You are on page 1of 42

Progress in Energy and Combustion Science 85 (2021) 100915

Contents lists available at ScienceDirect

Progress in Energy and Combustion Science


journal homepage: www.elsevier.com/locate/pecs

Combustion engine applications of waste tyre pyrolytic oil


Maciej Mikulski a,b, Marta Ambrosewicz-Walacik c, Jacek Hunicz d,∗, Szymon Nitkiewicz e,f
a
School of Technology and Innovation, Energy Technology, University of Vaasa, Wolffintie 34, FI-65200 Vaasa, Finland
b
Vaasa Energy Business and Innovation Centre (VEBIC), Yliopistonranta 3, FI-65200 Vaasa, Finland
c
Independent researcher
d
Faculty of Mechanical Engineering, Lublin University of Technology, Nadbystrzycka 36, 20-618 Lublin, Poland
e
Faculty of Technical Sciences, University of Warima and Mazury, Oczapowskiego 11, 10-719 Olsztyn, Poland
f
Collegium Medicum, University of Warima and Mazury, Warszawska 30, 10-082 Olsztyn, Poland

a r t i c l e i n f o a b s t r a c t

Article history: There is abundant worldwide research into combustion engine applications for tyre pyrolysis oil (TPO).
Received 13 March 2020 However, many of these studies demonstrate conflicting or ambiguous results, so although the huge num-
Accepted 20 February 2021
ber of used tyres promises good availability for TPO, its role as fuel for transport applications is still
uncertain. This reviewś goal is to clarify the case for TPO as transport fuel by means of a critical, wide-
Key words: ranging and updated review of TPO’s engine applications. The work gathers, collates and analyses the re-
Tyres sults of over 200 influential original research papers, aiming to answer the governing research questions
Pyrolysis related to TPO production and quality, post-processing and quality improvement and its final end-use
Pyrolytic oil engine validation. The work re-evaluates the environmental aspects of TPO technology, setting it against
Life cycle analysis
the latest backdrop of growing climate change concern and the urgency to find alternative fuels. The hard
Fuel quality
economics of TPO are also addressed, for example, assessing other end-of-life tyre management routes
Fuel characterisation
Combustion engines and competing fuel alternatives.
Compression ignition The critical discussion on the key issues, including the most relevant drivers and boundaries, points
Emissions towards TPO’s use as a fuel component in marine, off-road and heavy-duty road applications. The re-
sults indicate that state-of-the-art production methods yield fuel that could be used directly in bunkering
chains for marine transport as low-sulphur fuel oil. Discussion reveals that automotive applications are
limited to blends not exceeding 10% tyre pyrolytic oil: sulphur and polyaromatic hydrocarbons contents
and particulate emissions are the main constraints. Pyrolysis process efficiency is high and feedstock for
TPO is both available and flexible. Waste tyre-derived pyrolytic oils could function as a supplementary
solution to biofuels, blended to take advantage of their complementary properties.
The particular added value of this review is that it bridges the latest knowledge from several domains
related to TPO fuel: industrial management, process chemistry, fuel science and combustion/engine re-
search. The resultant analysis is expressed in terms that are accessible to all those domains. It underlines
how studies from an individual domain perspective fail to produce the holistic view. The review creates
a route towards modern multidisciplinary research supporting TPOś role in global transition to circular
economy.
© 2021 Elsevier Ltd. All rights reserved.

Contents

1. Introduction and background. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.1. Waste tyres - the scale of the problem, recycling routes and legislation framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2. Waste tyre pyrolytic oil as a fuel for combustion engines: drivers and boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3. Motivation, knowledge gaps and detailed scope of the present work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2. The state of the art in tyre pyrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1. From waste tyres to fuel: process outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2. Mechanisms of tyre pyrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


Corresponding author.
E-mail address: j.hunicz@pollub.pl (J. Hunicz).

https://doi.org/10.1016/j.pecs.2021.100915
0360-1285/© 2021 Elsevier Ltd. All rights reserved.
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

2.3.Tyre pyrolysis technology and the effect of operational parameters on product composition . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3.1. Temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.2. Feedstock size and heating rate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.3. Inert gases and catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3.4. Feedstock type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3.5. Co-pyrolysis of tyres with other materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4. TPO production efficiency and life cycle analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4.1. TPO energy and exergy efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4.2. Tyre pyrolytic plant life cycle analysis and total cost of ownership . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4.3. TPO vs other fuels - remarks on the economics and GHG footprint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5. General characterisation of pyrolysates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.1. Solid fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.2. Gaseous fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.3. Liquid fraction (tyre pyrolysis oil – TPO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3. Pyrolytic oil as engine fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1. Properties of crude TPO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2. Properties of TPO distillates (DTPO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3. Quality improvement measures for TPO/DTPO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4. Blending with other fuels and TPO addification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4. Engine tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.1. Mixture formation and combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2. Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.2.1. Nitrogen oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.2.2. Particulate matter and smoke opacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2.3. CO and unburnt hydrocarbons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5. A critical overview and outlook on further research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Declaration of Competing Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Appendix. Type of reactors used during pyrolysis and most relevant process conditions affecting TPO yield . . . . . . . . . . . . . . . . . . . . 35
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

1. Introduction and background a synergistic effect when co-pyrolysis of bio and synthetic material
is applied, reporting potential for process efficiency enhancement
Global demand for crude oil continues to increase, propelled by and increasing the fuel mass yield [5].
economic and population growth. Meeting the Paris Agreement’s Another aspect of the sustainability challenge is efficient re-
CO2 (carbon dioxide) reduction goals requires radical changes cycling of waste materials. Management of automotive shredder
around the world. A significant reduction in fossil fuels is needed residue and management of end-of-life tyres (ELTs) are of partic-
across all fronts, from power generation to manufacturing and ular concern [6,7]. Waste tyres cannot be used in landfill due to
transport. Although renewables’ share is increasing at a rate of 4% poor biodegradability and environmental hazards like fire and wa-
per year, it is foreseen that in 2030 they can still only cover half of ter stagnation (supporting the proliferation of insects and bacte-
the global increase in energy demand [1]. ria) [8]. The complex material structure which makes ELTs difficult
Looking at fuel production, growth of gasoline consumption is to process also creates a variety of residual streams. Steel fibres,
being retarded mainly by fast progress of electrification in the shreds, oils and carbon filler can be recovered from used tyres.
light-duty transport sector. Heavy-duty transport continues to rely Tyres’ high calorific value make them suitable as fuel substitute
on diesel fuel (DF) due to its high energy density and the high in waste-to-energy plants or as a feedstock to alternative thermo-
efficiency of compression-ignition (CI) engines [2]. Here, the fast chemical conversion plants for both energy and carbon material
track towards sustainability is mainly via scaling up new gen- production [9].
eration drop-in fuels that can be readily used in existing fleets. In the above context, pyrolysis of ELTs is currently a widely ex-
Currently, there are two feasible renewable fuel production tech- plored topic, with many works focusing on improving the process
nologies that have reached industrial-scale production volumes: efficiency and tyre pyrolysis oil (TPO) fraction yield [10–12]. At the
transesterification and hydrotreatment of carboxylic acids. Trans- same time, crude TPOs and also their different distilled fractions
esterification, producing mainly fatty acid methyl esters (FAME), have been studied by many researchers as standalone fuel [13,14].
is more widespread due to the simplicity of the process. Hy- Others have studied the large-scale blending of TPO with other fos-
drotreatment, yielding hydrotreated vegetable oil (HVO) is less sil fuels [15,16] and biofuels [17], or, in case of light distillates, as
common but offers greater potential because the end product a renewable fuel additive (improver) [18]. This research involves
has higher quality than esters or even diesel [3]. Moreover, HVO both laboratory analysis of fuel properties and evaluation of per-
can be produced using existing refinery infrastructure, provid- formance by engine tests. Evaluating the current status of engine
ing high process efficiency. Both processes have one significant research is particularly difficult due to a diversity of engine plat-
downside: limited flexibility on feedstock, which heavily relies on forms used in tests. For instance, contradictory results are quite of-
first-generation bio-components, potentially competing with food ten reported when exhaust emission effects are considered [13,15].
production [4]. Similar discrepancies can be observed in terms of crude TPO pro-
Pyrolysis is another promising concept for renewable diesel and duction, fractionation and quality improvement. These ambiguities
other types of fuels. The main advantage here is that a variety of result from differences in feedstocks used, processing technology,
bio and synthetic waste materials (plastics, rubber, etc…) can be engine displacements and their technology level or the combustion
used in the same reactor. More importantly, recent works indicate modes used during testing.

2
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Nomenclature and abbreviations SiO2 silica


SOC start of combustion
Al2 O3 aluminium oxide SOI start of injection
BF blast furnace SO3 sulphur trioxide
BMEP brake mean effective pressure SOX oxides of sulphur
BR butadiene rubber TGA thermogravimetric analysis
bTDC before top dead centre TiO2 titanium dioxide
C carbon TME tallow methyl ester
CA crank angle TPO tyre pyrolysis oil
CaCO3 calcium carbonate UCOME used cooking oil methyl ester
CaO calcium oxide UHC unburnt hydrocarbons
Ca (OH)2 calcium hydroxide ZnO zinc oxide
CCAI calculated carbon aromaticity index ZSM-5 zeolite Socony Mobil-5
CH3 COOH acetic acid
CI compression ignition
CN cetane number The objective of this work is a comprehensive assessment of
CO carbon monoxide the current state of the art in ELT conversion to mechanical power
CO2 carbon dioxide through internal combustion engine applications. The introduction
CR common rail section provides a background for the present work. It describes
DF diesel fuel the overall scale of the problem and its legislation boundaries,
DTPO distilled tyre pyrolytic oil placing pyrolysis in the context of other re-processing technolo-
DMC dimethyl carbonate gies. The introduction is summarised by discussing the premise of
DTG derivative thermogravimetry the present research in terms of its significance and the particu-
EGR exhaust gas recirculation lar knowledge gaps it aims to fill. The main body of the work re-
ECA Emission Control Area views the current state of the art in tyre pyrolytic fuels, following
EEDI energy efficiency design index the process chain from production (Section 2), through analytics
ELT end-of-life tyre and upgrading (Section 3) and finishing with end-use engine per-
EPR extended produced responsibility formance (Section 4). The work is summarised with an overview
ETRMA European Tyre and Rubber Manufacturers Asso- and outlook section (Section 5), culminating with conclusions.
ciation
EU European Union 1.1. Waste tyres - the scale of the problem, recycling routes and
FAME fatty acid methyl esters legislation framework
Fe2 O3 iron oxide
H hydrogen The number of used tyres in the European Union (EU) is trend-
H2 molecular hydrogen ing upwards, mainly because of the developing road transport and
H2 O2 hydrogen peroxide light-duty automotive sectors in eastern Europe. A report from
H2 S hydrogen sulphide the European Tyre and Rubber Manufacturers’ Association (ETRMA)
H2 SO4 sulphuric acid indicated there were 2.88×109 kg of used tyres (EU, Norway,
HFO heavy fuel oil Switzerland and Turkey) in 2013 and 3.58×109 kg in 2018, which
HHV higher heating value gives an average growth rate of 4% per year [19,20] A similar trend
HRR heat release rate is observed on a worldwide basis, with a global amount of ELT
HVO hydrotreated vegetable oil generated estimated in 2019 to be 29.1×109 kg [21].
IMEP indicated mean effective pressure Currently, the management of used tyres in most countries is
IMO International Maritime Organisation regulated by legislation. A comprehensive review of specific EU
JME jatropha methyl esters regulations affecting tyre deposition and treatment can be found in
K2 O potassium oxide [9]. Of over 15 sets of ELT legislation, certainly the most important
LHV lower heating value one is the Directive on the Landfill of Waste 10 0 0/31/EC, issued in
MgO magnesium oxide 1999. This initiative basically prohibited stockpiling of whole tyres
MJ/kg megajoules per kilogram in landfills, starting from July 2003, and scrapped tyres from July
MJ/Nm3 megajoules per normal cubic metre 2006. As a result, in 2010, 96% of the EU’s used tyres were re-
N nitrogen covered for residual management [19]. The EU Circular Economy
NaOH sodium hydroxide Package of December 2015 also helped close the loop in used-tyres
NCV net calorific value management. It put particular emphasis on product reuse and sup-
NEDC new European drive cycle porting sustainability by extending the product’s lifetime. Reuse,
NiMo-Al2 O3 nickel-molybdenum on alumina export and retreading (all classified as reuse) cumulatively account
NO nitric oxide for around 20% of all recycled tyres. In the case of ELTs (tyres that
NOX oxides of nitrogen cannot be reused), ERTMA identifies material recycling and energy
NBR natural polybutadiene rubbers recovery as primary handling routes. Fig. 1 depicts how utilisation
P2 O5 phosphorus pentoxide of waste tyres evolved as a consequence of legislation over the
PAH polycyclic aromatic hydrocarbons years. It is based on merged data from sources referenced in the
PM particulate matter figureś caption.
PRR pressure rise rate Generally, the ELT market is currently stable in Europe with
rpm revolutions per minute country-to-country variation being more significant than the
SBR styrene-butadiene rubber change in the global trend. Material recycling is now the main han-
dling route (50%), followed by energy recovery (29%). The ELT man-

3
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 1. ELT management routes and their percentage contribution in handling available material in EU. Based on data merged data from [22] - until 2009; [23] – until 2015;
[24] – year 2016, [25] – year 2017, [20] – year 2018.

