–134.
https://doi.org/10.1016/j.fuel.2012.07.073
Feachem, R., Bradley, D., Garelick, H., & Mara, D. (1984). Sanitation and Disease ‐
Health Aspects of Excreta and Waste‐water Management. JAWRA Journal of the
American Water Resources Association, 20(5), 803–803.
https://doi.org/10.1111/j.1752-1688.1984.tb04765.x
Fidalgo, B., Chilmeran, M., Somorin, T., Sowale, A., Kolios, A., Parker, A., Williams,
L., Collins, M., McAdam, E. J., & Tyrrel, S. (2019). Non-isothermal
thermogravimetric kinetic analysis of the thermochemical conversion of human
faeces. Renewable Energy, 132, 1177–1184.
https://doi.org/10.1016/j.renene.2018.08.090
Finnish Standards SFS 3026. (1986). Finnish standard for determination of total
phosphorus in water. Digestion with peroxodisulfate. In Finnish Standards
Association, Helsinki, Finland.
Flynn, J. H.; Wall, L. A. (1966). A quick, direct method for the determination of
activation energy from thermogravimetric data. Journal of Polymer Science Part B:
Polymer Letters, 4(5), 323–328. https://doi.org/10.1098/rstb.1988.0133
Foo, K. Y., & Hameed, B. H. (2010). Insights into the modeling of adsorption isotherm
systems. Chemical Engineering Journal, 156(1), 2–10.
https://doi.org/10.1016/j.cej.2009.09.013
Foo, K. Y., & Hameed, B. H. (2012). Porous structure and adsorptive properties of
pineapple peel based activated carbons prepared via microwave assisted KOH and
K2CO 3 activation. Microporous and Mesoporous Materials, 148(1), 191–195.
121
https://doi.org/10.1016/j.micromeso.2011.08.005
Frame, M. H., & Schandl, C. A. (2015). A case example of asphyxia due to occupational
exposure to airborne chemicals and review of workplace fatalities. Journal of
Forensic Sciences, 60(2), 521–524. https://doi.org/10.1111/1556-4029.12695
Fu, B., Ge, C., Yue, L., Luo, J., Feng, D., Deng, H., & Yu, H. (2016). Characterization
of Biochar Derived from Pineapple Peel Waste and Its Application for Sorption of
Oxytetracycline from Aqueous Solution. BioResources, 11(4).
https://doi.org/10.15376/biores.11.4.9017-9035
G.Manić, N., Ž.Janković, B., D.Stojiljković, D., V.Jovanović, V., & Radojević, M. B.
(2019). Tga-Dsc-Ms Analysis of Pyrolysis Process. Thermal Science, 23, 1457–
1472.
Gai, C., Dong, Y., & Zhang, T. (2013). The kinetic analysis of the pyrolysis of
agricultural residue under non-isothermal conditions. Bioresource Technology, 127,
298–305. https://doi.org/10.1016/j.biortech.2012.09.089
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), 1–19. https://doi.org/10.1371/journal.pone.0113888
Ganrot, Z., Dave, G., & Nilsson, E. (2007a). Recovery of N and P from human urine by
freezing, struvite precipitation and adsorption to zeolite and active carbon.
Bioresource Technology, 98(16), 3112–3121.
https://doi.org/10.1016/j.biortech.2006.10.038
122
Ganrot, Z., Dave, G., & Nilsson, E. (2007b). Recovery of N and P from human urine by
freezing, struvite precipitation and adsorption to zeolite and active carbon.
Bioresource Technology, 98(16), 3112–3121.
https://doi.org/10.1016/j.biortech.2006.10.038
Githeko., C. N. &. (2013). Effect of Binder Types and Amount on. International Journal
of Engineering Research and Science and Technology, 2(1), 12–20.
Glarborg, P., Miller, J. A., Ruscic, B., & Klippenstein, S. J. (2018). Modeling nitrogen
chemistry in combustion. Progress in Energy and Combustion Science, 67, 31–68.
https://doi.org/10.1016/j.pecs.2018.01.002
Glocheux, Y., Pasarín, M. M., Albadarin, A. B., Allen, S. J., & Walker, G. M. (2013).
Removal of arsenic from groundwater by adsorption onto an acidified laterite by-
product. Chemical Engineering Journal, 228, 565–574.
https://doi.org/10.1016/j.cej.2013.05.043
Gold, M., Niang, S., Niwagaba, C. B., Eder, G., Muspratt, A. M., Diop, S. P., & Strande,
L. (2014). Results from FaME (Faecal Management Enterprises) – can dried faecal
sludge fuel the sanitation service chain? 37th WEDC International Conference,
December, 1–6. http://wedc.lboro.ac.uk/resources/conference/37/Gold-2026.pdf
Grønli, M. G., Várhegyi, G., & Di Blasi, C. (2002). Thermogravimetric analysis and
devolatilization kinetics of wood. Industrial and Engineering Chemistry Research,
41(17), 4201–4208. https://doi.org/10.1021/ie0201157
123
Pratices.Regional wood energy development program in Asia, field document no.
46. Bangkok, Thailand: Food and Agriculture Organization of the United Nations;
Regional Wood Energy Development Programme in Asia, 46, 1–48.
Guo, J., Song, Y., Ji, X., Ji, L., Cai, L., Wang, Y., Zhang, H., & Song, W. (2019).
Preparation and characterization of nanoporous activated carbon derived from
prawn shell and its application for removal of heavy metal ions. Materials, 12(2).
https://doi.org/10.3390/ma12020241
Gwara, S., Wale, E., Odindo, A., & Buckley, C. (2021). Attitudes and perceptions on the
agricultural use of human excreta and human excreta derived materials: A scoping
review. Agriculture (Switzerland), 11(2), 1–30.
https://doi.org/10.3390/agriculture11020153
Habib, U., Habib, M., & Khan, A. U. (2014). Factors influencing the performance of
coal briquettes. Walailak Journal of Science and Technology, 11(1), 1–6.
https://doi.org/10.2004/wjst.v11i1.417
Hafford, L. M., Ward, B. J., Weimer, A. W., & Linden, K. (2018). Fecal sludge as a
fuel: Characterization, cofire limits, and evaluation of quality improvement
measures. Water Science and Technology, 78(12), 2437–2448.
https://doi.org/10.2166/wst.2019.005
Hale, S. E., Alling, V., Martinsen, V., Mulder, J., Breedveld, G. D., & Cornelissen, G.
(2013). The sorption and desorption of phosphate-P, ammonium-N and nitrate-N in
cacao shell and corn cob biochars. Chemosphere, 91(11), 1612–1619.
https://doi.org/10.1016/j.chemosphere.2012.12.057
Halim, S. F., Yong, S. K., & Tay, C. C. (2017). Ammonia nitrogen adsorption using
spent mushroom substrate biochar (Smsb). Pertanika Journal of Science and
Technology, 25(S7), 9–20.
124
Hameed, B. H., & El-Khaiary, M. I. (2008). Malachite green adsorption by rattan
sawdust: Isotherm, kinetic and mechanism modeling. Journal of Hazardous
Materials, 159(2–3), 574–579. https://doi.org/10.1016/j.jhazmat.2008.02.054
Harder, R., Wielemaker, R., Larsen, T. A., Zeeman, G., & Öberg, G. (2019). Recycling
nutrients contained in human excreta to agriculture: Pathways, processes, and
products. Critical Reviews in Environmental Science and Technology, 49(8), 695–
743. https://doi.org/10.1080/10643389.2018.1558889
Hayhurst, A. N., & Lawrence, A. D. (1996). The amounts of NOx and N2O formed in a
fluidized bed combustor during the burning of coal volatiles and also of char.
Combustion and Flame, 105(3), 341–357. https://doi.org/10.1016/0010-
2180(95)00215-4
He, C., Giannis, A., & Wang, J. Y. (2013). Conversion of sewage sludge to clean solid
fuel using hydrothermal carbonization: Hydrochar fuel characteristics and
combustion behavior. Applied Energy, 111, 257–266.
https://doi.org/10.1016/j.apenergy.2013.04.084
He, X., Liu, Z., Niu, W., Yang, L., Zhou, T., Qin, D., Niu, Z., & Yuan, Q. (2018).
Effects of pyrolysis temperature on the physicochemical properties of gas and
biochar obtained from pyrolysis of crop residues. Energy, 143, 746–756.
https://doi.org/10.1016/j.energy.2017.11.062
Hernández, A. B., Okonta, F., & Freeman, N. (2017). Thermal decomposition of sewage
sludge under N2, CO2 and air: Gas characterization and kinetic analysis. Journal of
Environmental Management, 196, 560–568.
https://doi.org/10.1016/j.jenvman.2017.03.036
Heydari, M., Rahman, M., & Gupta, R. (2015). Kinetic study and thermal decomposition
behavior of lignite coal. International Journal of Chemical Engineering, 2015.
https://doi.org/10.1155/2015/481739
125
Hikal, W. M., Mahmoud, A. A., Said-Al Ahl, H. A. H., Bratovcic, A., Tkachenko, K. G.,
Kačániová, M., & Rodriguez, R. M. (2021). Pineapple (<i>Ananas
comosus</i> L. Merr.), Waste Streams, Characterisation and Valorisation: An
Overview. Open Journal of Ecology, 11(09), 610–634.
https://doi.org/10.4236/oje.2021.119039
Höglund, C., Vinneras, B., Stenström, T. A., & Jönsson, H. (2000). Variation of
chemical and microbial parameters in collection and storage tanks for source
separated human urine. Journal of Environmental Science and Health - Part A
Toxic/Hazardous Substances and Environmental Engineering, 35(8), 1463–1475.
https://doi.org/10.1080/10934520009377047
Hu, X., Zhang, X., Ngo, H. H., Guo, W., Wen, H., Li, C., Zhang, Y., & Ma, C. (2020).
Comparison study on the ammonium adsorption of the biochars derived from
different kinds of fruit peel. Science of the Total Environment, 707, 135544.
https://doi.org/10.1016/j.scitotenv.2019.135544
Huang, H., Xiao, X., Yan, B., & Yang, L. (2010). Ammonium removal from aqueous
solutions by using natural Chinese (Chende) zeolite as adsorbent. Journal of
Hazardous Materials. https://doi.org/10.1016/j.jhazmat.2009.09.156
126
Hue, N. T., & Tung, N. H. (2018). Study on simultaneous adsorption of phosphate and
fluoride from water environment by modified laterite ore from Northern Vietnam.
Green Processing and Synthesis, 7(2), 89–99. https://doi.org/10.1515/gps-2016-
0136
Inyinbor, A. A., Adekola, F. A., & Olatunji, G. A. (2016). Kinetics, isotherms and
thermodynamic modeling of liquid phase adsorption of Rhodamine B dye onto
Raphia hookerie fruit epicarp. Water Resources and Industry, 15, 14–27.
https://doi.org/10.1016/j.wri.2016.06.001
Iriel, A., Bruneel, S. P., Schenone, N., & Cirelli, A. F. (2018). The removal of fluoride
from aqueous solution by a lateritic soil adsorption: Kinetic and equilibrium
studies. Ecotoxicology and Environmental Safety, 149(November 2017), 166–172.
https://doi.org/10.1016/j.ecoenv.2017.11.016
Jana, B. B., Rana, S., & Bag, S. K. (2012). Use of human urine in phytoplankton
production as a tool for ecological sanitation. Water Science and Technology,
65(8), 1350–1356. https://doi.org/10.2166/wst.2012.044
Kakali, G., Perraki, T., Tsivilis, S., & Badogiannis, E. (2001). Thermal treatment of
kaolin: The effect of mineralogy on the pozzolanic activity. Applied Clay Science,
20(1–2), 73–80. https://doi.org/10.1016/S0169-1317(01)00040-0
Kamtchueng, B. T., Onana, V. L., Fantong, W. Y., Ueda, A., Ntouala, R. F., Wongolo,
M. H., Ndongo, G. B., Ze, A. N., Kamgang, V. K., & Ondoa, J. M. (2015).
Geotechnical, chemical and mineralogical evaluation of lateritic soils in humid
tropical area (Mfou, Central-Cameroon): Implications for road construction.
International Journal of Geo-Engineering, 6(1), 1–21.
127
https://doi.org/10.1186/s40703-014-0001-0
Kanbara, A., Miura, Y., Hyogo, H., Chayama, K., & Seyama, I. (2012). Effect of urine
pH changed by dietary intervention on uric acid clearance mechanism of pH-
dependent excretion of urinary uric acid. Nutrition Journal, 11(1), 1–7.
https://doi.org/10.1186/1475-2891-11-39
Karahalios, T., Berner, C., & Njenga, M. (2018). Human Waste-to-fuel Briquettes as a
Sanitation and Energy Solution for Refugee Camps and Informal Urban
Settlements. Recovering Bioenergy in Sub-Saharan Africa: Gender Dimensions,
Lessons and Challenges, December, 96.
https://www.researchgate.net/profile/Mary_Njenga/publication/329557889_Recove
ring_Bioenergy_in_Sub-Saharan_Africa_Gender_Dimensions_Lessons_and_
Challenges/links/5c0f8b5e299bf139c74fe81e/Recovering-Bioenergy-in-Sub-
Saharan-Africa-Gender-Dimensions-Lessons-
Karak, T., & Bhattacharyya, P. (2011). Human urine as a source of alternative natural
fertilizer in agriculture: A flight of fancy or an achievable reality. Resources,
Conservation and Recycling, 55(4), 400–408.
https://doi.org/10.1016/j.resconrec.2010.12.008
Kaur, S., Rani, S., Mahajan, R. K., Asif, M., & Gupta, V. K. (2015). Synthesis and
adsorption properties of mesoporous material for the removal of dye safranin:
Kinetics, equilibrium, and thermodynamics. Journal of Industrial and Engineering
Chemistry, 22, 19–27. https://doi.org/10.1016/j.jiec.2014.06.019
Kavvada, O., Tarpeh, W. A., Horvath, A., & Nelson, K. L. (2017). Life-cycle cost and
environmental assessment of decentralized nitrogen recovery using ion exchange
from source-separated urine through spatial modeling. Environmental Science and
Technology, 51(21), 12061–12071. https://doi.org/10.1021/acs.est.7b02244
Keiluweit, M., Nico, P. S., Johnson, M., & Kleber, M. (2010). Dynamic molecular
128
structure of plant biomass-derived black carbon (biochar). Environmental Science
and Technology, 44(4), 1247–1253. https://doi.org/10.1021/es9031419
Kim, K. H., Kim, J. Y., Cho, T. S., & Choi, J. W. (2012). Influence of pyrolysis
temperature on physicochemical properties of biochar obtained from the fast
pyrolysis of pitch pine (Pinus rigida). Bioresource Technology, 118, 158–162.
https://doi.org/10.1016/j.biortech.2012.04.094
Kim, Y., Park, T., & Hong, D. (2021). Heating and emission characteristics of briquettes
developed from spent coffee grounds. Environmental Engineering Research, 27(4),
210063–0. https://doi.org/10.4491/eer.2021.063
Kirchmann, H., & Pettersson, S. (1994). Human urine - Chemical composition and
fertilizer use efficiency. Fertilizer Research, 40(2), 149–154.
https://doi.org/10.1007/BF00750100
Kitano, M., Inoue, Y., Yamazaki, Y., Hayashi, F., Kanbara, S., Matsuishi, S.,
Yokoyama, T., Kim, S. W., Hara, M., & Hosono, H. (2012). Ammonia synthesis
using a stable electride as an electron donor and reversible hydrogen store. Nature
Chemistry, 4(11), 934–940. https://doi.org/10.1038/nchem.1476
Kithome, M., Paul, J. W., Lavkulich, L. M., & Bomke, A. A. (1998). Kinetics of
Ammonium Adsorption and Desorption by the Natural Zeolite Clinoptilolite. Soil
Science Society of America Journal, 62(3), 622–629.
https://doi.org/10.2136/sssaj1998.03615995006200030011x
129
Kizito, S., Wu, S., Kipkemoi Kirui, W., Lei, M., Lu, Q., Bah, H., & Dong, R. (2015).