agement system in EU countries consists of three different models: In terms of drivers, TPO fuels face the same economics as other
Extended Producer Responsibility (EPR), the liberal system (free alternative fuels. Global fuel demand is rising by around 10% per
market) and the tax system. Under the first system, introduced year and at the same time crude oil exploration is becoming more
in the majority of EU member countries, manufacturers and im- expensive, increasing the economic feasibility for alternatives
porters of tyres are obliged to ensure that tyre wastes are dis- [1]. This momentum is reinforced by numerous CO2 reduction
posed of in an appropriate manner that does not harm the envi- initiatives, with the 2015 Paris agreement being the most relevant
ronment [19]. The free-market system assumes that ELTs are im- globally.
portant sources of valuable raw materials and that their recovery Tyre pyrolytic oils are produced from waste material and so are
and recycling are profitable, so the onus rests on commercial com- classified as second-generation fuels. As such, they have a more
panies to capture and harness the ELTs. That system operates in favourable CO2 factor on a well-to-wheel basis than popular biofu-
Austria, Switzerland and Germany. In the case of the tax system, els FAME and HVO, assuming these are still majorly produced from
applied in Denmark and Croatia, recovery/recycling organisations first-generation components (competing with food production) like
are responsible for used-tyres management. This is financed by the palm or rapeseed oil [30]. However, it is necessary to stress that a
state via a tax levied on tyre manufacturers [19]. detailed GHG emissions life cycle analysis for both fuel categories,
Depending on the country/system, different stakeholders take TPO and bio-oils, is extremely complex and heavily dependent on
primary responsibility for management of ELTs and, ultimately, for geo-economic factors (see Section 2.4 for details).
phasing-in new processing technologies. One of the main barriers The biggest challenge for FAME and HVO is limited flexibility
to application of tyres pyrolysis is unclear legal process definition. of feedstock. Data in the previous section and the percentage of
According to the EU, depolymerisation of tyres by gasification or collected ELTs [20,31] suggest that waste tyres potentially available
pyrolysis is classified as a destructive process, equivalent to in- for TPO production in 2019 could be estimated at around 2.7×109
cineration (thermal waste treatment with or without heat recov- kg per year in Europe alone, and growing at 4% per annum. This
ery). In fact, pyrolysis of ELTs should be considered as an energy is assuming the current ELTs stream split for applications other
and material co-recovery process. Pyrolysis oil (TPO) and remaining than energy. For comparison, European availability of oilseeds and
gas both can be used as energy sources or inputs for new materi- vegetable oils (which currently underpin over 86% of biofuel pro-
als synthesis [26]. Thus, pyrolysis shows considerable environmen- duction) accounts for 50×109 kg, but only around 13% of that (i.e.
tal benefits over combustion/incineration which releases significant about 6.5×109 kg) is available for fuel purposes due to strong com-
amounts of CO2 and extremely toxic substances. The former con- petition with food production [32]. Waste tyre availability for en-
tributes to climate change; the latter are dangerous for the envi- ergy generation would increase if more sustainable and econom-
ronment and human health [27,28]. Still, the emission legislation is ically feasible solutions are developed, making this use of ELTs
quite liberal for waste incineration processes, effectively inhibiting more attractive than direct material recovery. Pyrolysis (and gasifi-
the more energy-efficient and environmentally friendly processes. cation) also provide feedstock flexibility, meaning the same reactor,
In summary, current EU legislation suppresses the opportu- in principle, can be used to convert biomass, plastics and rubbers
nity for tyre pyrolysis to take a major share in ELTs management. from waste tyres. This co-pyrolysis of different feedstocks provides
To address this, closer collaboration between the joint alternative further pathways to improve TPO composition and yield [33].
management system of each country and the research community Aside from the general push to cut CO2 emissions, the light-
should be promoted. A transition from Extended Producer Respon- and heavy-duty road transport sectors have their own country-
sibility to the free-market system could offer a better opportunity specific (or EU-wide) fleet-level CO2 targets. The off-road sector is
for TPO. The ERTMA [29] and some research groups [8] are lobby- currently discussing the adoption of similar limits, while the ma-
ing for this, aiming to create a value-based drive towards more ef- rine sector has its EEDI (Energy Efficiency Design Index). Without
ficient and socially acceptable solutions in waste tyre management. going into detail, it is sufficient to say that, because all initiatives
refer to tank-to-wheel CO2 emissions, none of them supports the
1.2. Waste tyre pyrolytic oil as a fuel for combustion engines: drivers development of diesel-like alternative fuel. This applies equally to
and boundaries TPO, FAME and HVO since all have a similar carbon to hydrogen ra-
tio, dependent more on the feedstock and refining processes rather
For simplicity, this discussion concerning the drivers and than on the fuel category itself.
boundaries for TPO as engine fuel focuses purely on CI (diesel) en- The constraints for TPO as a fuel are mostly dependent on the
gines. These have the biggest demand for alternative fuels, as in- domain where the fuel is intended for use. Table 1 outlines the
dicated in Section 1. Fossil DF, and on the alternative side, FAME selected physicochemical properties of the fuels determined by the
and HVO, currently the most feasible alternative for CI engines, are most relevant standards in different application domains. Note that
mentioned in these discussions as references. the automotive and marine fuel regulations are considered the

4
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Table 1
Selected requirements for CI engine fuels in different transport domains with respective standards [34–38].

Sector Norm/standard Density Kinematic viscosity Sulphur content CN/CCAI Flash point Water content
Units / conditions / limits [kg/m3 ] @150 C [mm3 /s] @400 C [ppm] [-] min. [°C] min. [specified] max. Comments

On- CEN EN590 820-845 2-4.5 10 51 55 200 [mg/kg] EU commercial


road Diesel
CEN EN 14214 860-900 3.5-5 10 51 120 500 [mg/kg] EU neat biodiesel
ASTM D975 No. 2-D n/a 1.9-4.1 15 40 52 0.05 US commercial
[%vol.] Diesel
ASTM D6751 09 n/a 1.9-6 15 47 93 0.05 US neat biodiesel
[%vol.]
Off- DIN 51605 n/a 910-925 36 10 39 101 750 [mg/kg] german standard
road/ for SVO fuel
other applications
ASTM D975 No. 4-D n/a 5.5-24 n/a 30 55 0.05 heavy distillate for
[%vol.] const. speed eng.
Marine ISO 8127 ISO-F-DFA 890 max. 2-6 1000 40 60 n/a distillate fuels;
ISO 8127 ISO-F-DFB 900 max. 2-11 1500 35 60 n/a selected categories

ISO 8127 ISO-F-RM(A-E) 920 -991 10 -180 @50°C statutory 850-860 (CCAI) 60 0.05 residual fuels;
[%vol.] selected categories

ISO 8127 ISO-F-RMG 991 max. 180-700 @50°C statutory 870 (CCAI) 60 0.05
[%vol.]

For marine residual fuels, the customer defines the sulphur content in accordance with relevant statutory limitations.

two boundary regimes for fuel applicability (least and most in- Aside from the fuel standards outlined above, emissions legisla-
clusive, respectively) and from this perspective, we discuss them tion has an indirect impact on fuel phase-in. Currently, all domains
more thoroughly. Selected regulations for the off-road sector are where combustion engines operate are subject to strict emission
provided for reference. limits. In Europe, Euro VI is in force for light- and heavy-duty ve-
The EN 590 standard defines the fuel properties for compres- hicles, limiting emissions of nitrogen oxides (NOX ), unburnt hydro-
sion ignition engines for European road transport. In the US, the carbons (UHC), carbon monoxide (CO) and particulate matter (PM).
ASTM standard for automotive diesel fuels (No. 2-D) puts similar Stage V regulations cover the off-road sector (including inland wa-
requirements on fuel quality, although there is one major differ- terways), but impose slightly less stringent NOX , UHC and CO lim-
ence: ASTMś minimum cetane number (CN) is significantly lower its, dependent on engine power output. In the marine sector, the
than that of EN590. Generally, a well-known fact is that crude TPO IMO imposes NOX limits: Tier III regulations apply in ECAs and
complies with neither EN 590 nor ASTM D-2 as a standalone, drop- Tier II regulations apply in waters outside the ECAs. For a com-
in fuel because of its high viscosity and large amounts of impuri- plete overview of current emission norms, the reader is referred to
ties [17]. Thus, most authors typically consider blending TPO with DieselNet: Engine & Emission Technology Online [43].
DF and/or other alternatives. In terms of alternatives, note that The case of how emission legislation either supports or sup-
neat FAME as an infrastructurally mature fuel category has sep- presses individual fuel alternatives has three dimensions. One is
arate quality standards both in the EU and the US. For automo- the engine-out emissions (i), where different fuels have a different
tive fuels, these standards are quite strict and yet EN590 still limits internal predisposition to creating certain hazardous species. The
the blending of FAME in commercial diesel to 7% by volume, while second factor (ii) relates to how the fuel affects engine perfor-
ASTM D975 imposes a 5% limit. mance, since all emission limits are given in values specific to
The other commonly considered option is further refining to generated power. Finally, (iii) all of the discussed emission legis-
produce TPO more compliant with EN 590, with distillation as one lation currently in force, except IMO Tier II, require some sort of
of the more straightforward methods [39]. Poor miscibility with exhaust gas aftertreatment for CI engines to meet the imposed
diesel is another constraint if TPO is considered as an additive limits. However, TPO can have a profound effect on performance
[40]. In terms of further upgrading TPO, standard refinery infras- of catalysts used in aftertreatment. Notably, sulphur is limited in
tructure can be used to treat crude TPO with similar processes as automotive diesel fuel because it is a catalyst poison for diesel
those for fossil crude oil or vegetable oils, including hydrotreat- oxidation catalysts, catalysts for NOX reduction by urea SCR, and
ing, to arrive at ultra-clean distillates. These are covered in detail NOX adsorbed catalysts. The EU 10 ppm and the US 15 ppm sul-
in Sections 3 and 4, considering both energy demand/quality and phur limits are thus not negotiable and would have to be met for
end-use in engines. use in these applications. Section 4 of this review (Engine tests)
The marine domain is still quite inclusive in terms of fuel qual- looks in more detail at the TPO studies that evaluated legislated
ity. ISO 8217 is the current standard governing bunker fuel qual- and unlegislated emissions.
ity, and crude TPO, depending on the process and feedstock qual-
ity, can be directly considered in one of the distillate or residual 1.3. Motivation, knowledge gaps and detailed scope of the present
marine fuel categories (consult Table 1). Also, heavy fractions from work
distilled tyre pyrolysis oil (DTPO) can be used to valorise the off-
stream from lighter distillates production. High viscosity and high The above introduction clearly shows that ELTs are an attrac-
contamination do not pose a challenge, because marine diesel en- tive feedstock for energy generation. Pyrolysis as fuel production
gines have on-board fuel pretreatment using centrifugal separa- technology has potential to combine clean utilisation of tyres with
tors and preheaters. However, marineś sulphur legislation might an efficient way of material recycling and energy recovery, while
be challenging for TPO. The 2020 International Maritime Organi- avoiding the food-fuel dilemma. Furthermore, the flexibility of the
zation (IMO) sulphur cap limits a fuel’s total sulphur content to reactor permits greater feedstock availability through co-pyrolysis
0.5% on all international waters [41,42]. In the IMO’s Emission Con- of other synthetic waste (plastics) and biomass. In the face of the
trol Areas (ECAs) the fuel’s sulphur limit is only 0.1% (10 0 0 ppm) global CO2 reduction targets, this seems to be increasingly recog-
[41]. Desulphurisation methods for TPO are reviewed in detail in nised. The main hurdles for TPO to overcome relate to legislation,
Section 3.3. production technology development and the technical aspects of

5
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 2. Tyre decomposition through the pyrolysis route.

its use as an engine fuel. In all of those fields, there is currently As well as finding knowledge gaps, we recognize that there is
rapid development regarding TPO and many research works are not widespread familiarity with TPO as an alternative fuel for com-
bringing new insights, greater knowledge and more efficient solu- bustion engines. So, aiming to build a complete picture, this re-
tions. This is communicated in original research papers but there view also provides an updated description of the state-of-the-art
are also broader review works that summarise progress. Reviews process for tyre pyrolysis. The level of detail is sufficient to under-
by Torretta et al. [8] and Antoniou and Zabaniotou [9] provide good stand the dependencies and to valorise the off-streams. This is cov-
insight into legislation, feedstock availability and detail forecasts ered in Section 2, along with providing the input necessary for the
for ELTs management stream. The latter review by Antoniou and research question (i) and (ii). Consequently, the input for research
Zabaniotou [9] is by far the most comprehensive study covering questions (iii) and (iv) is gathered in Section 3. Section 4 provides
the topics of tyre waste handling through pyrolysis. It has an ex- a critical review of all aspects of the research question (v). The
tensive review of the production process, pyrolysis kinetics and its introduction section provides the necessary legislative framework
products, including TPO. This study gives an excellent overview of to conduct the review in all sections. Finally, the discussion and
the topic yet it does not cover the fuel applications in great depth summary in Section 5 combine the individual inputs into a concise
and does not reflect the current state of the art in end-user val- roadmap of TPO fuelling applications.
idation of TPO fuels through engine tests. Furthermore, it is now The work is directed to scientists and engineers working in the
seven years since the study was published, a period that has seen fuel industry and engine technology. Automotive and aviation sec-
significant development in tyre pyrolysis technology. tors are already claiming over 90% of available biofuel supplies,
In terms of the fuelling applications, the knowledge gap is par- leaving limited alternatives on the liquid fuel side for the marine
tially filled by a more recent review by Czajczyńska et al. [44]. and off-road sectors. Thus, these sectors might be particularly in-
This work is very detailed but focuses solely on the gaseous frac- terested in the work. Hence the discussion is led with both auto-
tion. Furthermore, the combustion applications are limited solely motive and bunkering fuel legislation in mind. All sectors are con-
to furnace combustion, providing heat to the pyrolysis process it- sidered when assessing engine emissions.
self. Due to higher energy density (by volume) and easier logistics,
it is the liquid fraction of pyrolysis (TPO), not the gaseous fraction, 2. The state of the art in tyre pyrolysis
that holds potential as a transport fuel.
The last decade has seen a surge in research into TPO’s use 2.1. From waste tyres to fuel: process outline
as combustion engine fuel [45]. Evaluating the current status is
very difficult due to the wide discrepancy in TPO quality, differ- Tyre pyrolysis can be described as a process of thermal decom-
ent blending strategies and the variety of engine platforms used position of solid organic compounds of ELT within limited access to
hitherto. It was this combination of rapid progress, coupled with oxygen. During the thermo-chemical reactions, a series of volatile,
the difficulty of building a complete and accurate picture from so liquid and solid components are being produced [46]. Fig. 2 shows
many disparate studies of TPO’s use for combustion engine appli- a typical tyre material decomposition route through pyrolysis.
cations, that motivated us to bridge this substantial knowledge gap. Most often the tyres’ steel components are removed before
The research questions addressed here are as follows: thermal processing and rubber parts are chopped, cleaned and
dried. Tyre pyrolysis generally consists of two stages: primary py-
rolysis and secondary cracking. During the initial phase, the vapour
i What is the current view on TPO well-to-tank environmental or volatile products are generated, consisting of a variety of hydro-
impact and economic feasibility? carbons. From the point of view of this study, the liquid fraction -
ii What is the progress in pyrolysis technology with respect to TPO - is of main interest due to its possible application as a fuel.
maximising the TPO yield and resulting fuel quality? The gaseous fraction can be reused to power the pyrolytic reactor
iii What are the available TPO-upgrading technologies and what is and make the process self-sustainable in energy terms. The solid
their effect on the end fuel quality? fraction (char) contains mainly carbon black, inorganic compounds
iv What is TPOs potential as a fuel component? and many other carbon-based solids. Char can be utilised in a va-
v How do different fuel compositions of TPO and TPO-derived fu- riety of ways including substituting carbon black in the production
els perform in the engine in terms of combustion characteris- of new tyres. Therefore, by-products of the fuel recovery can pro-
tics, efficiency and emissions? vide energy savings as well as financial savings, when the entire