Evaluation of slow pyrolyzed wood and rice husks biochar for adsorption of
ammonium nitrogen from piggery manure anaerobic digestate slurry. Science of the
Total Environment, 505, 102–112. https://doi.org/10.1016/j.scitotenv.2014.09.096
Krueger, B. C., Fowler, G. D., Templeton, M. R., & Moya, B. (2020). Resource
recovery and biochar characteristics from full-scale faecal sludge treatment and co-
treatment with agricultural waste. Water Research, 169, 115253.
https://doi.org/10.1016/j.watres.2019.115253
Kumar, D., Singh, A., & Gaur, J. P. (2008). Mono-component versus binary isotherm
models for Cu(II) and Pb(II) sorption from binary metal solution by the green alga
Pithophora oedogonia. Bioresource Technology, 99(17), 8280–8287.
https://doi.org/10.1016/j.biortech.2008.03.008
Kuntke, P., Sleutels, T. H. J. A., Saakes, M., & Buisman, C. J. N. (2014). Hydrogen
production and ammonium recovery from urine by a Microbial Electrolysis Cell.
International Journal of Hydrogen Energy, 39(10), 4771–4778.
https://doi.org/10.1016/j.ijhydene.2013.10.089
Lam, S. S., Liew, R. K., Cheng, C. K., Rasit, N., Ooi, C. K., Ma, N. L., Ng, J. H., Lam,
W. H., Chong, C. T., & Chase, H. A. (2018). Pyrolysis production of fruit peel
biochar for potential use in treatment of palm oil mill effluent. Journal of
Environmental Management, 213, 400–408.
https://doi.org/10.1016/j.jenvman.2018.02.092
130
Larsen, T. A., Alder, A. C., Eggen, R. L., Maurer, M., & Lienert, J. (2009). A new
planning and design paradigm to achieve sustainable resource recovery from
wastewater. Environmental Science and Technology, 43(16), 6126–6130.
https://doi.org/10.1021/es9010515
Ledezma, P., Kuntke, P., Buisman, C. J. N., Keller, J., & Freguia, S. (2015). Source-
separated urine opens golden opportunities for microbial electrochemical
technologies. Trends in Biotechnology, 33(4), 214–220.
https://doi.org/10.1016/j.tibtech.2015.01.007
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.
https://doi.org/10.1016/j.biortech.2013.08.135
Lemougna, P. N., Madi, A. B., Kamseu, E., Melo, U. C., Delplancke, M. P., & Rahier,
H. (2014a). Influence of the processing temperature on the compressive strength of
Na activated lateritic soil for building applications. Construction and Building
Materials, 65, 60–66. https://doi.org/10.1016/j.conbuildmat.2014.04.100
Lemougna, P. N., Madi, A. B., Kamseu, E., Melo, U. C., Delplancke, M. P., & Rahier,
H. (2014b). Influence of the processing temperature on the compressive strength of
Na activated lateritic soil for building applications. Construction and Building
Materials, 65, 60–66. https://doi.org/10.1016/j.conbuildmat.2014.04.100
Li, B., Huang, H. M., Boiarkina, I., Yu, W., Huang, Y. F., Wang, G. Q., & Young, B. R.
(2019). Phosphorus recovery through struvite crystallisation: Recent developments
in the understanding of operational factors. Journal of Environmental Management,
248(January), 109254. https://doi.org/10.1016/j.jenvman.2019.07.025
Li, H., Mahyoub, S. A. A., Liao, W., Xia, S., Zhao, H., Guo, M., & Ma, P. (2017). Effect
of pyrolysis temperature on characteristics and aromatic contaminants adsorption
131
behavior of magnetic biochar derived from pyrolysis oil distillation residue.
Bioresource Technology, 223, 20–26.
https://doi.org/10.1016/j.biortech.2016.10.033
Li, J., Ma, C., Ma, Y., Li, Y., Zhou, W., & Xu, P. (2007). Medium optimization by
combination of response surface methodology and desirability function: An
application in glutamine production. Applied Microbiology and Biotechnology,
74(3), 563–571. https://doi.org/10.1007/s00253-006-0699-5
Li, R., Wang, J. J., Zhou, B., Zhang, Z., Liu, S., Lei, S., & Xiao, R. (2017).
Simultaneous capture removal of phosphate, ammonium and organic substances by
MgO impregnated biochar and its potential use in swine wastewater treatment.
Journal of Cleaner Production, 147, 96–107. https://doi.org/10.1016/
j.jclepro.2017.01.069
Liang, H., Chen, L., Liu, G., & Zheng, H. (2016). Surface morphology properties of
biochars produced from different feedstocks. Iccte, 1205–1208.
https://doi.org/10.2991/iccte-16.2016.210
Lienert, J., Bürki, T., & Escher, B. I. (2007). Reducing micropollutants with source
control: Substance flow analysis of 212 pharmaceuticals in faeces and urine. Water
Science and Technology, 56(5), 87–96. https://doi.org/10.2166/wst.2007.560
Lim, A. C. R., Chin, B. L. F., Jawad, Z. A., & Hii, K. L. (2016). Kinetic Analysis of
Rice Husk Pyrolysis Using Kissinger-Akahira-Sunose (KAS) Method. Procedia
Engineering, 148, 1247–1251. https://doi.org/10.1016/j.proeng.2016.06.486
Lind, B. B., Ban, Z., & Bydén, S. (2000). Nutrient recovery from human urine by
struvite crystallization with ammonia adsorption on zeolite and wollastonite.
Bioresource Technology, 73(2), 169–174. https://doi.org/10.1016/S0960-
8524(99)90157-8
132
Liu, J., Jiang, J., Aihemaiti, A., Meng, Y., Yang, M., Xu, Y., Gao, Y., Zou, Q., & Chen,
X. (2019). Removal of phosphate from aqueous solution using MgO-modified
magnetic biochar derived from anaerobic digestion residue. Journal of
Environmental Management, 250(June), 109438.
https://doi.org/10.1016/j.jenvman.2019.109438
Liu, Y., He, Z., & Uchimiya, M. (2015). Comparison of Biochar Formation from
Various Agricultural By-Products Using FTIR Spectroscopy. Modern Applied
Science, 9(4), 246–253. https://doi.org/10.5539/mas.v9n4p246
Lun, L. W., Gunny, A. A. N., Kasim, F. H., & Arbain, D. (2017). Fourier transform
infrared spectroscopy (FTIR) analysis of paddy straw pulp treated using deep
eutectic solvent. AIP Conference Proceedings, 1835(April 2017).
https://doi.org/10.1063/1.4981871
Luo, L., Wang, G., Shi, G., Zhang, M., Zhang, J., He, J., Xiao, Y., Tian, D., Zhang, Y.,
Deng, S., Zhou, W., Lan, T., & Deng, O. (2019). The characterization of biochars
derived from rice straw and swine manure, and their potential and risk in N and P
removal from water. Journal of Environmental Management, 245(January), 1–7.
https://doi.org/10.1016/j.jenvman.2019.05.072
Ma, Z., Wang, J., Yang, Y., Zhang, Y., Zhao, C., Yu, Y., & Wang, S. (2018).
Comparison of the thermal degradation behaviors and kinetics of palm oil waste
under nitrogen and air atmosphere in TGA-FTIR with a complementary use of
model-free and model-fitting approaches. In Journal of Analytical and Applied
Pyrolysis (Vol. 134). Elsevier B.V. https://doi.org/10.1016/j.jaap.2018.04.002
133
as adsorbent for Fe (III) from aqueous solution. Beni-Suef University Journal of
Basic and Applied Sciences, 4(2), 142–149.
https://doi.org/10.1016/j.bjbas.2015.05.008
Maiti, A., Basu, J. K., & De, S. (2012). Experimental and kinetic modeling of As(V) and
As(III) adsorption on treated laterite using synthetic and contaminated
groundwater: Effects of phosphate, silicate and carbonate ions. Chemical
Engineering Journal, 191, 1–12. https://doi.org/10.1016/j.cej.2010.01.031
Maiti, A., Thakur, B. K., Basu, J. K., & De, S. (2013). Comparison of treated laterite as
arsenic adsorbent from different locations and performance of best filter under field
conditions. Journal of Hazardous Materials, 262, 1176–1186.
https://doi.org/10.1016/j.jhazmat.2012.06.036
Malone Rubright, S. L., Pearce, L. L., & Peterson, J. (2017). Environmental toxicology
of hydrogen sulfide. Nitric Oxide - Biology and Chemistry, 71(412), 1–13.
https://doi.org/10.1016/j.niox.2017.09.011
Mambo, W. (2016). Optimal compaction pressure, particle size and binder ratio for
quality briquettes made from maize cobs.
http://ir.jkuat.ac.ke/handle/123456789/2387
Mansing R, P., & Rout, P. . (2013). Removal of phosphorus from sewage effluent by
adsorption on Laterite. International Journal of Engineering Research &
Technology, 2(9), 551–559.
Martin, T. M. P., Esculier, F., Levavasseur, F., & Houot, S. (2020). Human urine-based
fertilizers: A review. Critical Reviews in Environmental Science and Technology,
0(0), 1–47. https://doi.org/10.1080/10643389.2020.1838214
Mathurasa, L., & Damrongsiri, S. (2018). Low cost and easy rice husk modification to
efficiently enhance ammonium and nitrate adsorption. International Journal of
134
Recycling of Organic Waste in Agriculture, 7(2), 143–151.
https://doi.org/10.1007/s40093-018-0200-3
Matsui, Y., Nakao, S., Sakamoto, A., Taniguchi, T., Pan, L., Matsushita, T., &
Shirasaki, N. (2015). Adsorption capacities of activated carbons for geosmin and 2-
methylisoborneol vary with activated carbon particle size: Effects of adsorbent and
adsorbate characteristics. Water Research, 85, 95–102. https://doi.org/10.1016/
j.watres.2015.08.017
Maurer, M., Pronk, W., & Larsen, T. A. (2006a). Treatment processes for source-
separated urine. Water Research, 40(17), 3151–3166. https://doi.org/10.1016/
j.watres.2006.07.012
Maurer, M., Pronk, W., & Larsen, T. A. (2006b). Treatment processes for source-
separated urine. In Water Research (Vol. 40, Issue 17, pp. 3151–3166).
https://doi.org/10.1016/j.watres.2006.07.012
Mishra, R. K., & Mohanty, K. (2018). Pyrolysis kinetics and thermal behavior of waste
sawdust biomass using thermogravimetric analysis. Bioresource Technology, 251,
63–74. https://doi.org/10.1016/j.biortech.2017.12.029
Mitra, S., Thakur, L. S., Rathore, V. K., & Mondal, P. (2016). Removal of Pb(II) and
Cr(VI) by laterite soil from synthetic waste water: single and bi-component
adsorption approach. Desalination and Water Treatment, 57(39), 18406–18416.
https://doi.org/10.1080/19443994.2015.1088806
Mouni, L., Belkhiri, L., Bollinger, J. C., Bouzaza, A., Assadi, A., Tirri, A., Dahmoune,
F., Madani, K., & Remini, H. (2018). Removal of Methylene Blue from aqueous
solutions by adsorption on Kaolin: Kinetic and equilibrium studies. Applied Clay
Science, 153, 38–45. https://doi.org/10.1016/j.clay.2017.11.034
Muspratt, A. M., Nakato, T., Niwagaba, C., Dione, H., Kang, J., Stupin, L., Regulinski,
135
J., Mbéguéré, M., & Strande, L. (2014). Fuel potential of faecal sludge: Calorific
value results from Uganda, Ghana and Senegal. Journal of Water Sanitation and
Hygiene for Development, 4(2), 223–230.
https://doi.org/10.2166/washdev.2013.055
Myers, R. H., Montgomery, D. C., Geoffrey Vining, G., Borror, C. M., & Kowalski, S.
M. (2004a). Response Surface Methodology: A Retrospective and Literature
Survey. Journal of Quality Technology, 36(1), 53–78.
https://doi.org/10.1080/00224065.2004.11980252
Myers, R. H., Montgomery, D. C., Geoffrey Vining, G., Borror, C. M., & Kowalski, S.
M. (2004b). Response Surface Methodology: A Retrospective and Literature
Survey. In Journal of Quality Technology (Vol. 36, Issue 1, pp. 53–78).
https://doi.org/10.1080/00224065.2004.11980252
Nellie Oduor, Emily Kitheka, Celestine Ingutia, Nathan Nyamai, James Kimwemwe, &
Kevin Juma. (2019). Quality and Emission Analysis of Charcoal from Various
Species of Wood Using Improved Carbonization Technologies in Kenya. Journal
of Environmental Science and Engineering A, 8(1). https://doi.org/10.17265/2162-
5298/2019.01.002
Neset, T. S. S., & Cordell, D. (2012). Global phosphorus scarcity: Identifying synergies
for a sustainable future. Journal of the Science of Food and Agriculture, 92(1), 2–6.
https://doi.org/10.1002/jsfa.4650
Ng, C. W. W., Akinniyi, D. B., Zhou, C., & Chiu, C. F. (2019). Comparisons of
weathered lateritic, granitic and volcanic soils: Compressibility and shear strength.
Engineering Geology, 249, 235–240. https://doi.org/10.1016/j.enggeo.2018.12.029
Ngusale, G. K., Luo, Y., & Kiplagat, J. K. (2014). Briquette making in Kenya: Nairobi
and peri-urban areas. Renewable and Sustainable Energy Reviews, 40, 749–759.
https://doi.org/10.1016/j.rser.2014.07.206
136
Nguyen-Viet, H., Zinsstag, J., Schertenleib, R., Zurbrügg, C., Obrist, B., Montangero,
A., Surkinkul, N., Koné, D., Morel, A., Cissé, G., Koottatep, T., Bonfoh, B., &
Tanner, M. (2009). Improving environmental sanitation, health, and well-being: A
conceptual framework for integral interventions. EcoHealth, 6(2), 180–191.
https://doi.org/10.1007/s10393-009-0249-6
Nguyen, T. T. N., Némery, J., Gratiot, N., Strady, E., Tran, V. Q., Nguyen, A. T., Aimé,
J., & Peyne, A. (2019). Nutrient dynamics and eutrophication assessment in the
tropical river system of Saigon – Dongnai (southern Vietnam). Science of the Total
Environment, 653, 370–383. https://doi.org/10.1016/j.scitotenv.2018.10.319
Njenga, M., Karanja, N., Jamnadass, R., Kithinji, J., Sundberg, C., & Jirjis, R. (2013).