6
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Seidelt et al. [52] proved that curves of a rubber DTG were char-
acterised by three temperature peaks, such as at 378°C, 458°C and
468°C. The first step was attributed to depolymerisation of the BR,
while at 468°C the decomposition of the cyclised and cross-linked
polymer residue was observed. Han et al. [53] however, conducting
a thermogravimetric mass spectrum analysis of waste tyres, iden-
tified four stages of this product’s thermal decomposition. The first
stage started below 320°C and contributed a minor reduction of
mass due to vaporisation of water and decomposition of plasticiser.
The next stage occurred at 320-400°C and related to the decompo-
sition of natural rubber. The third stage, at 400-520°C, involved the
decomposition of synthetic rubber. Very little weight loss was ob-
served during the fourth and final stage, above 520°C. The authors
associated the first three phases with the decomposition of differ-
ent polymer categories in the tyres: plasticisers (temperatures up
to 320°C), natural rubber (320-400°C), and synthetic rubber (400-
Fig. 3. Typical TGA curve for pyrolysis process, based on [50].
520°C). As a final note, it should be pointed out that in terms of
temperature range and ELT material degradation stages, there is a
process is evaluated. Notably, all material and energy recovery re- good agreement within various studies.
duces carbon footprint. For a more fundamental perspective, the TGA analysis is com-
Further processing can improve the properties of crude TPO. bined with chemical kinetics studies. Four general mechanisms are
That usually involves sulphur removal, distillation and cracking. In usually related to the decomposition of organic polymers: ran-
practice, the end fuels for engines are produced from TPO via frac- dom chain scission, end chain scission, chain stripping and cross-
tionation (fraction blending to obtain desired fuel properties) and linking [54]. Small molecules of volatile alkenes and cycloalkenes
thus are referred to as distilled tyre pyrolytic oil or DTPO [47]. are produced form the degradation of rubbers due to chain scis-
sion processes. The main products of chain alkenes are: 2-pentene,
2.2. Mechanisms of tyre pyrolysis 1,3-butadiene and isoprene. These are formed from the depoly-
merisation of natural and synthetic rubber [55]. The main cyclic
The process of thermal decomposition is most often charac- alkenes, formed from the degradation of natural and synthetic
terised by curves of the mass loss versus temperature. Thermo- rubber or the cyclisation of chain alkenes consist of: 5-methyl-
gravimetry, often called thermogravimetric analysis (TGA), is an an- 1,3-cyclopentadiene, 1-methyl-1,4-cyclohexadiene, D-limonene and
alytical technique that measures the mass over a period of time as cyclonona-1,2,6-triene [56,57]. As the reactor bed temperature in-
the temperature changes [48]. So-called derivative thermogravime- creases, alkenes and cycloalkenes undergo a series of reactions
try (DTG) functions, using the same method but with different pre- such as the Dies-Alder reaction and aromatisation, forming ben-
sentation data, are used to show the rate of decomposition. zene and its derivatives. Along with the increase of the tempera-
The tyre rubber pyrolysis process TGA curve in Fig. 3 exhibits ture, the ring-opening and recyclisation reactions yield the forma-
two main stages. The first is the slow stage, related to the decom- tion of other compounds belonging to alkenes and cycloalkenes.
position of oils, plasticisers and additives present in rubber. The Mentioned reactions mostly occur in the initial stage of the whole
second is the active pyrolysis stage which involves cracking and process. During the second stage, benzene and its derivatives un-
rapid decomposition of the other rubber components. The inter- dergo further reactions to form more aromatic compounds such as
molecular associations and weaker chemical bonds are destroyed polycyclic aromatic hydrocarbons (PAHs) [56], which are problem-
during this second stage, resulting in gas and liquid fraction sepa- atic from the point of view of exhaust emissions. The favourable
ration. Beyond a certain temperature, active pyrolysis is essentially conditions for aliphatic compounds aromatisation to PAHs are high
completed, with only minor weight loss thereafter. This interpre- temperatures and residence time [58]. This stage of pyrolysis is
tation of pyrolysis mechanisms was introduced by Williams and also associated with pyrolysis of D-limonene and styrene [56,59].
Besler [48] and commonly adopted by others, e.g. Juma et al. [49].
Leung and Wang [46] stated that the first stage’s temperature 2.3. Tyre pyrolysis technology and the effect of operational
range was between 150°C and 350°C. This is where oil, moisture, parameters on product composition
plasticisers and other additives decomposed. During the second
stage, above these temperatures but below 550°C, natural polybu- Commonly used pyrolysis technologies are distinguished by
tadiene and polybutadiene-styrene rubbers (NBR and SBR respec- the heating rate and residence time. The reactor type used can
tively) were decomposed. Using DTG function, the maximum de- be: fixed, moving, fluidised, spouted and vacuum types (refer to
composition rates were observed between 380°C and 450°C. Prad- Table 2 for detail). There are also some novel technologies, such as
han and Singh [50] conducted a TGA analysis of waste bicycle tyres microwave [60,61] and ultrasonic [62–64]. Moreover, various cat-
in a laboratory-scale batch reactor. They observed that the first alysts can be applied to improve the quality of TPO, using a pro-
stage of decomposition occurred at between 250°C and 380°C, fol- cess called catalytic pyrolysis [65–73]. Finally, it should be noted
lowed by the second stage at between 380°C and 550°C. This is de- that feedstock preprocessing can vary from the preparation of rub-
picted in Fig. 3. Above 550°C, the decomposition process was prac- ber powder (required for fast heating), through centimetres-large
tically finished. TGA analysis conducted by Berrueco et al. [51] pre- pieces, to pyrolysis of whole tyres [74,75]. Whole-tyre pyrolysis is
sented the thermal decomposition of a used car tyre (Michelin ra- achieved using special reactors designed for both intermittent and
dial X type) sliced into 20 mm × 20 mm pieces. Decomposition continuous operation [76].
started at 200°C and complete conversion of the sample had oc- The content, as well as the properties and composition of in-
curred when the temperature reached 500°C. However, the cited dividual pyrolysates fractions, may vary upon the technology and
authors isolated three stages. The first was between 200°C and conditions employed during the process. The amounts of polymer
325°C, the second between 325°C and 400°C and the third stage types present in rubber significantly influence the product fraction
was between 400°C and 500°C. A DTG analysis of tyre pyrolysis by distribution, as well as the content of different hydrocarbons such

7
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Table 2
Type of reactors used during pyrolysis and most relevant process conditions affecting TPO yield.

Type of Reactor vol. [dm3 ] Inert gas used Process description (time) TPO Yield [%] Temp.∗ [deg C] References
reactor

fluidized bed 0.9 nitrogen + air continuous 25-38 700 Lee et al. (1995) [77]
n.a. nitrogen n.a. n.a. n.a. Williams and Besler (1995) [48]
nitrogen/steam fast pyrolysis 51-65 500 Kaminsky and Mennerich (2001) [78]
steam n.a. 30-50 450 Dai et al. (2001) [79]
various over 36 600 Kaminsky et al. (2009) [80]
nitrogen continuous, fast pyrolysis 25-45 400 Raj et al. (2013) [81]
0.23 nitrogen fast pyrolysis; catalytic n.a. n.a. Zhang et al. (2014) [82]
4.1 nitrogen continuous, fast pyrolysis 26-38 400-450 Wang and Jan (2018) [83]
n.a. hydrogen intermittent (4-60 min) 42-47 375 Mastral et al. (2000) [84]
fixed bed 1.5 helium intermittent (15 min) 30-38 550 Dıéz et al. (2004) [85]
1.2 nitrogen intermittent (60 min); catalytic 35-56 430 Williams and Brindle (2002) [65]
0.4 nitrogen intermittent (60 min); catalytic 46-60 425 Kar (2011) [70]
3.5 nitrogen intermittent (30 min) 5-39 700 De Marco Rodríguez et al. (2001) [86]
2.2 nitrogen intermittent (120 min) 30-38 500 Choi et al. (2014) [87]
0.59 nitrogen intermittent (60 min) 47-56 n.a. Ucar et al. (2005) [88]
0.95 nitrogen intermittent (120 min) 30-43 500 Berrueco et al. (2005) [51]
0,004 nitrogen intermittent (15 min) 36-45 n.a. Murillo et al. (2006) [89]
2.1 nitrogen intermittent (50 min) up to 49 475 Islam et al. (2008) [90]
0.08 n.a. intermittent 20-58 n.a. Acevedo et al. (2013) [91]
1.7 nitrogen n.a. 33-41 Choi et al. (2013) [92]
6.7 nitrogen intermittent (160 min) 38-88 Al Mamun al. (2016) [93]
n.a. nitrogen intermittent (20 min) 41-53 450 Hopa et al. (2017) [94]
2.8 n.a. intermittent (16 min) 25-55 n.a. Odejobi et al. (2020) [95]
179 nitrogen intermittent 30-42 400 Aziz et al. (2018) [96]
0.5 nitrogen intermittent 45-65 400 Kumar Singh et al. (2018) [10]
moving bed 0.15 nitrogen continuous (240 min) 54 n.a. Aylón et al. (2008) [97]
n.a. nitrogen continuous 28-48 Aylón et al. (2010) [98]
Vacuum 0.1 -4 kPa vacuum intermittent 33-50 480 Zhang et al. (2008) [69]
n.a. -3 kPa vacuum intermittent 10-55 415 Roy et al. (1990) [99]
848 -13 kPa vacuum continuous 50 n.a. Benallal et al. (1995) [100]
848 -80 kPa vacuum continuous n.a. Chaala and Roy (1996) [101]
n.a. -70 kPa vacuum intermittent Roy et al. (1997) [102]
848 -10 kPa vacuum n.a. 43-56 Roy et al. (1999) [103]
n.a. n.a. 30-53 Rombaldo et al. (2008) [104]
n.a. Vihar et al. (2015) [105]
40640 intermittent 40-45 Tudu et al. (2016) [106]
not specified n.a. nitrogen intermittent (30 min) 18-56 550-575 González et al. (2001) [107]
helium n.a. n.a. n.a. Seidelt et al. (2006) [52]
nitrogen intermittent (60 min); catalytic ˜
Dung et al. (2009) [108]
nitrogen intermittent (180 min) 28-41 400 Czajczyńska et al. (2020) [109]
spouted bed 11.2 n.a. continuous (30 min) 44-55 n.a. López et al. (2010) [110]
reactor n.a. continuous (50 min) 54-58 Alvarez et al. (2017) [58]
28.3 intermittent; catalytic 40 Ayanoğlu and Yumrutaş (2016) [71]
0.16 hydrogen intermittent (90 min); catalytic 10-37 550 Ramirez-Canon et al. (2018) [111]
rotary (auger) n.a. nitrogen continuous (80 min) 39-51 n.a. Choi et al. (2016) [112]
reactor n.a. n.a. n.a. Martínez et al. (2014) [113]

Temperature for optimum TPO yield.Reactor Type

as aromatics and olefins, in gas and liquid fractions. This section peratures and takes minutes to hours. Rotary kiln and fluidised bed
gathers the state of the art in tyre pyrolysis technology and the reactors are the main types for fast processes, which are distin-
knowledge of how different pyrolysis parameters influence the fi- guished by small material particles, fast heating rates and vapour
nal product yield. The most relevant data is illustrated in Table 2. residence times no longer than a few seconds. Czajczyńska et al.
Note that a more extensive yet visually less impactful version of [44] also compared different types of reactors while differentiat-
this table is included as Appendix 1. The tabularised data is fur- ing between slow and fast pyrolysis, but with the focus on gas
ther followed by a factor-by factor discussion. The focus is on max- yield. The main conclusion was that fast pyrolysis yields more liq-
imising the yield of TPO, as this fraction is central to the present uid fraction, due to avoided secondary reactions, whereas the main
review. The previous section summarising the mechanisms of py- product of slow pyrolysis is a solid fraction. This statement is ob-
rolysis underpins the following discussion. viously a simplification since the actual yield depends heavily on
Table 2 presents an overview of generic types of reactors used process conditions. Nevertheless, it qualitatively illustrates the gen-
for ELT pyrolysis. Additionally, maximum TPO yields achieved by eral trend.
different authors and key process parameters are provided. Note Another review by Lewandowski et al. [76] confirms this the-
that there are a few comprehensive reviews of ELTs pyrolysis tech- sis by comparing results from different reactors. A more thorough
nologies. Therefore, our review covers only the most important analysis of pyrolysis product composition related to reactor type is
technological findings, focusing instead on properties of the py- depicted in Fig. 4, summarising results obtained using different re-
rolytic products and their engine application. actors, but at roughly the same process temperature of 500°C. It is
Martínez et al. [114] differentiated between slow and fast py- evident from Fig. 4 that fluidised bed reactors provide maximisa-
rolysis, which are distinguished by heating rate and vapour resi- tion of the TPO yield. Attainable TPO production efficiency for fixed
dence time. Generally, fixed-bed reactors and larger material parti- bed and mechanically mixed reactors is approximately 10% lower.
cles normally are used for slow pyrolysis, which runs at lower tem- At this point, it should be emphasised that other process param-

8
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Note that a handful of studies report TPO yield saturation at


certain temperature levels. Berrueco et al. [51] carried out pyrolysis
of waste tyres in an atmospheric-bed batch reactor at temperatures
between 300°C and 700°C. There was a lack of sensitivity in TPO
yield below 400°C but yield was observed to gradually increase as
the temperature rose from 400°C, reaching a maximum at 500°C.
Further temperature increase did not affect pyrolysates yields. Sim-
ilar results were presented by de Marco Rodríguez et al. [86], con-
ducting pyrolysis at a wide range of temperatures (30 0, 40 0, 50 0,
60 0 and 70 0°C) in a nitrogen atmosphere. The cited authors ob-
served that a significant yield increase occurred at 30 0-50 0°C and
peaked at 500°C (38.0%). The yield remained at a similar level at
temperatures above 500°C. The same tendency was noticed for the
gaseous fraction. Laresgoiti et al. [117] carried out pyrolysis in iden-
tical conditions to de Marco Rodríguez et al. [86] and presented the
same results. Laresgoiti et al. [116] observed the same tendency in
earlier work too. Aydın and İlkılıç [118] also noticed that the max-
imum TPO yield (40%) was achieved at 500°C and that it remained
at a similar level as the temperature was increased to 750°C. How-
ever, the cited authors observed no correlation between gases and
oil fractions formation, contrary to the results of previous works.
The third general trend implies that TPO yield diminishes with
Fig. 4. Product lumps in different types of reactors [76]. Reproduced with permis- increasing temperature. This was reported by Williams and Brindle
sion from Journal of Analytical and Applied Pyrolysis, Elsevier. [65], Galvagno et al. [121], Mastral et al. [122] and Aylón et al. [98].
Aylón et al. [98] conducted nitrogen atmosphere pyrolysis at 600,
eters, e.g. heating rate, feedstock, particle size etc. affect the TPO 70 0 and 80 0°C. They observed that as the temperature rose there
yield to a large extent. The reported results are obtained on large was a significant decrease in TPO yield fraction, an increase of gas
capacity, pilot-scale or model (laboratory-scale) installations. Addi- fraction and a relatively constant solid yield. The decrease of the
tionally, liquid fractions can be condensed at various temperatures oil fraction and an increase of gases were explained by the cracking
and the yield analysis can be made on a wet or dry basis. Nev- of liquids. The same tendency was found by Galvagno et al. [121],
ertheless, the following paragraphs aim to describe the individual who examined three pyrolysis temperatures: 550, 600 and 680°C.
parameter effects. The highest TPO yield (38.12%) occurred at 550°C while the lowest
(31.82%) was at 680°C. The gas fraction yield increased from 2.39%
2.3.1. Temperature at the lowest pyrolysis temperature to 10.75% at the highest.
One of the most significant factors affecting pyrolysis is the For a great number of works, presented in Table 2 the limited
temperature at which the process is performed. Therefore, the ma- range of temperature sweeps performed makes it difficult to de-
jority of the research has focused on process efficiency, with tem- termine with absolute certainty whether saturation or change in
perature being one of the most important parameters in this en- monotony determined TPO yield as pyrolysis temperature rose. Fig-
deavour. From the perspective of valorisation of pyrolysis product ure 5 visualises the discrepancies already mentioned between in-
streams, maximum process efficiency coincides with maximising dividual results. Despite these discrepancies, some qualitative con-
the quantity and the heating value of the liquid fraction. clusions can be drawn from Fig. 5 and the above paragraphs. The
The existence of a distinct extremum (maximum) in TPO yield results indicate that there is a temperature window between 400°C
with pyrolysis temperature was noted by many researchers, includ- and 800°C in which TPO production is maximised. However, based
˜
ing: Murillo et al. [89], Rombaldo et al. [104], Dung et al. [108], on different TGA curves, e.g. [50,70,123,124] the pyrolysis process
Pradhan and Singh [50] and Luo and Feng [115]. Pradhan and Singh should be completed below 550°C. This is because the pyrolysis is
[50] reported high TPO yield in a laboratory-scale batch reactor not static, so the effects of temperature on TPO yield from a given
fed with scrapped bicycle tyres in 10 mm chunks. The peak TPO feedstock can be thoroughly analysed only, if at minimum, heating
conversion (49.6%) was reported at 600°C, a considerably higher rate, heat transfer in the pyrolysed material and gas retention time
temperature than in the other referred works by Berrueco et al. are taken into consideration [125]. The first two factors determine
[51], de Marco Rodríguez et al. [86], Laresgoiti et al. [116,117] and the feedstock decomposition, while the third determines secondary
Aydın and İlkılıç [118]. The temperature dependency of TPO yield cracking of the pyrolisates, which produces gas at the expense of
was explained, following Islam et al. [90] (originally referring to TPO. Therefore, apparently contradictory effects of temperature, in
Cunliffe and Williams [119]), by strong cracking of tyre rubber dur- fact, can lead to coherent conclusions. Another important issue is
ing second-stage reactions in the mid-temperature range (refer to the location of temperature measurement. In the case of contin-
Fig. 3). The insensitivity of TPO yield to temperature beyond the uous processes, usually final temperature is provided, whereas in
maximum yield point, following González et al. [107] and Nisar the case of an intermittent process there is a wide-span tempera-
et al. [120], can be explained with the observation that the liquid ture stratification.
fraction was subsequently cracked to a gaseous fraction. This was
further attributed to the rapid extraction of generated products 2.3.2. Feedstock size and heating rate
from the autoclave. Results presented by González et al. [107], and While scoping Table A1 it becomes evident that, due to the
Dai et al. [79] also indicated that peak TPO yields were reached kinetic character of the process, the size of the tyre shred influ-
at a certain temperature (respectively, 55.6% at 550°C and approx- ences pyrolysis efficiency, and hence TPO yield. More extensively
imately 50% at 450°C). Further increases in temperature (explored meshed feed material (mesh of millimetres) promotes heat en-
range up to 70 0 and 80 0°C respectively) led to a reduction in oil trainment during the process. This is especially the case when re-
fraction. This was correlated with the increase of the gaseous frac- actor design can take advantage of the particle size – e.g. a flu-
tion yield. idised bed reactor [78]. The most suitable size of the feeding ma-