Quality of cooking fuel briquettes produced locally from charcoal dust and sawdust
in Kenya. Journal of Biobased Materials and Bioenergy, 7(3), 315–322.
https://doi.org/10.1166/jbmb.2013.1355
Nyberg, K. A., Ottoson, J. R., Vinnerås, B., & Albihn, A. (2014). Fate and survival of
Salmonella Typhimurium and Escherichia coli O157:H7 in repacked soil lysimeters
after application of cattle slurry and human urine. Journal of the Science of Food
and Agriculture, 94(12), 2541–2546. https://doi.org/10.1002/jsfa.6593
Obernberger, I., Brunner, T., & Bärnthaler, G. (2006). Chemical properties of solid
biofuels-significance and impact. Biomass and Bioenergy, 30(11), 973–982.
https://doi.org/10.1016/j.biombioe.2006.06.011
Onabanjo, T., Patchigolla, K., Wagland, S. T., Fidalgo, B., Kolios, A., McAdam, E.,
Parker, A., Williams, L., Tyrrel, S., & Cartmell, E. (2016). Energy recovery from
human faeces via gasification: A thermodynamic equilibrium modelling approach.
Energy Conversion and Management, 118, 364–376. https://doi.org/10.1016/
j.enconman.2016.04.005
137
Osei, J., Gawu, S. K., Schäfer, A. I., Atipoka, F. A., & Momade, F. W. (2016). Impact of
laterite characteristics on fluoride removal from water. Journal of Chemical
Technology and Biotechnology, 91(4), 911–920. https://doi.org/10.1002/jctb.4656
OSHA. (2010). Carbon Dioxide Health Hazard Information Sheet. The FSIS
Environmental Safety and Health Group (ESHG), 2–4.
https://www.osha.gov/dts/chemicalsampling/data/CH_225400.html
Otieno, A. O., Home, P. G., Raude, J. M., Murunga, S. I., Ngumba, E., Ojwang, D. O.,
& Tuhkanen, T. (2021). Pineapple peel biochar and lateritic soil as adsorbents for
recovery of ammonium nitrogen from human urine. Journal of Environmental
Management, 293(May), 112794. https://doi.org/10.1016/j.jenvman.2021.112794
Öztürk, N., & Bektaş, T. E. (2004). Nitrate removal from aqueous solution by adsorption
onto various materials. Journal of Hazardous Materials, 112(1–2), 155–162.
https://doi.org/10.1016/j.jhazmat.2004.05.001
Palmonari, A., Cavallini, D., Sniffen, C. J., Fernandes, L., Holder, P., Fagioli, L.,
Fusaro, I., Biagi, G., Formigoni, A., & Mammi, L. (2020). Short communication:
Characterization of molasses chemical composition. Journal of Dairy Science,
103(7), 6244–6249. https://doi.org/10.3168/jds.2019-17644
Parikh, J., Channiwala, S. A., & Ghosal, G. K. (2007). A correlation for calculating
elemental composition from proximate analysis of biomass materials. Fuel, 86(12–
138
13), 1710–1719. https://doi.org/10.1016/j.fuel.2006.12.029
Patel, S. U., Jeevan Kumar, B., Badhe, Y. P., Sharma, B. K., Saha, S., Biswas, S.,
Chaudhury, A., Tambe, S. S., & Kulkarni, B. D. (2007). Estimation of gross
calorific value of coals using artificial neural networks. Fuel, 86(3), 334–344.
https://doi.org/10.1016/j.fuel.2006.07.036
Pillai, M. G., Simha, P., & Gugalia, A. (2014). Recovering urea from human urine by
bio-sorption onto Microwave Activated Carbonized Coconut Shells: Equilibrium,
kinetics, optimization and field studies. Journal of Environmental Chemical
Engineering. https://doi.org/10.1016/j.jece.2013.11.027
Pilusa, T. J., Huberts, R., & Muzenda, E. (2013). Emissions analysis from combustion of
eco-fuel briquettes for domestic applications. Journal of Energy in Southern Africa,
24(4), 30–36. https://doi.org/10.17159/2413-3051/2013/v24i4a3143
Pradhan, S. K., Mikola, A., Heinonen-Tanski, H., & Vahala, R. (2019). Recovery of
nitrogen and phosphorus from human urine using membrane and precipitation
process. Journal of Environmental Management, 247, 596–602.
https://doi.org/10.1016/j.jenvman.2019.06.046
139
Qin, Y., Zhu, X., Su, Q., Anumah, A., Gao, B., Lyu, W., Zhou, X., Xing, Y., & Wang,
B. (2019). Enhanced removal of ammonium from water by ball-milled biochar.
Environmental Geochemistry and Health, 5. https://doi.org/10.1007/s10653-019-
00474-5
Ragheb, S. M. (2013). Phosphate removal from aqueous solution using slag and fly ash.
HBRC Journal, 9(3), 270–275. https://doi.org/10.1016/j.hbrcj.2013.08.005
Randall, D. G., Krähenbühl, M., Köpping, I., Larsen, T. A., & Udert, K. M. (2016a). A
novel approach for stabilizing fresh urine by calcium hydroxide addition. Water
Research, 95, 361–369. https://doi.org/10.1016/j.watres.2016.03.007
Randall, D. G., Krähenbühl, M., Köpping, I., Larsen, T. A., & Udert, K. M. (2016b). A
novel approach for stabilizing fresh urine by calcium hydroxide addition. Water
Research, 95, 361–369. https://doi.org/10.1016/j.watres.2016.03.007
Reid, W. V., Ali, M. K., & Field, C. B. (2020). The future of bioenergy. Global Change
Biology, 26(1), 274–286. https://doi.org/10.1111/gcb.14883
Ren, L. F., Deng, J., Li, Q. W., Ma, L., Zou, L., Laiwang, B., & Shu, C. M. (2019).
Low-temperature exothermic oxidation characteristics and spontaneous combustion
risk of pulverised coal. Fuel, 252(March), 238–245.
https://doi.org/10.1016/j.fuel.2019.04.108
140
Rocha, J., & Klinowski, J. (1990). 29Si and 27Al magic-angle-spinning NMR studies of
the thermal transformation of kaolinite. Physics and Chemistry of Minerals, 17(2),
179–186. https://doi.org/10.1007/BF00199671
Rose, C., Parker, A., Jefferson, B., & Cartmell, E. (2015). The characterization of feces
and urine: A review of the literature to inform advanced treatment technology.
Critical Reviews in Environmental Science and Technology, 45(17), 1827–1879.
https://doi.org/10.1080/10643389.2014.1000761
Saastamoinen, H., & Leino, T. (2019). Fuel Staging and Air Staging to Reduce Nitrogen
Emission in the CFB Combustion of Bark and Coal. Energy and Fuels, 33(6),
5732–5739. https://doi.org/10.1021/acs.energyfuels.9b00850
Saeed, A. A. H., Harun, N. Y., Bilad, M. R., Afzal, M. T., Parvez, A. M., Roslan, F. A.
S., Rahim, S. A., Vinayagam, V. D., & Afolabi, H. K. (2021). Moisture content
impact on properties of briquette produced from rice husk waste. Sustainability
(Switzerland), 13(6). https://doi.org/10.3390/su13063069
Sait, H. H., Hussain, A., Salema, A. A., & Ani, F. N. (2012). Pyrolysis and combustion
kinetics of date palm biomass using thermogravimetric analysis. Bioresource
Technology, 118, 382–389. https://doi.org/10.1016/j.biortech.2012.04.081
141
Salar-García, M. J., Ortiz-Martínez, V. M., Gajda, I., Greenman, J., Hernández-
Fernández, F. J., & Ieropoulos, I. A. (2017). Electricity production from human
urine in ceramic microbial fuel cells with alternative non-fluorinated polymer
binders for cathode construction. Separation and Purification Technology, 187,
436–442. https://doi.org/10.1016/j.seppur.2017.06.025
Saleem, J., Shahid, U. Bin, Hijab, M., Mackey, H., & McKay, G. (2019). Production and
applications of activated carbons as adsorbents from olive stones. Biomass
Conversion and Biorefinery, 9(4), 775–802. https://doi.org/10.1007/s13399-019-
00473-7
Sanchez, M. E., Otero, M., Gómez, X., & Morán, A. (2009). Thermogravimetric kinetic
analysis of the combustion of biowastes. Renewable Energy, 34(6), 1622–1627.
https://doi.org/10.1016/j.renene.2008.11.011
Sarkar, M., Banerjee, A., Pramanick, P. P., & Sarkar, A. R. (2006). Use of laterite for
the removal of fluoride from contaminated drinking water. Journal of Colloid and
Interface Science, 302(2), 432–441. https://doi.org/10.1016/j.jcis.2006.07.001
Sarkhot, D. V., Ghezzehei, T. A., & Berhe, A. A. (2013a). Effectiveness of biochar for
sorption of ammonium and phosphate from dairy effluent. Journal of
Environmental Quality, 42(5), 1545–1554. https://doi.org/10.2134/jeq2012.0482
Sarkhot, D. V., Ghezzehei, T. A., & Berhe, A. A. (2013b). Effectiveness of Biochar for
Sorption of Ammonium and Phosphate from Dairy Effluent. Journal of
Environmental Quality, 42(5), 1545–1554. https://doi.org/10.2134/jeq2012.0482
142
Sha, D., Li, Y., Zhou, X., Zhang, J., Zhang, H., & Yu, J. (2021). Influence of Volatile
Content on the Explosion Characteristics of Coal Dust. ACS Omega, 6(41), 27150–
27157. https://doi.org/10.1021/acsomega.1c03803
Shaaban, A., Se, S. M., Mitan, N. M. M., & Dimin, M. F. (2013). Characterization of
biochar derived from rubber wood sawdust through slow pyrolysis on surface
porosities and functional groups. Procedia Engineering, 68, 365–371.
https://doi.org/10.1016/j.proeng.2013.12.193
Shakya, A., & Agarwal, T. (2019). Removal of Cr(VI) from water using pineapple peel
derived biochars: Adsorption potential and re-usability assessment. Journal of
Molecular Liquids, 293, 111497. https://doi.org/10.1016/j.molliq.2019.111497
Shen, J., & Duvnjak, Z. (2005). Adsorption kinetics of cupric and cadmium ions on
corncob particles. Process Biochemistry, 40(11), 3446–3454.
https://doi.org/10.1016/j.procbio.2005.02.016
Shirai, H., Ikeda, M., & Aramaki, H. (2013). Characteristics of hydrogen sulfide
formation in pulverized coal combustion. Fuel, 114, 114–119.
https://doi.org/10.1016/j.fuel.2012.03.028
Sial, T. A., Khan, M. N., Lan, Z., Kumbhar, F., Ying, Z., Zhang, J., Sun, D., & Li, X.
(2019). Contrasting effects of banana peels waste and its biochar on greenhouse gas
emissions and soil biochemical properties. Process Safety and Environmental
Protection, 122, 366–377. https://doi.org/10.1016/j.psep.2018.10.030
Sica, M., Duta, A., Teodosiu, C., & Draghici, C. (2014). Thermodynamic and kinetic
study on ammonium removal from a synthetic water solution using ion exchange
resin. Clean Technologies and Environmental Policy, 16(2), 351–359.
https://doi.org/10.1007/s10098-013-0625-3
143
Simha, P., Lalander, C., Vinnerås, B., & Ganesapillai, M. (2017). Farmer attitudes and
perceptions to the re–use of fertiliser products from resource–oriented sanitation
systems – The case of Vellore, South India. Science of the Total Environment, 581–
582, 885–896. https://doi.org/10.1016/j.scitotenv.2017.01.044
Somorin, T., Parker, A., McAdam, E., Williams, L., Tyrrel, S., Kolios, A., & Jiang, Y.
(2020). Pyrolysis characteristics and kinetics of human faeces, simulant faeces and
wood biomass by thermogravimetry–gas chromatography–mass spectrometry
methods. Energy Reports, 6, 3230–3239.
https://doi.org/10.1016/j.egyr.2020.11.164
State, E., Polytechnic, T. F., Ekiti, A., & State, E. (2018). 2018 Volume 7 , Issue 1 : 10-
18 Physico-Chemical Properties of Lateritic Soils in Ado-Ekiti , South Western
Nigeria Abstract : Universal Journal of Environmental Research and Technology.
7(1), 10–18.
Stoops, G., & Marcelino, V. (2018). Lateritic and Bauxitic Materials. In Interpretation
of Micromorphological Features of Soils and Regoliths (pp. 691–720). Elsevier.
https://doi.org/10.1016/b978-0-444-63522-8.00024-3
Strande, L., Mariska, R., & Brdjanovic, D. (2014). Enduse of Treatment Products. In
Faecal Sludge Management: Systems Approach for Implementation and Operation.
http://www.eawag.ch/forschung/sandec/publikationen/ewm/dl/fsm_book.pdf
144
Sumaraj, & Padhye, L. P. (2017). Influence of surface chemistry of carbon materials on
their interactions with inorganic nitrogen contaminants in soil and water.
Chemosphere, 184, 532–547. https://doi.org/10.1016/j.chemosphere.2017.06.021
Supatata, N., Buates, J., & Hariyanont, P. (2013). Characterization of Fuel Briquettes
Made from Sewage Sludge Mixed with Water Hyacinth and Sewage Sludge Mixed
with Sedge. International Journal of Environmental Science and Development,
4(2), 179–181. https://doi.org/10.7763/ijesd.2013.v4.330
Tanui, J. K., Kioni, P. N., Kariuki, P. N., & Ngugi, J. M. (2018). Influence of processing
conditions on the quality of briquettes produced by recycling charcoal dust.
International Journal of Energy and Environmental Engineering, 9(3), 341–350.
https://doi.org/10.1007/s40095-018-0275-7
Tokan, A., Sambo, A. S., & Kyauta, J. S. (2014). Effects of Particle Size on the Thermal
Properties of Sawdust, Corncobs and Prosopis Africana Charcoal Briquettes.
American Journal of Engineering Research (AJER), 03(January), 369–374.
www.ajer.org
Udert, K. M., Larsen, T. A., Biebow, M., & Gujer, W. (2003). Urea hydrolysis and
precipitation dynamics in a urine-collecting system. Water Research, 37(11), 2571–
2582. https://doi.org/10.1016/S0043-1354(03)00065-4
145
Ugwu, K. (2013). Evaluation of Binders in the Production of Briquettes from Empty
Fruit Bunches of Elais Guinensis. International Journal of Renewable and
Sustainable Energy, 2(4), 176. https://doi.org/10.11648/j.ijrse.20130204.17
Urban, D. L., & Antal, M. J. (1982). Study of the kinetics of sewage sludge pyrolysis
using DSC and TGA. Fuel, 61(9), 799–806. https://doi.org/10.1016/0016-
2361(82)90306-4
Usman, A. R. A., Abduljabbar, A., Vithanage, M., Ok, Y. S., Ahmad, M., Ahmad, M.,
Elfaki, J., Abdulazeem, S. S., & Al-Wabel, M. I. (2015). Biochar production from
date palm waste: Charring temperature induced changes in composition and surface
chemistry. Journal of Analytical and Applied Pyrolysis, 115, 392–400.
https://doi.org/10.1016/j.jaap.2015.08.016
Uzun, H., Yıldız, Z., Goldfarb, J. L., & Ceylan, S. (2017). Improved prediction of higher
heating value of biomass using an artificial neural network model based on
proximate analysis. Bioresource Technology, 234, 122–130.
https://doi.org/10.1016/j.biortech.2017.03.015
Vijayaraghavan, K., Padmesh, T. V. N., Palanivelu, K., & Velan, M. (2006). Biosorption
of nickel(II) ions onto Sargassum wightii: Application of two-parameter and three-
parameter isotherm models. Journal of Hazardous Materials, 133(1–3), 304–308.
https://doi.org/10.1016/j.jhazmat.2005.10.016
Vinnerås, Björn, Nordin, A., Niwagaba, C., & Nyberg, K. (2008). Inactivation of
bacteria and viruses in human urine depending on temperature and dilution rate.