9
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

pyrolysis (refer to Table 2). The gases hinder the generation of


oxides and limit the partial pressures of substrates. They also act
as carriers of volatile products of the process, while reducing the
condensation rate of aromatics and supporting carbon-black bind-
ing. Ramirez-Canon et al. [111] and Kan et al. [127], independently
presented that better quality TPO can be obtained by the use of
hydrogen as an inert gas. Hydrogenative pyrolysis stabilises the
hydrocarbon chains by hindering the re-polymerisation process.
It is also proven to remove sulphur from the liquid fraction (see
Section 3.3 for details).
Excluding the synergistic effect of hydrogen which is reactive
in the conditions of pyrolysis, for other inert gases, the vapours
retention time (practically regulated by the gas flow rate) is the
dominant factor influencing yield of liquid and gaseous fractions.
Shorter retention times support high yield of liquid products; pro-
longed presence of vapours inside the reactor leads to reactions
promoting gaseous products [79,80].
The use of a catalyst is usually designed to enhance the qual-
ity of a particular fraction. In general, catalytic pyrolysis is per-
formed to achieve monoaromatic compounds like benzene, toluene
or xylene [128]. On the other hand, catalysts may also reduce the
quantity of undesirable components (such as polycyclic aromatic
hydrocarbons) in the final product. Typical catalysts are copper ni-
Fig. 5. The effects of temperature on liquid fraction yield. The measurement points
are adopted from [10,11,51,58,65,79,81,88,94,126] and fitted with quadratic func-
trate, ZSM-5 zeolite, magnesium oxide (MgO), calcium carbonate
tions for different types of reactors. The points surrounded with circles refer to (CaCO3 ), sodium hydroxide (NaOH), sodium carbonate (NaCO3 ), ze-
truck tyres. olite and activated alumina. The efficiency of the catalysts depends
on the size of the pores: microporous catalysts are among those
terial in terms of TPO production are shreds between 1.5 and 30 offering the highest efficiency. Such structures are, however, more
mm, while sub-millimetre particles (tyre ground) promote the pro- prone to clogging by post-pyrolysis vapours, leading to deactivation
duction of gaseous products [58,59,70,77,78,87,88,92,97]. of the catalyst [65,70,82,108,128].
Scale of production is a significant aspect to consider when Williams and Brindle [65] conducted pyrolysis using Y-zeolite
specifying the feed material structure. Note that most of the re- (CBV-400) and ZSM-5 catalysts. While examining the effect of dif-
liable data on the influence of tyre meshing on pyrolysis that leads ferent catalyst bed temperatures (40 0-60 0°C, at 50°C increments)
to the above-mentioned generalisation comes from laboratory- they found the highest TPO yield (40% and 45% for each catalyst
scale reactors operating intermittently. The optimal feed material respectively) in the temperature range of 425-475°C. Further in-
granulation may vary when scaling-up the reactor size or shift- creases in temperature increased gaseous fraction production but
ing to continuous operation [96,103,106]. Furthermore, the opti- reduced the oil fraction: it was approximately 33% at 600°C for
mum heating rate for TPO production will be largely dependent both catalysts. It is worth noting that in the cited work, pyrolysis
on the fragmentation of the feed material. Haydary et al. [123], was also conducted without the catalyst, which actually resulted in
using a screw-type flow reactor performed programmed experi- higher TPO yield (55.8% at 500°C). This trend was also confirmed
ments, supported by the heat transfer model, to find correlations by other researchers [128] and the high-level conclusion is that
between heating rate and particle size. The comparison of mate- the increase of the temperature of a catalyst bed causes increased
rial conversion rates for rubber particles from 4 mm to 8 mm production of the gaseous products, at the expense of the liquid
showed that at the pyrolysis temperature of 550°C, the time re- fraction.
quired for 90% conversion was 25 and 95 seconds, respectively.
Mkhize et al. [124] analysed the effects of the temperature and 2.3.4. Feedstock type
heating rate on TPO yields, using a fixed-bed, slow pyrolysis re- The influence of different categories of tyre feedstock has been
actor with a nitrogen atmosphere. They established for individual studied by many researchers. A handful of published comparative
temperatures (309, 350, 450, 550, 591°C) that the maximum oil analyses leads to a common conclusion that the type of feedstock
fraction yield varied greatly depending on the heating rate (0.86, (e.g. car or truck tyre) affects TPO yield, due to differences in their
5, 15, 25, 29.14°C/min). Generally, the best results (around 45% oil proportions of natural rubber and synthetic rubbers.
mass fraction yield) were achieved in medium ranges of tempera- A cohesive view of this issue can be formulated when assess-
tures and heating rates. Similarly, other authors usually claim that ing data from two works. Fig. 6 combines the results of Ucar et al.
heating rate is a relevant control parameter yet fail to deliver clean [88] and Kumar Singh et al. [10]. Both studies used different cat-
sweeps of heating rate undisturbed by variations of other parame- egories of waste tyres at similar pyrolysis temperature ranges, but
ters. Thus, we believe currently it is impossible to draw any mean- with different reactors: fixed-bed and stirred tank respectively. In-
ingful conclusions on the effect of heating rate, so we leave this as terestingly, both studies show consistent trends in terms of TPO
one of the knowledge gaps for further works to cover. yield and feedstock categories. Independently of process conditions
and reactor type, heavy-duty tyres yielded significantly higher liq-
2.3.3. Inert gases and catalysts uid fraction compared to the light-duty tyre feedstock. The latter
The atmosphere (the injection of gases or vapour) is an impor- yields high amounts of char. Material analyses of the feedstock
tant variable during pyrolysis. The flow of the inert gas stabilises composition consistently showed that truck tyres have a higher
and equalises the temperature inside the reactor. In the case content of natural rubber and a lower fraction of synthetic rub-
of the fluidised-bed reactor, this also aims to provide a more ber when compared to passenger car tyres. Additionally, Ucar et al.
homogeneous temperature distribution in the bed. Nitrogen and [88] pointed out that truck tyres used SBR as a synthetic compo-
helium are frequently used to replace atmospheric air during nent, whereas car tyres contained mainly BR.

10
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 6. Product distributions from pyrolysis of scrap tyres; data elaborated based on Ucar et al. [88] and Kumar Singh et al. [10]. The exact numbers can slightly differ than
presented in original research papers, because the data were converted to the dry basis.

Kumar Singh et al. [10] used TGA to study the interaction of


different categories of waste tyres during pyrolysis. Higher frac-
tions of natural rubber (predominant in heavy-duty vehicle tyres)
was reported to reduce the degradation temperature, while syn-
thetic rubber from car tyres had an opposite effect. Importantly, a
synergistic effect was noted from interactions of both rubber frac-
tions, so solid and gaseous fraction yields were reported respec-
tively higher and lower for co-pyrolysis than predicted from indi-
vidual fractions pyrolysis data. However, the TPO yield showed a
much lower synergistic effect. More information on co-pyrolysis of
waste tyres with other materials follows in the next subsection.

2.3.5. Co-pyrolysis of tyres with other materials


As mentioned in the introduction, one particular benefit of the
pyrolysis process is its flexibility on feedstock. Some aspects of co-
pyrolysis of waste tyres with other materials deserve particular at-
tention because there seems to be an added value of material in-
teraction for improving TPO composition and yield.
Almost all available studies of co-pyrolysis focus on biomass as
the primary feedstock, and examine the large oxygenates content
Fig. 7. Sankey diagram for tyre pyrolysis at 500°C. Averaged values for HHV and
of bio-oils, leading to corrosion problems, the low calorific value of yields based on data gathered in Section 2.
the end fuel and its instability [129]. The co-pyrolysis of biomass
and low oxygen content / high calorific value polymers provides tions of 4:6 with over 30% reduction in activation energy with re-
effective mitigation to the above issue [113]. spect to the 100% biomass reference.
The publications to date in this area focus on the co-processing Bearing in mind the lack of studies concerning co-pyrolysis
of biomass and plastics: there are fewer studies of co-pyrolysis of with waste tyre as the primary feedstock and the promising ini-
biomass and scrap tyres. Abnisa and Wan Daud [130] studied co- tial results by Chen et al. [5], it appears that TPO production has
pyrolysis of palm shells and scrap tyres, changing the process tem- still some uncharted improvement potential.
perature, time and mixing ratio. They showed that co-processing
can improve the yield and quality of crude oil while compensat- 2.4. TPO production efficiency and life cycle analysis
ing for the shortages of biomass in fuel production. Farooq et al.
[131] conducted experiments with wheat straw and waste tyres 2.4.1. TPO energy and exergy efficiency
and found that, compared to oil derived from neat biomass, the The energy density of waste tyres is estimated at around 31.4
co-pyrolysis product was more stable with significantly lower alde- MJ/kg, which is roughly double that of mixed biomass [40,132].
hyde content. Chen et al. [5] were the first to explain this syn- In terms of energy efficiency it is sufficient to say that when
ergistic effects by investigating the kinetics of biomass (tobacco optimised for oil production, tyre pyrolysis can be self-sustainable,
stalk) and scrap tyre co-pyrolysis. According to the authors, the with only around 9.5% of total tyre energy used to power the
alkyl groups largely present amongst tyre decomposition products process and with only around 2% of total energy released to
can interact with carbonyl groups in tobacco stalks and lower the the ambient surroundings [133]. The rest of the products can be
CO2 production of the pyrolysis. This increases the oil fraction yield properly valorised. The process Sankey diagram (Fig. 7) show-
and lowers the energy input of the whole process. ing the energy flow uses averaged heating values and yields of
Ultimately, the study of Chen et al. [5] is the only one that the individual fractions, garnered from the detailed review in
brings forward the primary role of waste tyres in the co-pyrolysis. Sections 2.3 and 2.5. Fig. 7 is considered self-explanatory and left
Contrary to earlier mentioned studies [130] and [131], the authors without detailed comments.
explored tyre/biomass ratios above 1:1. They investigated seven More recent studies related to TPO production efficiency are
different feedstock compositions, equally distributed between the based on exergy analysis, which quantifies the process capability
100% tobacco stalk and 100% tyre scrap references. Importantly for to recover the energy and its losses related to irreversibility. Al-
the prospect of waste tyre as feedstock, the authors reported the tayeb et al. [134] calculated the exergetic efficiency of the pyroly-
greatest synergistic benefits for co-pyrolysis (tyre/biomass) frac- sis process, based on a detail analysis of chemical exergy of sub-

11
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

trates and products. The overall exergetic efficiency of the rotary


kiln reactor was found to be 69.9%, which is equivalent to an ex-
ergy destruction of 30.1% due to process irreversibility. Notably, the
authors indicated there is almost no further potential for efficiency
improvement with change of reactor or process conditions.
The obtained exergy efficiency is comparable to the value com-
monly reported for gasification plants (roughly 70%) [135]. For
reference, the exergy efficiency of transesterification to obtain
biodiesel from different feedstock is estimated to be significantly
higher, at around 80% [136]. Note, however, that in the works re-
lated to FAME, exergy losses of methanol production are not in-
cluded, so the number does not represent the complete feedstock-
product chain as with ELTs.

2.4.2. Tyre pyrolytic plant life cycle analysis and total cost of
ownership
The mentioned work of Altayeb et al. [134] is ultimately one
of the few attempting to cover the complete life cycle analysis Fig. 8. Comparison of net benefits (10 years depreciation period and total impact)
(LCA) of TPO production. This study, based on the simulated rotary for the four ELT treatment technologies in China. Merged data from Li et al. [138].
kiln reactor adjusted for optimal TPO yield temperature (∼500°C),
was made using dedicated software (SimaPro) and taking the geo-
economical perspective of Saudi Arabia. The system boundaries of pyrolysis with other tyre processing technologies and TPO with
included waste tyre transport, shredding, pyrolysis, TPO recov- other fuel alternatives. We have not found any reliable studies
ery, distribution and end-use. The analysis also took into account comparing TPO LCA with that of other alternative fuels. Only
the emissions and costs related to the transport of valorised off- Li et al. [138] have assessed pyrolysis together with other ELT
streams (steel and char). The authors analysed a set of scenar- treatment technologies. This Chinese case study is ultimately by
ios in diversifying the feedstock-product routes and reactor pow- far the most thorough and accurate analysis available on TPO
ering scenarios, distinguishing over 20 environmental impact cat- LCA. It accommodates similar system boundaries as the study of
egories while diversifying the contribution of individual TPO pro- Buadit et al. [137], from feedstock to products, excluding transport,
duction stages. However, it did not perform a comparative analysis and accounts for material recovery from avoided products. For
with other TPO processing technologies. Nor did it provide any ref- example, 350 kg of crude TPO yielding from 10 0 0 kg of scrap
erence to other alternative fuel production processes. Crucially, it tyres is assumed to substitute 175 kg of diesel fuel. Under the
presented results in a normalised form instead of revealing abso- same boundaries, LCA extended over economic analysis is applied
lute values of the individual outcomes. These are important short- to four processes currently used in China to process waste tyres.
comings in the context of the ongoing GHG-impact discussion of These four processes are ambient grinding (producing crumb
individual fuel/waste management options. Nevertheless, there is rubber to be used mainly in construction); active de-vulcanisation
one conclusion that is valuable to note. In the most feasible sce- (currently the dominant processing technology in China); py-
nario, where all non-condensable gases are used to power the re- rolysis; and Illegal oil extraction. The last process is analogous
actor (as in the diagram in Fig. 7), the total normalised GHG impact to pyrolysis but performed without proper equipment, emission
(CO2 and CH4 accounted) of TPO lifecycle was negligible compared control and without recovering the off-gases and other substances,
to toxic emissions. Toxic emissions in the study were divided into which are instead illegally dumped. This illegal practice with its
several impact categories related to air/water/terrestrial toxicity high environmental impact currently handles over 14% of ELTs in
but all relating to CO, NOX , PM and waste metals were combined. China. The end results of the LCA and environmental–economic
The highest-emitting process here was the combustion of the non- analysis performed by Li et al. [138] are depicted in Fig. 8.
condensable gases to power the reactor. These formed over 70% According to these results, pyrolysis is the best ELT processing
of the total TPO toxic emission-related impact and ultimately over technology in terms of cumulative GHG reduction and second-best
65% of the total environmental impact of TPO production. Next was in terms of fossil fuel substitution. In the second category, it is sur-
TPO end-combustion, accounting for roughly 20% and 17% impact passed by devulcanization, the main route for ELTs in China. Py-
on toxicity and total environmental impact respectively. rolysis, however, does not pose the same excessive impact on hu-
A similar study, using similar methodology was carried from a man health as devulcanisation. However, it is necessary to point
perspective of Thailand by Buadit et al. [137]. This uses the same out at that at the time of the study (2010), the problem of air-
software for LCA with similar impacts yet excludes transport of borne pollution in China was regarded as more pronounced than
materials and final combustion of TPO, focusing purely on the py- the impact of GHG. This was recognized by Li et al. [138] because
rolysis process itself. This study, however, evaluates the environ- the weighting factor of all respiratory categories was double com-
mental impacts of waste tyre pyrolysis both with and without ac- pared to climate change. If applying the same approach with the
counting for material recovery from avoided products. Under the current constraints that attach greater value to CO2 reduction, the
second assumption, on a problem-oriented level, three impact cat- weighted eco-indicators points in Fig. 8 would further reinforce the
egories - global warming, terrestrial ecotoxicity and fossil resource position of tyre pyrolysis as environmentally feasible technology.
scarcity - result in negative impacts. At a damage level, similar to Still, in the framework of the discussed study, according to Fig. 8,
Altayeb et al. [134], human health emerges as the dominant cate- pyrolysis scores best in overall environmental impact and at the
gory, resulting mainly from fine particulate matter formation from same time is considered the most economically feasible ELT man-
the combustion of the gaseous fraction that powers the reactor. agement technology. Over a 10-year depreciation period, the net
As one can note from the above two paragraphs, the LCA benefit of a new pyrolysis plant in China (2500 tonnes/year ca-
results heavily depend on the assumptions that are made and pacity) would be 32% higher compared to the current devulcani-
the geo-socio-economical perspective. Thus, the most relevant sation facility with similar capacity. Note that in this study the as-
arguments for TPO feasibility come from a comparative analysis sumed plant capacities are rather small because they are intended

12
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Table 3
The current market price of most common fuel alternatives including TPO. Prices exclude value added or fuel related taxes and are averaged over the period indicated in the
date column.

Fuel category Price [$/1000 kg] Remarks Date Reference

RME 1049 Rapeseed methyl ester (RME) according to RED scheme with Nov. Greenea (Hillairet and
quality EN 14214, water 350 ppm, CFPP 11°C, sulphur max 12 2020 Madiot [140])
ppm, FOB ARA
TME 1040 Tallow methyl ester (TME) according to RED scheme with quality
EN 14214, water 350 ppm, CFPP 11°C, sulphur max 12 ppm, FOB
ARA
UCOME 1198 Used cooking oil methyl ester (UCOME) according to RED
schemes with GHG saving >90% with quality EN 14214, water
350 ppm, CFPP 0°C FOB ARAD
UCO 730 Used cooking oil (UCO) according RED schemes with FFA max 5%,
Water & Impurities max 2%, Sulfur max 40 ppm, IV min 70-80
Tallow cat. 1 520 Animal fat cat. 1 produced under the terms of the Regulation
(EC) No. 1069/2009 – method 1) with FFA max 20%, impurities
0.15%, water max 0.6%
SVO rapeseed 950 Straight vegetable oil for fuel applications according to DIN Nov. NESTE [141]
51605 European Rapeseed oil reference (Crude Dutch /EU) 2020
SVO palm 730 Straight vegetable oil - Malaysian palm oil / Rotterdam (NL)
TPO 350 No quality standards exist, price depends on the quality and Mar. 2019 DOING Holdings [142]
varies from country to country (Americas 342-404 $/1000 kg; EU
326-349 $/1000 kg; Asia 355-379 $/1000 kg)
GASOIL 420 European low sulfur gasoil, large market variations ± 47%; Nov. 2020 Focus Econ. [143]

for decentralised applications. The only relevant data related to the


economical feasibility of large tyre pyrolysis plants (100 tonnes/h
capacity) come from 1999. This legacy work of Shelley and El-
Halwagi [139] assessing tyre pyrolysis plant economics in UK con-
ditions suggested, for an optimised scenario, a total investment
cost of 18 million euros, with a fairly good return of investment
rate of 18%. The authors, however, assumed hydrotreatment of
crude TPO as part of the process, revalorising ready fuel compo-
nents. The cited studies are therefore not comparable, neither in
time nor socio-geographical conditions. Consequently, we consider
the current state of the art in pyrolysis economics is still uncertain.
As a final note on tyre pyrolysis economics, we must mention
that while performing this review we have come across several re-
cent publications on this topic which showed severe errors in their
principal assumptions. These include, for example, overestimating
TPO value by an order of magnitude, thus making claimed con-
clusions unfeasible. These studies were deliberately excluded from
Fig. 9. Comparison of specific CO2 emissions from the production of different fuel
this review in order to avoid their corrupting influence. We will alternatives. Biodiesel values reproduced from [144]. Value of TPO calculated based
return to the bigger issue of reliability of the information on TPO on the assumptions discussed in the text.
in Section 5.

2.4.3. TPO vs other fuels - remarks on the economics and GHG first-generation biofuels. This figure does not assume any product
footprint valorisation of removed CO2 due to product substitution.
In the light of the evident lack of comparative LCA of TPO Tyres as waste are commonly counted as zero-CO2 feed-
with other fuel alternatives, we conclude this section with a brief, stock. The CO2 emission for TPO accounts mainly for burning
absolute value-based discussion of selected GHG and economic the non-condensable gases to power the reactor (including de-
factors directly relevant to the issue of fuel feasibility. To this moisturisation of the feedstock). This is commonly reported for
end, Table 3 collates available information about current market tyre pyrolysis at a level of 68.06 (kg/tonne waste tyres) [137,145].
prices of different fuel alternatives. It is not the goal of the present Energy consumption and so hence also the CO2 footprint related
work to discuss the global trends in fuel markets: the main point to ELT transport, pre-processing (grinding, mechanical separation)
from Table 3 is that TPO currently has the lowest market price and end product valorisation (extraction, transportation to end-
of all available options. The price of TPO (excluding value-added users) is usually assumed at a level of 15-19% of the pyrolysis
tax or fuel duty) is comparable with that of fossil gasoil and is process, depending on scenario [134,137,139]. Assuming TPO yields
over 50% lower than the cheapest large–scale, straight vegetable reported for the state-of-the-art reactors at the level of 45-55%
oil, palm oil. This, coupled with the fact that unlike all other bio (refer to Table 2) and lower heating values of 40 – 42 MJ/ kg (refer
components, TPOś price did not significantly change during the to Table 6), the final TPO well-to-tank emissions are on average
past 10 years, shows there is still only minor interest in the fuel 3.76 g CO2 e/MJ.
industry for using this as a fuel component for further refining. In the case of SVOs and FAME, assuming they are still produced
On the upside, there is potential to accommodate TPO directly as mainly from first-generation feedstock, it is large CO2 footprint of
engine fuel in transport sectors (such as marine) which cannot farming activities (mainly fertilisers) that is responsible for the vast
afford expensive fuels. amount of their well-to-tank CO2 emissions. This is not the case for
A comparison of the direct CO2 footprints of different alterna- biofuels derived from waste, non-edible by-products (see Table 3 –
tive fuels (Fig. 9) reveals another incentive for TPO compared to tallow cat 1) or waste cooking oil. Note that tallow biodiesel,

13
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

however, is characterised by its large energy demand to extract oil yield generally increases with temperature [126]. González et al.
from fat, entailing boiling and mechanical separation [144]. [107] pointed out that production of hydrogen increases 10 times,
As shown in Fig. 9, TPO production has roughly the same car- from 0.57 g/kg of scrap tyre to 5.3 g/kg, with the increase of
bon footprint as oil extraction for rapeseed and soy (3.12 and 4.04 temperature from 400°C to 700°C. Across the same temperature
g CO2 e/MJ respectively) and is comfortably superior to any first- range, production of methane went from 2 g/kg to 79 g/kg. Pro-
generation FAME or SVO in the overall balance. Note that CO2 duction of ethylene and ethane rose from 0 to 128 g/kg and 31
emissions of soy and canola SVOs can be derived directly from g/kg respectively. Production of CO and CO2 was not differentiated
Fig. 9 by neglecting the biodiesel combustion contribution. Further, by temperature to any great extent. The increase in hydrogen and
note that the currently most economically viable palm oil (Table 3) hydrocarbons production at nearly constant CO and CO2 yield
is considered the worst in terms of CO2 footprint [146]. resulted in higher gas calorific value, rising from 12 MJ/Nm3 , (the
As a final point, it is apparent that the price of all the men- value of pure CO) to 42.9 MJ/Nm3 , which is typical for natural gas.
tioned bio components relates to their opportunity to provide GHG Analysis of Table 4 reveals that different process parameters
reduction. This leads to the current bizarre situation in which used provide significantly different pyrolytic gas compositions for sim-
cooking oils (UCO) and their esters (UCOME) have greater market ilar temperatures. This can be attributed to different feedstocks,
value than their un-used originates. There is definitely a value in reactor types, processing times and heating rates. What is more
the green image for fuel companies. meaningful is that the different studies produced different trends.
Having said that, we acknowledge that a detailed GHG emis- For example, Galvagno et al. [121] observed a reduction of hy-
sions life cycle analysis for both fuel categories, TPO and bio-oils, drogen content with temperature, whereas the fraction of CH4 in-
is extremely complex and heavily dependent on geo-economic fac- creased. A similar trend in H2 concentration for the same temper-
tors or even fundamental assumptions. Consequently, the results ature range was observed by Berrueco et al. [51]. However, Luo
presented in Fig. 9 should be treated as indicative. Qualitatively, and Feng [115] reported an increase of H2 content with a paral-
however the differences are evident, and in the context of the lel decrease in CH4 . It is also worth to notice that Luo and Feng
above discussion, TPO can be considered more favourable in terms achieved H2 concentrations roughly twice as high as other studies,
of CO2 footprint than all first-generation biofuels, and is ultimately attributed to the use of a catalyst. This research also extended its
the cheapest alternative. range of process temperatures higher than other studies.
Dai et al. [79] systematically analysed the effects of temperature
2.5. General characterisation of pyrolysates and residence time on the composition of pyrolytic gas. The tests
were realised in a temperature range from 350°C to 800°C. For
2.5.1. Solid fraction the given temperature range, the fraction of hydrogen increased
The solid fraction is made up of hydrocarbons with more than monotonically from 1% to 15% and the fraction of CH4 increased
10 carbon atoms and consists of aromatics, non-aromatics and in- from 7% to 40%. The fraction of CO increased with temperature
gredients containing sulphur, nitrogen and oxygen [147]. Analysing from 1% to 7%, whereas the fraction of CO2 decreased from 8% to
the elemental composition of solid fraction, de Marco Rodríguez 2%. In general, concentrations of unsaturated hydrocarbons were
et al. [86] found that it was predominantly carbon, accounting found to be higher than that of saturated, which was attributed
for approximately 83% by weight. The share of other elements to double carbon bonds within the rubber molecules. Additionally,
was significantly lower. Hydrogen was 2.4%wt. at 400°C and ap- the rise in temperature increased ethylene concentration and de-
prox. 0.6%wt. at 50 0, 60 0, 70 0°C. Nitrogen was approx. 0.3%wt., creased ethane concentration, in the same range for both compo-
sulphur was approx. 2.4%wt. Oxygen and others accounted for nents. Analysing the effects of residence time, it was found that,
approximately 1%wt. The cited authors also found that the ob- at 600°C, an increase of residence time from one second to five
tained pyrolytic residues were characterised by the concentration seconds increased the content of H2 , CH4 and C2 H4 . At the same
of ash (from 9.0 to 13.2%wt.), far higher than for commercial car- time, concentrations of CO2 and saturated hydrocarbons decreased.
bon black, which has an admissible concentration of no more than This sensitivity to residence time was attributed to secondary re-
0.5%wt. The ash in the pyrolytic residues comes from the original actions, including char reduction, tar cracking and shift reaction.
tyre inorganic fillers. López et al. [148] presented that tyre rub- Similarly, Leung et al. [126] noticed that concentrations of hydro-
ber ashes consisted of SiO2 (52%wt.), ZnO (27%wt.), Na2 O (8.6%wt.), gen, methane and ethylene increase with residence time, followed
CaO (4.6%wt.), Al2 O3 (2.8%wt.). The concentrations of MgO, P2 O5 , by the reduction of heavier hydrocarbons.
K2 O, Fe2 O3 and TiO2 were between 0.4%wt. and 0.9%wt. It should be emphasised that the obtained gaseous fraction is
The solid fraction obtained by tyre pyrolysis can be used as characterised by a good calorific value (Table 4). Theoretically, af-
a solid fuel [149], a carbon black for rubber manufacturing [86], ter desulphurisation, it could be used for combustion engine ap-
a filler in road bitumen [150–152] or an adsorbent material. For plications, especially as a complement to natural gas or liquefied
the latter application, however, the carbon residue requires activa- petroleum gas in a gas network, providing large- scale availabil-
tion, which may be accomplished by either physical or chemical ity and infrastructure compatibility. It is customary for the gaseous
methods [9,153–155]. Physical activation is conducted by partial fraction to provide internal heat for the pyrolysis process itself. The
gasification using steam or CO2 as activating agents [153,154,156]. pyrolytic off-gas may cover the full energy needs of a pyrolysis
Chemical activation entails treating the char with nitric acid, hy- plant optimised for liquid yield [9]. Bearing that in mind, combus-
drogen peroxide, ammonium persulfate or impregnated by potas- tion engine applications for pyrolytic gases will not be considered
sium hydroxide and phosphoric acid: these act as a dehydrated ac- further in this work.
tivating agent [9,155].
2.5.3. Liquid fraction (tyre pyrolysis oil – TPO)
2.5.2. Gaseous fraction The oil fraction obtained as a result of tyre pyrolysis (TPO) is
The gaseous compounds include H2 , CO, CO2 , CH4 , composed of hydrocarbons between C5 and C20 and is generally
C2 H6 , C2 H4 and slight amounts of higher hydrocarbons a mixture of alkanes, cycloalkanes, alkenes, cycloalkenes and aro-
[51,65,79,107,115,121,126,148,157,158]. Table 4 provides a compre- matic compounds [80,89,103,107,108,115,147]. The exact composi-
hensive review of gas fraction composition for different pyrolysis tion depends on the type of feedstock and parameters of the pro-
temperatures. Note that the various studies use different feedstock cess. Fig. 10 compares liquid composition and is compiled from dif-
or types of reactors. At this point, it is worth noting that gas ferent studies (non-catalytic pyrolysis only). It can be noted that

14
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Table 4
Basic characterisation of pyrolytic gases fraction