Water Research, 42(15), 4067–4074. https://doi.org/10.1016/j.watres.2008.06.014
146
Vinnerås, Bjorn, Palmquist, H., Balmér, P., & Jönsson, H. (2006). The characteristics of
household wastewater and biodegradable solid waste - A proposal for new Swedish
design values. Urban Water Journal, 3(1), 3–11.
https://doi.org/10.1080/15730620600578629
Viskari, E.-L., Grobler, G., Karimäki, K., Gorbatova, A., Vilpas, R., & Lehtoranta, S.
(2018). Nitrogen Recovery With Source Separation of Human Urine—Preliminary
Results of Its Fertiliser Potential and Use in Agriculture. Frontiers in Sustainable
Food Systems, 2. https://doi.org/10.3389/fsufs.2018.00032
Vyazovkin, S., Burnham, A. K., Criado, J. M., Pérez-Maqueda, L. A., Popescu, C., &
Sbirrazzuoli, N. (2011). ICTAC Kinetics Committee recommendations for
performing kinetic computations on thermal analysis data. Thermochimica Acta,
520(1–2), 1–19. https://doi.org/10.1016/j.tca.2011.03.034
Wang, B., Lehmann, J., Hanley, K., Hestrin, R., & Enders, A. (2015). Adsorption and
desorption of ammonium by maple wood biochar as a function of oxidation and
pH. Chemosphere. https://doi.org/10.1016/j.chemosphere.2015.05.062
Wang, C., Gu, L., Liu, X., Zhang, X., Cao, L., & Hu, X. (2016). Sorption behavior of
Cr(VI) on pineapple-peel-derived biochar and the influence of coexisting pyrene.
International Biodeterioration and Biodegradation, 111, 78–84.
https://doi.org/10.1016/j.ibiod.2016.04.029
147
Wang, J., Ye, X., Zhang, Z., Ye, Z. L., & Chen, S. (2018). Selection of cost-effective
magnesium sources for fluidized struvite crystallization. Journal of Environmental
Sciences (China), 70, 144–153. https://doi.org/10.1016/j.jes.2017.11.029
Ward, B. J., Yacob, T. W., & Montoya, L. D. (2014). Evaluation of solid fuel char
briquettes from human waste. Environmental Science and Technology, 48(16),
9852–9858. https://doi.org/10.1021/es500197h
Wei, X., Viadero, R. C., & Bhojappa, S. (2008). Phosphorus removal by acid mine
drainage sludge from secondary effluents of municipal wastewater treatment plants.
Water Research, 42(13), 3275–3284. https://doi.org/10.1016/j.watres.2008.04.005
WHO. (2006). WHO Guidelines for the Safe Use of Wasterwater Excreta and Greywater
- World Health Organization - Google Books.
https://books.google.fi/books?hl=en&lr=&id=uJJ3UIPGtFIC&oi=fnd&pg=PR9&d
q=who+2006+reuse+of+huma+excreta+standards+and+guidelines&ots=wQboTW
mibd&sig=25U5IInpqPdLtZEVDyt_kDteVYk&redir_esc=y#v=onepage&q=who
2006 reuse of huma excreta standards and guidelines
Winker, M., Vinnerås, B., Muskolus, A., Arnold, U., & Clemens, J. (2009). Fertiliser
products from new sanitation systems: Their potential values and risks. Bioresource
Technology, 100(18), 4090–4096. https://doi.org/10.1016/j.biortech.2009.03.024
Wu, D., Zhang, B., Li, C., Zhang, Z., & Kong, H. (2006). Simultaneous removal of
148
ammonium and phosphate by zeolite synthesized from fly ash as influenced by salt
treatment. Journal of Colloid and Interface Science, 304(2), 300–306.
https://doi.org/10.1016/j.jcis.2006.09.011
Wunderlich, S. (2015). Nutrient Recovery from Wastewater Using Biochar and Zeolite.
2013–2015.
Xiong, J., Qin, Y., Islam, E., Yue, M., & Wang, W. (2011). Phosphate removal from
solution using powdered freshwater mussel shells. Desalination, 276(1–3), 317–
321. https://doi.org/10.1016/j.desal.2011.03.066
Xu, K., Zhang, C., Dou, X., Ma, W., & Wang, C. (2019). Optimizing the modification of
wood waste biochar via metal oxides to remove and recover phosphate from human
urine. Environmental Geochemistry and Health, 41(4), 1767–1776.
https://doi.org/10.1007/s10653-017-9986-6
Xu, L., & Yuan, J. (2015). Online identification of the lower heating value of the coal
entering the furnace based on the boiler-side whole process models. Fuel,
161(August), 68–77. https://doi.org/10.1016/j.fuel.2015.08.009
Yabe, J., Nakayama, S. M. M., Ikenaka, Y., Yohannes, Y. B., Bortey-Sam, N., Kabalo,
A. N., Ntapisha, J., Mizukawa, H., Umemura, T., & Ishizuka, M. (2018). Lead and
cadmium excretion in feces and urine of children from polluted townships near a
lead-zinc mine in Kabwe, Zambia. Chemosphere, 202, 48–55.
https://doi.org/10.1016/j.chemosphere.2018.03.079
Yacob, T. W., (Chip) Fisher, R., Linden, K. G., & Weimer, A. W. (2018). Pyrolysis of
human feces: Gas yield analysis and kinetic modeling. Waste Management, 79,
214–222. https://doi.org/10.1016/j.wasman.2018.07.020
Yahav Spitzer, R., Mau, V., & Gross, A. (2018). Using hydrothermal carbonization for
sustainable treatment and reuse of human excreta. Journal of Cleaner Production,
149
205, 955–963. https://doi.org/10.1016/j.jclepro.2018.09.126
Yank, A., Ngadi, M., & Kok, R. (2016). Physical properties of rice husk and bran
briquettes under low pressure densification for rural applications. Biomass and
Bioenergy, 84, 22–30. https://doi.org/10.1016/j.biombioe.2015.09.015
Yao, Y., Gao, B., Chen, J., & Yang, L. (2013). Engineered biochar reclaiming phosphate
from aqueous solutions: Mechanisms and potential application as a slow-release
fertilizer. Environmental Science and Technology, 47(15), 8700–8708.
https://doi.org/10.1021/es4012977
Yao, Y., Gao, B., Chen, J., Zhang, M., Inyang, M., Li, Y., & Alva, A. (2013).
Bioresource Technology Engineered carbon ( biochar ) prepared by direct pyrolysis
of Mg-accumulated tomato tissues : Characterization and phosphate removal
potential. Bioresource Technology, 138, 8–13.
https://doi.org/10.1016/j.biortech.2013.03.057
Yao, Y., Gao, B., Zhang, M., Inyang, M., & Zimmerman, A. R. (2012). Effect of biochar
amendment on sorption and leaching of nitrate, ammonium, and phosphate in a
sandy soil. Chemosphere, 89(11), 1467–1471.
https://doi.org/10.1016/j.chemosphere.2012.06.002
Zevenhoven, R., & Kilpinen, P. (2001). Control of pollutants in flue gases and fuel
gases.