Temp. Basic gases fraction composition [%vol.] Calorific value


References [°C] [MJ/Nm3 ]
H2 CO CO2 CH4 C2 H 6 C2 H 4 C3 H 8 C4 H 8 H2 S

de Marco Rodríguez et al. [86] 400 n.a. 4.2 10.7 4.4 4.5 4.3 n.a. n.a. 2.6 84.0
(2001) 500 4.8 9.4 19.8 9.1 9.4 5.1 75.5
600 6.5 8.8 20 9 9.7 3.6 75.3
700 10.4 11.4 20.6 8.1 8.9 3.9 68.5
Galvagno et al. (2002) [121] 550 20.5 0.42 8.8 9.2 7.1 6.7 n.a. n.a. 3.4 22.1
600 17.2 1.1 8.6 13.6 8.2 8.2 2.6 24
680 10.5 0.1 6.4 22.8 9.5 12.3 2.05 29.1
Leung et al. (2002) [126] 800 20.7 2.6 1.8 44.5 4.4 17.3 n.a. n.a. n.a. 38.1
Kyari and Williams (2005) [157] 500 21.5 5.1 26.2 17.3 8.2 8.7 n.a. n.a. n.a. 30.3
Berrueco et al. (2005) [51] 400 2.61 1.08 1.41 1.02 0.43 0.25 n.a. n.a. n.a. n.a.
500 14.17 0.25 1.01 4.29 1.54 0.66
550 17.86 0.26 0.78 5.64 1.75 0.83
700 10.1 0.38 1.18 5.43 1.56 1.15
Luo and Feng (2017) [115] 600 19.39 18.63 17.03 18.25 7.36 n.a. n.a. 1.79 n.a. n.a.
700 22.07 17.78 13.76 14.73 11.8 1.86
800 28.18 16.02 10.26 12.69 13.54 1.41
900 32.24 10.57 8.84 17.01 12.81 1.8
1000 39.06 8.11 9.5 9.32 17.03 1
López et al. (2011) [148] 550 22.27 1.98 2.54 21.32 4.19 2.32 n.a. n.a. n.a. 68.7
Fernández et al. (2012) [158] 550 7.9 9 n.a. 9.9 5.9 n.a. n.a. n.a. 2.9 65.6
Kumar Sigh et al. (2018) [10] 650 18.4 4.6 3.1 11.7 18.9 4.6 6.7 n.a. n.a. n.a.
750 28.7 4.2 1.2 7 3.6 0.3 1.2

Fig. 10. Composition of TPO from different reactors at variable temperatures. Data from de Marco Rodríguez et al. [86], Ucar et al. [88], Debek
˛ and Walendziewski [74] and
Alvarez et al. [58]. Two-colour filling for the first five bars indicates aliphatic hydrocarbons cumulatively.

substantially different TPO compositions can be obtained, depend- methylethenyl-benzene and styrene. These all had concentrations
ing upon the reactor type, conditions and feedstock. It is also above 0.5%wt.
evident that the effect of temperature depends on reactor type. Ucar et al. [88] compared the composition of TPO derived from
As a side note, the list of identified species and their classifica- passenger car and truck tyres, processed under identical condi-
tion into particular hydrocarbon groups can differ from study to tions. They found that the TPO profile from cars is shaped in the
study. following order: paraffins < aromatics < olefins. The TPO profile
Debek
˛ and Walendziewski [74,159] performed pyrolysis at the from heavy-duty truck tyres showed: paraffins < olefins < aro-
relatively low temperature of 400°C. The TPO produced, on a mass matics. These effects were attributed to different rubber compo-
basis, consisted of approximately 11.9% alkanes, 7.7% dienes, 4.5% sition. The higher content of natural rubber used in truck tyres
cycloalkanes, 3.7% cycloalkenes and 47.4% aromatic hydrocarbons. produced more paraffinic hydrocarbons, whereas synthetic rub-
Arabiourrutia et al. [147] found that raising the pyrolysis tempera- ber present as the main component of car tyres produced more
ture from 425°C to 610°C increased the share of aromatics almost aromatics [10,88]. Interestingly, increasing the pyrolysis temper-
three-fold. The aromatics content (unfavourable for engine appli- ature did not affect the share of the individual types of com-
cations due to their contribution to in-cylinder soot production) pounds. The same conclusion was drawn by Murillo et al. [89],
reached 16.44% of the total TPO mass at the highest pyrolysis tem- who carried out pyrolysis at 400°C, 500°C and 600°C. They
perature. But even that was three times less than aromatics con- found that the share of aromatic compounds was approximately
tent obtained by Debek˛ [159]. Simultaneously, the non-aromatic 73%, with the saturated ones amongst them being approximately
compounds (C5 – C10 ) decreased from 30.48%wt. at 425°C to 6-7% and polar aromatics contributing 20% at all temperature
23.7%wt. at 610°C. This increase in aromatics was associated with conditions.
a reduction in liquid fraction yield, which was mainly due to the Research by de Marco Rodríguez et al. [86], on the other hand,
reduction in C10 lump. Furthermore, the cited authors noted that proved a strong thermal dependency for TPO composition. Simi-
the predominant aromatic compounds presented in the TPO frac- lar findings were later mentioned by Arabiourrutia et al. [147] and
tion obtained at 425°C were 1-methyl-4-(1-methylethnyl)-benzene, Choi et al. [59]. The original study by de Marco Rodríguez et al.
1-methyl-4-(1-methylethyl)-benzene, 1,2,3-trimethyl-benzene, 1- [86] observed that the share of aromatic compounds gradually in-

15
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Table 5
Elementary analysis of TPO and reference diesel engine fuels

TPO elemental composition [%wt.]


References Temp. [°C] Experimental conditions
C H N S O + others

Luo and Feng (2017) [115] 600 pyrolysing using BF 78.65 16.97 0.39 1.71 2.37
700 (blast-furance) slag (B/T 80.55 15.45 0.58 1.28 3.14
800 ratio 1:1) 84.23 9.28 0.66 1.76 4.07
900 87.36 8.1 0.81 1.36 2.37
1000 88.47 8.03 0.97 1.23 1.3
Rada et al. (2012) [163] 400 reactor – a quartz tube 84.6 14.2 0.2 n.a. n.a.
batch
Czajczyńska et al. [109] 400 reactor – a stainless steel 85.5 8.5 1.5 1.2 n.a.
(2020) 500 tube batch 86.3 8.3 1.3 1.2
600 87.1 7.8 1.2 1.3
Aylón et al. (2010) [98] 600 reactor – a moving-bed 88.58 10.4 1.43 1.05 n.a.
700 reactor, atmosphere N2 91.00 8.17 1.39 1.12
800 90.56 7.66 1.28 0.93
Ucar et al. (2005) [88] 650 passenger car tyre 87.57 10.35 <1 1.35 n.a.
650 truck tyre 86.47 11.73 <1 0.83 n.a.
Mastral et al. (2002) [122] 450 reactor – fixed-bed reactor, 85.62 11.11 0.2 0.9 n.a.
500 atmosphere N2 85.59 11.12 0.21 0.91
550 85.67 11.08 0.19 0.89
Galvagno et al. (2002) [121] 550 reactor – a rotary kiln, 85.62 11.55 0.44 2.49 n.a.
600 atmosphere N2 87.82 9.42 0.55 2.21
680 87.35 10.01 0.65 1.99
de Marco Rodríguez [86] 300 reactor – an autoclave, 86.50 10.7 0.3 1 1.4
et al. (2001) 400 atmosphere N2 85.90 10.6 0.3 1.1 2
500 85.60 10.1 0.4 1.4 2.5
600 86.20 10.2 0.4 1.2 2.1
700 86 10.2 0.4 1.2 2.2
Roy et al. (1999) [103] 520 vacuum pyrolysis, 86.5 10.8 0.50 2.20 0.67
passenger car tyres
500 vacuum pyrolysis, 87.6 10.50 0.40 1.50 0.69
passenger car tyres with
silica filler
485 vacuum pyrolysis, truck 87.9 11.2 0.70 0.20 0.65
tyres
Diesel elemental composition [%wt.]
Hunicz et al. (2020) [3] diesel fuel 86.5 13.5 0 0.0006 0
Biodiesel elemental composition [%wt.]
Lotero et al. (2006) [164] biodiesel (FAME) 77.00 12.00 0 0.05 11.00

creased from 34.7% at 300°C to 75.6% at 600°C. Further elevating temperatures. For example Li et al. [162] produced 8.6%wt. PAHs at
the pyrolysis temperature to 700°C reduced aromatic content to 450°C and 13.1%wt. at 500°C in a rotary-kiln reactor. Interestingly,
57.4%. The inverse trend was noticed for aliphatic compounds [59]. further temperature rise did not increase PAHs production.
The authors, conducting single-stage pyrolysis with an atmosphere Table 5 summarises the elemental composition of different TPO
of product gas at 497°C and 614°C; and with nitrogen gas at 516°C samples reported by individual authors and compares it the com-
and 617°C, also noted an inverse correlation between aromatics position of DF and the most popular biodiesel (FAME). The pyroly-
and aliphatic compounds. Conversion of aliphatic compounds to sis process temperature is also given.
aromatics resulted from the secondary atomisation reaction, with Analysis of Table 5’s summary of the TPO compositions shows
the Dies-Alder mechanism and condensation occurring at higher two tendencies are emerging. Luo and Feng [115] found that pyrol-
temperatures [102,119,160]. ysis temperature significantly affected the profile of individual el-
De Marco Rodríguez et al. [86] stated that the correlation be- ements: raising the temperature significantly increased the shares
tween aromatics and aliphatics is related to the recombination re- of C and N, while the H share gradually decreased. In contrast, de
action among free radicals of these compounds. Choi et al. [59,112] Marco Rodríguez et al. [86], Galvagno et al. [121] and Aylón et al.
observed that the dominant compounds of aromatic fraction, in- [98] did not observe significant changes in the elemental composi-
dependent of the pyrolysis parameters, were xylenes, while DL- tion of TPO with temperature.
limonene was the main aliphatic compound. Notably, these obser- The quality of crude pyrolytic oil obtained from waste tyres
vations were irrespective of the process conditions, reactor type or makes it generally suitable for use as fuel in a CI engine. TPO’s
fluidised-bed material. heating value of approximately 43 MJ/kg is close to those of other
Among other aromatics, PAHs pose a challenge because they engine fuels, such as diesel (43.3 MJ/kg), methane (50 MJ/kg), gaso-
are not only highly toxic as an exhaust gas compound but also line (44.4 MJ/kg) and biodiesel (36.8 MJ/kg) [165,166]. It also has
enhance soot formation. Aromatisation of aliphatic compounds to similar molecular composition [86,103,115,163,164,167–171]. The
PAHs is supported by high temperatures and residence time. Cun- mass-based hydrogen to carbon ratio (affecting tank-to-wheel CO2
liffe and Williams [161], using a fixed-bed reactor, observed an in- emissions) for crude TPO averages 0.128, similar to that of refined
crease of the PAH content from 1.5%wt. to 3.4%wt. when shifting diesel oil.
temperature from 450°C to 600°C. Alvarez et al. [58], using a con- TPO’s properties relevant to its use in a combustion engine are
ical spouted-bed reactor, produced PAH fractions lower than 5%wt discussed in detail in the following section. One should bear in
at temperatures of 475°C and lower. Elevating the temperature to mind that TPO’s composition, and hence its related combustion pa-
575°C increased the polyaromatic content to 12.6%wt. However, rameters, depends on reactor type, inert gas used, process temper-
some studies reported higher PAH concentrations, even in lower ature, quality of feedstock and their pre-treatment. Subsequent re-

16
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

fining used as a post-treatment will also have a bearing on TPOś Most of the kinematic viscosity values for TPO (measured
combustion parameters. at 40°C) were in accordance with EN 590ś permissible range
for diesel fuels, 2.0-4.5 mm2 /s [17,71,72,88,106,115,173,176,178–
3. Pyrolytic oil as engine fuel 182,184]. However, Ramirez-Canon et al. [111] and Miandad et al.
[73] both reported low viscosities - below 1.90 mm2 /s. On the
Already in the early 1980s, it was found that pyrolytic oil sam- other hand, Islam et al. [175], Sebola et al. [177], Islam and
ples and condensates produced by petroleum residues coking (by Nahian [183], Bhatt and Prajapati [186], Hürdoğan et al. [187] and
thermal and steam cracking of gasoil fraction and gasoline) are Ambrosewicz-Walacik et al. [39] obtained highly viscous TPO,
similar to olefins [103]. For this reason, these products can be well above the level for proper operation of most fuel injection
blended and thermally treated under the same conditions. Fur- systems. Roy et al. [103] for instance, recorded kinematic viscosity
thermore, Miandad et al. [73] suggested that TPO generated by (at 50°C) of 23.5 mm2 /s for his TPO obtained from passenger
catalytic and non-catalytic pyrolysis consisted mainly of aromatics car tyres filled with silica (fixed-bed reactor). It should be noted
with some aliphatic compounds, which have similar physical prop- that viscosity depends heavily on temperature, with hyperbolic
erties as conventional diesel fuel. characteristics differing from fuel to fuel. For the discussed sample
As discussed in the introduction, this review focuses on CI the viscosity at 100°C was 5.2 mm2 /s. Its potential direct engine
engine applications because these are the most feasible from the application, for marine use, for instance, would require preheating
future demand perspective. TPO samples should be evaluated to around 120°C.
against legal and functional requirements in order to verify if TPO Water and sulphur concentrations in crude TPO are usu-
is suitable as fuel for CI engines. Important properties, among oth- ally above the levels specified as maxima for EN 590 diesel
ers, are kinematic viscosity, density, acidity, water content, sulphur (water, 200 mg/kg; sulphur, 10 mg/kg). According to the avail-
content, flash point, distillation curve and cetane number. The able data, water contents varied between zero and 46 g/kg
following chapters review TPO’s engine-relevant properties and [93,115]. Turning to sulphur content, all the samples presented
possible quality improvement measures. Table 6 summarises the in Table 6 significantly exceeded the sulphur limit for EN 590
results of the assessment of TPO as a CI-engine fuel. Automotive diesel. Values ranged from 88 mg/kg to 15 g/kg [17,39,71–
diesel EN 590 standard is used as a reference in this table and in 73,93,103,111,173,174,176,177,179,180,182,186]. Desulphurisation
the discussion. methods are discussed in Section 3.3.
The flash point is the lowest temperature at which a liq-
3.1. Properties of crude TPO uid can form an ignitable mixture with air, when the surface of
the liquid is exposed to an ignition source. It is related to the
Heating value (or calorific value) is one of the most important volatility of the fuel, and is therefore set by distillation param-
indicators for fuel as it translates to the vehicle/vessel range and eters. The EN 590 standard regulates the flash point of diesel
is a key factor in the well-to-wheel CO2 assessment and the fuel to be at least 55°C, enabling safe storage and handling. How-
cost of transport. It should be noted here that mass-based heat- ever, most research determines TPO’s flash point to be less than
ing value of hydrocarbon fuels is straightforwardly shaped by the 30°C [71,73,88,103,175,180–182]. Some put it higher, from 31°C
ratio of hydrogen to carbon and the presence of oxygen and nitro- to 54°C [17,39,72,111,173,174,183–186]. Only a few studies as-
gen. In the case of biofuels and also TPO, water content can signif- sessed TPO’s flash point to be in accordance with the require-
icantly harm calorific value. Heating values for the TPO samples in ments of the EN 590 standard, reporting values from 58°C to 94°C
Table 6 are depicted in one of three forms: higher and lower heat- [103,106,159,177–179].
ing value (HHV and LHV respectively) or as gross calorific value One of the most important fuel characteristics is the distillation
(GCV). This is because the individual study authors did not all curve, which can explain the trends in several parameters already
adopt the same measurement principle. The difference between discussed: flash point, volatility and low-temperature properties.
the three approaches relates to how water and ash content is han- Fig. 11 shows distillation curves of TPO produced at different reac-
dled in the calculation. For further elaboration of this, refer, for in- tor temperatures, determined by Alvarez et al. [58]. For reference,
stance, to Heywood [188]. gasoline, diesel and vacuum gas oil distillation curves are included
The energy density of crude TPO can be optimised to give in Fig. 11. Note, however, that the distillation curve of TPO is sim-
diesel-like values, with LHV ranging from 38.9 to 42.5 MJ/kg ulated – based on the detailed chemical analysis. Fig. 12 shows ex-
[115,179,180]. Reported values of GCV usually range from 41.6 to perimental distillation curves determined by de Marco Rodríguez
44.8 MJ/kg [88,90,103,174,175,177,183]. Some studies, with calorific et al. [86] and Yazdani et al. [11].
values ranging from 39.2 to 48.7 MJ/kg, fail to specify the method The distillation range of TPOs in Fig. 11 shows that pyrolytic
used for its determination [72,106,111,178,184,187]. This spread of oils consist of a wide variety of chemical products with different
values stems from differences in feedstock quality and different py- boiling points. The exception to this is case 6, which was pyrol-
rolysis process parameters. Studying Table 6 reveals that there is ysed at the highest temperature (575°C) and which shows a rela-
no straightforward correlation between heating value and process tively flat distribution, consisting mostly of fractions with a boiling
temperature. For instance, Luo and Feng [115] observed that both point around 160°C. The other curves show that the gasoline frac-
parameters increase simultaneously, while Ramirez-Canon et al. tion of TPO, also called light naphtha fraction, with boiling temper-
[111] observed a significant drop in HHV after elevating the process atures < 160°C, varies from 14% to 42%. The heavy naphtha fraction
temperature above 600°C. More detailed analysis, however, reveals (160-200°C) reported by Yazdani et al. [11] for pyrolysis tempera-
the direct correlation of heating value with TPO yield. Importantly, ture 500°C was only 4%, whereas from Fig. 11 it can be read that
there is no trade-off here: maximising yield also results in a higher at the process temperature 575°C this fraction constitutes approx-
energy density of the end fuel. As with the case of mass fraction imately 70% of the TPO. The middle distillate fraction (from 200°C
yield, catalytic pyrolysis tends to produce more energetic fuel [72]. to 350°C) is reported to be between 35% and 38%, and the heavy
Note that, according to Ramirez-Canon et al. [111], the increase of oil fraction (> 350°C) ranges from 1% to 47% [11,58,86]. Generally,
mass fraction yield translates to higher heating value by virtue of a wider distillation temperature range than for DF inhibits the low-
increased density of the TPO sample. The increased density was temperature properties of TPO, at the same time causing issues
observed and ranged from 830 to 987 kg/m3 at 15°C, the latter related to low flash-point. The low-temperature issues come from
value significantly beyond that of EN 590 diesel. large fractions of heavy components. While typically 95% of diesel

17
Table 6
Physicochemical characteristics of TPO and DTPO

Reference Additional information Temperature Crude or Calorific Density Kinematic Water Sulphur Flash Cetane
of pyrolysis distilled value @15°C viscosity content content point number
[°C] [MJ/kg] [kg/m3 ] @40°C [%wt.] [%wt.] [°C] [-]
[mm2 /s]

EN 590 [172] - - - 42-44 (LHV) 820-845 2-4.5 0.02 0.001


M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

≥ 55 ≥ 51
TPO, DTPO or fraction of TPO (light, medium and heavy)

Roy et al. (1999) [103] passenger car tyres 520 Crude 43.6 (GCV) 950 (20°C) 9.7 (@50°C), 0.3 0.8 28 44
3.2 (@100°C)
passenger car tyres filled with silica 500 Crude 43.8 (GCV) 976 (20°C) 23.5 (@50°C), 1.6 1.5 59 n.a.
5.2 (@100°C)
fixed-bed reactor, truck tyres 485 Crude 44.8 (GCV) 939 (20°C) 17.8 (@50°C), 1.5 1.0 22
6.0 (@100°C)
Ucar et al. (2005) [88] fixed-bed reactor, passenger car tyre 650 Crude 41.6 (GCV) 943 4.62 n.a. n.a. <30
truck tyre 650 Crude 42.4 (GCV) 913 3.85 <30
Murugan et al. (2008) [173] steel pyrolysis reactor, atmosphere 450-650 Crude 38.0 (net 934 3.77 0.72 43
vacuum calorific
value)

18
Murugan et al. (2008) [174] steel pyrolysis reactor, 450-650 Crude 42.8 (GCV) 935 3.20 0.95 43
Distilled 45.6 (GCV) 871 1.70 0.03 36
Islam et al. (2008) [90] fixed-bed reactor 575 Crude 42.0 (GCV) 957 4.75 (@30°C) n.a. ≤ 32
Islam et al. (2010) [175] fixed-bed reactor 475 Crude 41.6 (GCV) 943 4.62 ≤ 30
İlkılıç and Aydın (2011) [176] pyrolysis reactor was electrically 700 Crude n.a. 945 (20°C) 3.85 0.90 50 44
heated
Sebola et al. (2013) [177] n.a. n.o. Crude 43.0 (GCV) 926 (20°C) 9.0 3.54 0.91 94 n.a.
Distilled 43.7 (GCV) 807 (20°C) 1.50 0.065 0.41 44
Wongkhorsub and [178] batch pyrolysis reactor 350-400 Distilled 43.2 n.a. 2.69 n.a. n.a. 68
Chindaprasert (2013)
Frigo et al. (2014) [179] thermo-mechanic cracking reactor 500 Crude 41.96 (LHV) 903 2.90 0.033 0.97 58
Martínez et al. (2014) [180] continuous auger reactor 550 Crude 40.5 (LHV) 917 2.39 0.068 0.83 23
Tudu et al. (2014) [181] rotary type reactor 375-440 Light fraction 39.2 910 3.06 n.a. 1.17-3.00 30 25-30
Al-Lal et al. (2015) [182] n.a. no data Crude 40.5 (LHV) 917 2.40 0.069 0.83 23 n.a.
Vihar et al. (2015) [105] vacuum pyrolysis 600-700 Crude 40.6 (LHV) n.a. 3.22 (@20°C) 0.012 n.a. n.a.
Ayanoğlu and Yumrutaş (2016) [71] n.a. 450 Crude 41.0 (HHV) 830 3.21 n.a. 0.013 28.1
Al Mamum et al. (2016) [93] fixed-bed reactor 320-390 Distilled - 840 0.668 nil 0.629 8
Islam and Nahian (2016) [183] cylindrical chamber as a reactor 450-600 Crude 42.0 (GCV) 956 16.39 n.a. n.a. 50
450-600 Distilled 43.6 (GCV) 835 0.89 < 10
(continued on next page)
Progress in Energy and Combustion Science 85 (2021) 100915
Table 6 (continued)
Wang et al. (2016) [184] lab-scale fixed-bed reactor, N2 350 Crude 42.1 887.2 2.71 n.a. n.a. 39 44
atmosphere 400 42.1 889.1 2.74 40 43
450 42.2 890.4 2.81 43 45
Ambrosewicz-Walacik and [185] n.o. 450-500 Crude n.a. 955 5.68 0.49 53 n.a.
Walacik (2017)
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

Bhatt and Prajapati (2017) [186] rotary reactor 430 Crude 910 8.65 Nil 0.0088 32 42
Hürdoğan et al. (2017) [187] axis rotary type reactor 500-550 Crude 40.1 907 4.98 n.a. n.a. > 100 n.a.
Luo and Feng (2017) [115] sole pyrolysis 1000 Crude 38.5 (HHV) 982.7 3.91 1.56 n.a.
pyrolysis with ceramin balls 1000 Crude 40.2 (HHV) 944.8 3.62 2.07
catalytic pyrolysis with BF slags 1000 Crude 44.1 (HHV) 935.2 2.10 4.57
pyrolysis using BF slag 600 Crude 38.9 (LHV) 941.2 3.69 1.56
700 Crude 40.6 (LHV) 961.9 3.06 2.68
800 Crude 41.0 (LHV) 986.9 3.29 3.97
900 Crude 41.6 (LHV) 955.0 2.35 4.11
1000 Crude 42.1 (LHV) 982.7 2.10 4.57
Sharma and Murugan (2017) [17] n.a. n.a. Crude 38.1 n.a. 3.35 n.a. 0.95 49 28
Ambrosewicz-Walacik et al. [39] 450-500 Crude n.a. 955.0 5.68 0.49 53 n.a.

19
(2018) pyrolytic oil distilled at three Light 770.0 0.823 0.40 < 3.5
different temperatures Medium 1003.0 0.842 0.60 12.0
Heavy 1360.0 0.891 0.85 25.0
Banihani and Bani Hani (2018) [72] pyrolysis without catalyst 550 Crude 43.5 911.7 4.60 1.13 50 45
(@20°C)
pyrolysis with alumina catalyst 551 Crude 45.5 862.8 3.90 0.37 55 45
(@20°C)
pyrolysis with zeolite catalyst 552 Crude 48.3 858.6 3.60 0.34 56 48
(@20°C)
Miandad et al. (2018) [73] pyrolysis with activated alumina 600 Crude 42.5 (HHV) 900 1.90 n.a. 27 n.a.
Ramirez-Canon et al. (2018) [111] pyrolysis with hydrogen steam 450 Crude 48.7 830 1.59 0.90 1.09 32.0
500 47.2 850 1.88 0.75 1.06 32.5
550 47.4 880 2.77 0.75 0.89 30.5
600 46.9 860 2.65 0.35 1.07 35.5
450 with H2 45.9 860 2.35 0.55 0.87 34.5
500 with H2 48.1 850 2.41 0.40 0.78 30.5
550 with H2 47.3 870 2.48 0.25 0.67 31.0
600 with H2 47.9 880 3.07 0.30 0.80 31.0
Progress in Energy and Combustion Science 85 (2021) 100915
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Such values indicate that compression ignition in typical engine


conditions is still possible, but with significant ignition delay. How-
ever, lower CN does not preclude TPO use in other engine applica-
tions, such as off-road or marine (see Table 1 for reference). One
should note that the cetane number is shaped by chemical compo-
sition of hydrocarbons. Generally, normal paraffinic hydrocarbons
exhibit the best auto-ignition properties, whereas branched paraf-
fins, olefins and aromatics reduce the cetane number. Additionally,
a greater number of C atoms leads to better auto-ignition proper-
ties, but carries the penalty of higher viscosity [188,189].
Another obstacle to TPOś application is its acidity, which af-
fects corrosive activity and propensity to create deposits. Acidity
usually does not pose a challenge for mineral DF, while for biofu-
els it is an issue, mainly due to some content of carboxylic acids.
Both European and US standards (EN 14214 and ASTM D6751) limit
acid value of biofuels to 0.5 mg KOH/g. TPO also contains some
amounts of carboxylic acids and additionally can contain sulphuric
acids. However, different research provides very different acidity
levels. Ambrosewicz-Walacik et al. [39] determined acidity to be
4.33 mg KOH/g. Antoniou and Zorpas [190] reported a much lower
value of 0.45 mg KOH/g. Raj et al. [81] demonstrated diesel-like
acidity (0.1-0.21 mg KOH/g). Alvarez et al. [58] and Yazdani et al.
Fig. 11. Simulated distillation curves of TPOs obtained at 442°C, 475°C and 575°C [11], independently indicated that the content of carboxylic acids
with reference curves of commercial fuels and vacuum gas oil. Reproduced from in TPO drops with increasing pyrolysis temperature. However, com-
[58] with permission from Energy, Elsevier. parison of other studies’ acid values and process temperatures do
not support this statement. These discrepancies suggest that acid-
ity of TPO is affected by many factors, not yet sufficiently studied
in the literature.
Finally, long-term storage capabilities are an important con-
sideration for new fuels or their components. Storage capabili-
ties of fuels are affected mainly by oxidation, thermal or thermal-
oxidative decomposition, hydrolysis or microbial contamination
from dust or water [JH003]. Oxidative stability is one of the pri-
mary concerns, but high water-content in TPO might also involve
issues of hydrolysis and microbial growth. Nevertheless, available
reports mention satisfactory storage capabilities of TPO. Alvarez
et al. [58] stated that TPO properties remain unchanged with time.
More quantitatively, Ambrosewicz-Walacik et al. [185] determined
oxidative stability of the TPO distillate (EN 14112 method), to be
above 20 hours – a value considered sufficient for EN 590. Sharma
and Murugan [17] pointed out that TPO contains some phenolic
compounds, which may act as antioxidants. They observed that
20% TPO admixture in biofuel – jatropha methyl ester (JME) - im-
proved oxidation stability approximately 2.5 times. Moreover, the
stability effect of TPO admixture was greater than that of DF. Ac-
cording to the above remarks, TPO seems to support oxidative sta-
bility of biofuels when blends are considered.

3.2. Properties of TPO distillates (DTPO)


Fig. 12. Distillation curves of TPOs obtained at 500°C by de Marco Rodríguez et al.
[86] and at 550°C by Yazdani et al. [11]. Reproduced from [11] with permission from Distillation is one of the main methods to improve TPO fuel
Waste Management, Elsevier. quality, with different fractions being more suitable for different
applications, either as fuel or as a fuel additive. The distillation
fuel distillates below 340°C, this characteristic temperature for TPO curves of TPOs (see previous subsection for reference) show that
can be as high as 500°C. The low flash point (flammability) issue pyrolytic oils consist of wide variety of chemical products with dif-
relates to some fraction of gasoline-like components with low boil- ferent boiling points. Literature specifying the properties of differ-
ing temperatures. ent TPO fractions is, however, limited. Where available, properties
The distillation curves of crude TPO suggest that both the flash of TPO distillates are included in Table 6, providing comparison
point and low-temperature properties could be improved if nar- with the properties of crude TPO obtained via the same pyrolysis
rower fractions are isolated by further distillation. The properties process.
of distilled TPO are presented in the next subsection. The recent work by Ambrosewicz-Walacik et al. [39] outlines
Cetane number (CN) is of great importance because it deter- the systematic fractionation of TPO, along with an assessment of
mines the fuel’s ability to auto-ignite under CI engine conditions. the physicochemical properties of the fractions obtained. Three lev-
The EN 590 standard for automotive-grade fuels specifies a min- els of distillation were considered: A) Light Fraction: consisting of
imum cetane number of 51, whereas values for crude TPO are components with boiling temperature < 160°C. This fraction was
rather lower, ranging from 25 to 48 [17,72,103,106,176,184,186]. distilled in order to separate components with high volatility, sig-