Zhai, Y., Wang, T., Zhu, Y., Peng, C., Wang, B., Li, X., Li, C., & Zeng, G. (2018).
Production of fuel pellets via hydrothermal carbonization of food waste using
molasses as a binder. Waste Management, 77, 185–194.
150
https://doi.org/10.1016/j.wasman.2018.05.022
Zhang, Gaosheng, Liu, H., Liu, R., & Qu, J. (2009). Removal of phosphate from water
by a Fe-Mn binary oxide adsorbent. Journal of Colloid and Interface Science,
335(2), 168–174. https://doi.org/10.1016/j.jcis.2009.03.019
Zhang, Guojie, Sun, Y., & Xu, Y. (2018). Review of briquette binders and briquetting
mechanism. Renewable and Sustainable Energy Reviews, 82(January 2017), 477–
487. https://doi.org/10.1016/j.rser.2017.09.072
Zhang, J., Liu, J., Evrendilek, F., Xie, W., Kuo, J., Zhang, X., & Buyukada, M. (2019).
Kinetics, thermodynamics, gas evolution and empirical optimization of cattle
manure combustion in air and oxy-fuel atmospheres. Applied Thermal Engineering,
149, 119–131. https://doi.org/10.1016/j.applthermaleng.2018.12.010
Zhang, Liang, Hong, S., He, J., Gan, F., & Ho, Y. S. (2011). Adsorption characteristic
studies of phosphorus onto laterite. Desalination and Water Treatment, 25(1–3),
98–105. https://doi.org/10.5004/dwt.2011.1871
Zhang, Ling, Wu, W., Liu, J., Zhou, Q., Luo, J., Zhang, J., & Wang, X. (2014). Removal
of phosphate from water using raw and activated laterite: Batch and column studies.
Desalination and Water Treatment, 52(4–6), 778–783.
https://doi.org/10.1080/19443994.2013.826786
Zhang, Y., Li, Z., & Mahmood, I. B. (2014). Recovery of NH 4+ by corn cob produced
biochars and its potential application as soil conditioner. Frontiers of
Environmental Science and Engineering, 8(6), 825–834.
https://doi.org/10.1007/s11783-014-0682-9
Zhang, Z., & Baixiaofeng. (2009). Comparison about the three central composite
designs with simulation. Proceedings - International Conference on Advanced
Computer Control, ICACC 2009, 3, 163–167.
151
https://doi.org/10.1109/ICACC.2009.48
Zhou, Y., Zhang, L., & Cheng, Z. (2015). Removal of organic pollutants from aqueous
solution using agricultural wastes: A review. Journal of Molecular Liquids, 212,
739–762. https://doi.org/10.1016/j.molliq.2015.10.023
Zhu, K., Fu, H., Zhang, J., Lv, X., Tang, J., & Xu, X. (2012). Studies on removal of
NH4+-N from aqueous solution by using the activated carbons derived from rice
husk. Biomass and Bioenergy, 43, 18–25.
https://doi.org/10.1016/j.biombioe.2012.04.005
Zhu, R., Zhu, L., Zhu, J., Ge, F., & Wang, T. (2009). Sorption of naphthalene and
phosphate to the CTMAB-Al13 intercalated bentonites. Journal of Hazardous
Materials, 168(2–3), 1590–1594. https://doi.org/10.1016/j.jhazmat.2009.03.057
152
APPENDICES
Appendix I: Composition and Structures of Lateritic Soil (LS) and Pineapple Peel
Biochar (PPB)
Figure A1: Nitrogen Adsorption Isotherm at 77 K of (a) Lateritic Soil, and (b)
Pineapple Peel Biochar
Figure A2: BET Plots of (a) Lateritic Soil, and (b) Pineapple Peel Biochar
Determined using a Linear Section of Fig. A1 above at Relative Pressures (p/po)
Over the Range of 0.05 – 0.3
153
Figure A3: EDS Mapping of Lateritic Soil
154
Figure A4 (a): EDS Graph of Lateritic Soil at Spectrum 1
155
Figure A4 (c): EDS Graph of Lateritic Soil at Spectrum 3
156
Figure A5: EDS Mapping of Pineapple Peel Biochar
157
Appendix II: Langmuir, Freundlich, and Dubinin-Radushkevich Isotherm and
Equilibrium Data for NH4+-N Adsorption on Lateritic Soils and Pineapple Peel
Biochar
158
Table B3: Freundlich Isotherm Data of Lateritic Soil
159
Table B6: Dubinin-Radushkevich Isotherm Data of Pineapple Peel Biochar
160
Table B9: Ammonium Nitrogen (NH4+-N) Adsorbed (mg/g) at Initial Concentration
of 398.92 mg/L
161
60 4.48±0.08 6.15±0.06
90 4.90±0.03 6.16±0.05
120 4.91±0.03 6.16±0.02
162
Appendix III: Calibration Curves, Equilibrium, and Langmuir, Freundlich, and
Dubinin-Radushkevich Isotherm Data for P Adsorption on Lateritic Soils and
Pineapple Peel Biochar
1.6 2.5
Co: 680.21 mg/L Co: 555.04 mg/L
2.0
1.2
Absorbance
Absorbance
1.5
0.8
1.0
y=0.014x+0.1657 y=0.0027x+0.1674
0.4
R2=0.9667 0.5 R2=0.9996
0.0 0.0
0 200 400 600 800 0 200 400 600 800
Concentration in mg/L Concentration in mg/L
1.6 2.0
Co:439.95 mg/L Co:309.82 mg/L
1.6
1.2
Absorbance
Absorbance
1.2
0.8
0.8
0.0 0.0
0 200 400 600 800 0 200 400 600 800
Concentration in mg/L Concentration in mg/L
2.5 2.5
Co: 283.55 mg/L Co: 164.12 mg/L
2.0 2.0
Absorbance
Absorbance
1.5 1.5
1.0 1.0
y=0.0026x+0.1253
y=0.0029x+0.0057
0.5 0.5 R2=0.9985
R2=1
0.0 0.0
0 200 400 600 800 0 200 400 600 800
Figure C1: Calibration Curves for P Adsorption at Different Initial Concentrations (Co)
163
Table C1: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 680.21 mg/L
Table C2: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 555.04 mg/L
Table C3: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 439.95 mg/L
Table C4: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 309.82 mg/L
Table C5: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 283.55 mg/L
164
Initial Conc. Contact time (min) P adsorbed by P adsorbed by
lateritic soil pineapple peel
biochar
283.55 30 4.78±0.05 7.54±0.03
60 5.68±0.03 10.06±0.02
90 7.58±0.06 9.46±0.03
120 7.58±0.02 9.61±0.03
Table C6: Phosphorous (P) Adsorbed (mg/g) at Initial Concentration of 164.12 mg/L
165
Table C9: Dubinin-Radushkevich Isotherm Data of Lateritic Soil
166
Appendix IV: Graphical Presentation of the Interaction of Contact Time,
Adsorbent Loading and Agitation Speed on Ammonium Nitrogen (NH4+-N)
Adsorption on Pineapple Peel Biochar (PPB)
167
Appendix V: Graphical Presentation of the Interaction of Contact Time, Adsorbent
Loading and Agitation Speed on Phosphorous (P) Adsorption on Lateritic Soil
168
Appendix VI: Emissions Monitored During Combustion of Charcoal and Co-
combustion of Charcoal with Briquettes
169
Appendix VII: Pictures of Equipment Used in the Study
170
Plate G3: Image of Ignited Charcoal and Briquettes in the Cook Stove
171