20
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

it did not meet ISO 8217 for bunker fuel quality. Sulphur content
and moisture levels of the TPO fuel present the biggest issues. At
this point, it is necessary to mention that, in principle, TPO can
be used as feedstock to standard refinery processes to produce
automotive-grade diesel and other fuels. However, this route has
not been demonstrated openly and consequently will not be cov-
ered in this review.
Moisture removal does not pose a particular challenge and can
be achieved with a variety of well-established methods which
Fig. 13. Crude TPO sample (left) and products of its distillation analysed in
are standard for many alternative fuel components. These meth-
Ambrosewicz-Walacik et al. [39]; A – light naphtha fraction, B – medium fraction, ods involve mechanical separation (centrifugation), physical pro-
C – heavy fraction; author’s archive. cesses (percolation through a bed of anhydrous substance such as
sodium sulphate, vacuum drying) or thermal processing (distilla-
nificantly influencing the low flash point of the pyrolytic oil. B) tion) [47,174,181]. There are also chemical methods based on sup-
Medium Fraction: boiling temperature from 160°C to 204°C. This plementation with additives exhibiting water-binding properties
fraction was considered as a diesel fuel component and subjected [191]. As mentioned, these methods are commonplace and thus
to further engine tests in [39]. C) Heavy Fraction: components with will not be separately discussed. Moisture removal helps increase
boiling temperature from 205°C to 350°C. Note that at over 350°C, the heating value and improve the CN of the fuel and is important
only a tar-like substance remained in the distillation flask. As well from the perspective of preventing engine corrosion and tribologi-
as the physicochemical comparison given in Table 6, the differ- cal issues [106].
ences in the appearance of crude TPO and its distillates are seen Sulphur content of liquid engine fuels has been drastically re-
in Fig. 13. It clearly shows that the added value of distillation is duced in the past few decades, cutting the sulphur compounds
the reduction of ash and tar content in the fuel [39]. in exhaust emissions. These are environmentally harmful, mainly
It was shown that the distillation process dramatically reduced because of their potential to cause acid rain. Furthermore, fuel
kinematic viscosity (to below 0.9 mm2 /s for all samples) while sulphur increases particulate matter and soot emissions in inter-
also contributing to the unfavourable decrease of the flash point nal combustion engines [192]. There are also technical reasons for
(Table 6). The flash point for the light fraction was below 3.5°C; cutting fuel sulphur: sulphur causes corrosion and has a delete-
it was 12°C and 25°C for the medium and heavy fractions respec- rious effect on advanced catalytic systems used for CO, HC, NOx
tively. The crude TPO’s flash point was 53°C, almost meeting EN and particulate reduction [193,194]. This underlines the impor-
590’s threshold value of 55°C for diesel. tance of desulphurisation for TPO, especially bearing in mind the
Other authors had studied only selected distillates, usually possible direct use of TPO in shipping, where IMO has introduced
medium and light fractions, but the above trends were confirmed 0.5/0.1%wt. fuel sulphur limits. The automotive EN 590 standard
by Al Mamun et al. [93], Islam and Nahian [183] and Murugan is already very stringent, limiting fuel sulphur content to just 10
et al. [174]. Describing what they called a light TPO fraction, Tudu mg/kg (0.001%wt.).
et al. [181] reported viscosity of 3.06 mm2 /s (within EN 590’s per- The sulphur in pyrolytic oil is related to the fact that this
mitted range) and a moderate flash point of 30°C. compound is used as a cross-linking agent during tyre manufac-
Murugan et al. [174], Sebola et al. [177] and Islam and Nahian turing. Sulphur-containing compounds undergo different reactions
[183] recorded that the distillation process contributed to the in- (recombination and/or disproportionation) during pyrolysis. As dis-
crease in calorific value of analysed samples. Their medium frac- cussed in Sections 3.1 and 3.2, TPO/DTPO may contain different
tions had respectively 2.8 MJ/kg, 0.7 MJ/kg and 1.6MJ/k higher en- amounts of sulphur, depending on the process conditions and dis-
ergy density than their individual crude TPO samples. At the same tillation temperature.
time, it is difficult to assess how distillation affects CN. Actually, A conventional refinery hydrodesulphurisation (HDS) process
only Tudu et al. [181] reported that their fractionated light DTPO can be used to reduce sulphur content in fuel, yet its appli-
product had a CN of 25-30 (depending on the pyrolysis tempera- cation is primarily limited to distillates (light to medium frac-
ture) while having a more or less constant calorific value of 39.2 tions) [195]. Debek
˛ and Walendziewski [74] applied hydrotreat-
MJ/kg. However, that work lacks any comparison with the crude ment with a NiMo–Al2 O3 catalyst (a standard refinery configura-
TPO. Its authors also reported the light fraction’s very high sen- tion for petroleum-based products), cutting sulphur content from
sitivity to sulphur penetration, depending upon pyrolysis process 1.1% to 0.2%. Additionally, double carbon bonds were effectively de-
temperature. Its sulphur content varied from 1.2% to above 3%wt. stroyed. This required quite strict hydro-refining conditions, using
From Ambrosewicz-Walacik et al. [185] we can deduce that the 360°C and 5 MPa H2 pressure. Hita et al. [196,197] also performed
sulphur content of fractions follows the same trend as density. HDS with a NiMo catalyst, while trying six different porous sup-
The light fractions accumulated less sulphur than the baseline TPO ports, including a conventional composition with -Al2 O3 , as in the
(0.40%wt. vs 0.49%wt.), while having significantly lower density previously referred study. The authors selected –Al2 O3 , SiO2 –Al2 O3
than crude TPO (770 kg/m3 vs 955 kg/m3 ). The medium fraction and a mesoporous material referred to as MCM-41 as the most ac-
already had both indicators slightly above baseline TPO. The heavi- tive catalytic supports. All of them provided roughly equal sulphur
est fraction had a density of 1360 kg/m3 and sulphur concentration removal potential of around 90% for TPO samples with 2% and 1%
of 0.85%wt. sulphur content, leaving 0.2% and 0.1% sulphur respectively in the
end samples. The differences were observed in H2 uptake and fi-
3.3. Quality improvement measures for TPO/DTPO nal fuel composition. The SiO2 –Al2 O3 -based catalyst provided the
highest yield of diesel (57%wt.) and naphtha (25%wt.), while the
Section 3.1’s and 3.2’s review of TPO/DTPO parameters related conventional –Al2 O3 setup gave the highest fuel quality, charac-
to use as fuel in a CI engine reveals a large spread in the results terised by an increased paraffinic fraction and improved CN. The
obtained in different reactors. Nevertheless, it is clear that most of latter was as much as five points higher than the base TPO.
TPO’s physicochemical parameters did not meet the EN 590 stan- The above numbers indicate that state-of-the-art HDS has po-
dard. In the case of a few determinants, depending on pyrolysis tential to reduce the sulphur content of crude TPO to the level
methods, nor did TPO comply with other standards. For example, of 0.1%, independently of the base sampleś sulphur content. It

21
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

should be noted that high consumption of hydrogen, expensive the maximum sulphur reduction, cutting the TPO’s sulphur con-
molybdenum catalysts and low stability of the support materi- tent from the baseline approximately 1%wt. to 0.63%wt. Interest-
als deemed effective for TPO HDS are considered as disadvantages ingly, calcium oxide showed only a marginal desulfurisation abil-
of this method [195,197]. Refining processes that remove sulphur ity although the cited authors did refer to the already mentioned
from the fuel also simultaneously reduce the fuel’s lubricity. successful result by Aydın and İlkılıç [118]. Ahmad et al. supposed
Another popular desulphurisation method uses oxidation of the that using calcium oxide at higher temperatures would give a big-
sulphur compounds, either by methanol extraction or silica gel col- ger sulphur reduction [199]. They also found that the viscosity of
umn adsorption of the more polar, oxidized sulphur compounds tyre pyrolysis oil fell significantly with increasing addition of sul-
from methanol. Al-Lal et al. [182] used both methods on TPO pro- phuric acid and the hydrogen peroxide/acetic acid combination.
vided by a third party, yet failed to lower the sulphur content to a Viscosity was not similarly affected by the addition of calcium ox-
significant degree. ide, however. Also evaluating ionic solutions, Koc and Abdullah
There were a couple of successful attempts to reduce TPO’s [200] started desulphurisation by preparing a binary solution of
sulphur by optimising the pyrolysis process itself, either by ded- tetraoctylammonium bromide and hydrogen peroxide (30% purity),
icated reactor conditioning or by adding catalysts. Analysis of to which a phosphotungstic acid catalyst was added. The presented
Table 4 shows that the application of catalysts during pyrolysis solution cut the sulphur content in the TPO sample from 0.768%wt.
could be effective in lowering TPO’s sulphur content. Banihani and to 0.321%wt.
Bani Hani [72] employed this technique to cut sulphur in pyrolytic In summary, recent years have brought significant progress in
oil by around 70%. The sulphur content of 1.13% in their TPO with new desulphurisation techniques for TPO. Despite this progress,
baseline tyre pyrolysis without a catalyst was cut to 0.37% and none of the reviewed works actually managed to demonstrate
0.34% when alumina and zeolite catalysts respectively were added compliance with EN 590’s stringent sulphur limit of just 10 mg/kg
to the process. Importantly, in both catalytic cases, the reduction in (0.001%wt.) for automotive diesel. However, the IMO ECA sulphur
sulphur was accompanied by improvement of all the other proper- limit of 0.1%wt. is apparently already within range for TPO and fur-
ties of the TPO (refer to Table 6 for details). ther sulphur reduction can be achieved by combining the individ-
Note that already in 20 0 0 Murena [198] considered using hy- ual measures discussed in this section [186].
drogenative pyrolysis to reduce TPO’s sulphur content. Coupled
with optimising the feedstock pre-treatment (tyre scrap versus tyre 3.4. Blending with other fuels and TPO addification
powder), the author observed that the use of tetralin as a hydro-
gen donor could cut sulphur by 30%. However, the tyre/tetralin ra- The previous sections´discussions on TPO’s physicochemical
tio used to achieve this degree of sulphur reduction was over 1:3, properties such as density, viscosity, sulphur content, flash point
so the large amount of hydrogen donor solvent required could sig- and CN, clearly lead to the conclusion that it is not feasible to
nificantly influence the economics of tyre pyrolysis rendering this use pure TPO as an automotive drop-in fuel. However, blending
method uncompetitive. TPO or DTPO with DF, biodiesel or SVO is possible, while keeping
Choi et al. [59,112] tackled sulphur reduction by using differ- blended fuel properties within standards. Moreover, some specific
ent reactor conditions and reactor addification. They concluded properties of the TPO/DTPO can be utilised for regulation of the
that, independent of the pyrolysis process (single-stage or two- overall properties of the fuel. This section reviews the state of re-
stage) and type of inert gas (product gases or N2 ), raising py- search into TPO/DTPO blends with other fuels and is summarised
rolysis temperature generally increases the sulphur content, with in Table 7.
different sensitivities for individual pyrolysis types [59]. Using N2 Currently, there appears to be little attention paid to research
as the fluidising medium with two-stage pyrolysis generally facil- into use of crude TPO in binary mixtures with mineral diesel. Hür-
itated the lowest sulphur in crude TPO. Sulphur accumulation de- dogan et al. [187] analysed blends consisting of 50, 80 and 90%vol.
creased from 0.92%wt. at baseline conditions (single-stage; prod- of crude TPO and corresponding amounts of DF. The authors ob-
uct gases; temperature optimised for liquid yield), to 0.55% at op- served a gradual reduction of the mixture’s density and kinematic
timised operation (sulphur content co-optimised with temperature viscosity as the proportion of DF was increased. Values of den-
for maximum TPO yield and heating value). In the follow-up work, sity and kinematic viscosity for TPO50 (50% TPO) were significantly
Choi et al. started from the previously optimised pyrolytic reactor lower than for neat TPO and were in accordance with the require-
and experimented with different process additives (dolomite and ments of EN 590. Note that the flash point of their baseline TPO
olivine, both natural and calcined) in search of a further reduction was surprisingly high (> 100°C), so the authors did not consider
in the TPO’s sulphur [112]. The TPO sample produced with natu- this as a constraint. CN, however, was not reported.
ral and calcined olivine contained low amounts of sulphur (0.46 Martínez et al. [180] investigated a blend of 5%vol. crude TPO
and 0.45%wt. respectively). Interestingly, the addition of natural and 95%vol. of DF. They achieved a fuel compatible with the au-
dolomite increased sulphur to 0.70%wt., while calcine dolomite had tomotive diesel standard in all aspects except water and sulphur
no significant effect. content: their values exceeded EN 590 limits by 100%. This was
Aydın and İlkılıç [118] investigated the effect of adding CaO due to a rather high content of both species in the crude TPO.
and Ca(OH)2 as sulphur-mitigating agents to a pyrolytic reactor fed Sharma and Murugan [201] used biodiesel produced from jat-
with tyre scraps. After optimising the amount and ratio of both ropha oil to prepare blends using the addition of crude pyrolytic
compounds the authors demonstrated a reduction in sulphur con- oil, from 10% to 50%. The cited authors found that increasing the
tent from 1.4%wt. to 0.9%wt., without significantly impacting other share of jatropha methyl esters reduced the blend’s density, while
TPO parameters. Additional post-pyrolysis desulphurisation with kinematic viscosity and flash point increased. Additionally, the an-
sulphuric acid (10% aqueous solution) cut the sulphur content to tioxidative properties of TPO were utilised to compensate for weak
0.2% wt. biofuel oxidative stability. A similar tendency was observed in the
This latter post-treatment method is commonly referred to as later work of the same authors [17]. The authors claimed that
desulphurisation through ionic solutions or liquid-liquid extraction. biodiesel and TPO are complementary in terms of viscosity (too
Ahmad et al. [199] analysed the impact of various desulphurising high for biodiesel), flashpoint (too high for TPO) and CN (too low
ionic agents on TPO, using sulphuric acid; a combination of hydro- for biodiesel). The 20% admixture of TPO was considered the near-
gen peroxide and acetic acid; Fuller’s earth; and calcium oxide. A est to meeting EN 590: only its sulphur content and viscosity were
15%wt. solution of sulphuric acid as the extracting medium gave slightly too high. This could have been corrected by more efficient

22
Table 7
Basic characteristics of monocomponent, binary and ternary mixtures of TPO/DTPO
Composition of mixtures [%vol.] Calorific Density @15°C Kinematic Water content Sulphur content Flash point Cetane
Reference value [kg/m3] viscosity [%wt.] [%wt.] [°C] number
Rapeseed oil
[MJ/kg] @40°C [-]
TPO DTPO DF Biodiesel (SVO) Others
[mm2 /s]
Monocomponent mixtures (TPO/DTPO)
Murugan et al. [174] 100 - - - - - 42.80 935 3.2 n.a. 0.95 43.0 n.a.
(2008b) (GCV)
- 100 - - - - 45.60 871 1.7 0.03 36.0
(GCV)
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

Murugan et al. [47] - 100 - - - - 43.7 807 (20°C) 1.5 0.065 0.41 44.0
(2008c) (GCV)
Sharma and Murugan [201] 100 - - - - - 39.2 920 (20°C) 5.4 n.a. n.a. 43.0
(2013)
Martínz et al. (2014) [180] 100 - - - - - 42.70 917 2.39 0.069 0.83 23.0
(HHV),
40.49
(LHV)
Al Mamun et al. [93] - 100 - - - - - 840 0.668 nil 0.629 8.0
(2016)
Tudu et al. (2016) [106] - 100 - - - - 39.20 910 3.06 n.a. n.a. 75.0 25-30
- - - - - 100 DMC 15.78 1.079 0.625 18.0 35-36
Ambrosewicz-Walacik [185] - 100 - - - - n.a. 870 1.05 1.22 12 n.a.
and Walacik (2017)
Hürdoğan et al. [187] 100 - - - - 40.13 907 4.98 n.a. >100
(2017)

23
Sharma and Murugan [17] 100 - - - - - 38.10 913 3.35 0.95 49.0 28
(2017)
Ambrosewicz-Walacik [39] - 100 - - - n.a. 955 5.68 0.49 53 n.a.
et al. (2018)
Binary mixtures
Murugan et al. [174] - 80 20 - - - 46.40 832 1.94 n.a. 0.23 47.0 n.a.
(2008b) (GCV)
- 90 10 - - - 45.80 865 1.73 0.23 37.0
(GCV)
Murugan et al. [47] - 80 20 - - - 45.90 860 1.76 0.21 39.0
(2008c) (GCV)
- 90 10 - - - 45.90 865 1.73 0.23 37.0
(GCV)
- 5 95 - - - 44.8 837 (20°C) 2.9 0.021 0.067 50.0
(GCV)
- 10 90 - - - 44.8 837 (20°C) 2.8 0.021 0.085 48.0
(GCV)
- 15 85 - - - 44.7 837 (20°C) 2.5 0.014 0.106 42.0
(GCV)
- 20 80 - - - 43.7 837 (20°C) 2.4 0.017 0.206 44.0
(GCV)
- 25 75 - - - 43.7 836 (20°C) 2.2 0.011 0.226 38.0
(GCV)
- 30 70 - - - 43.7 836 (20°C) 2.0 0.007 0.233 34.0
(GCV)

(continued on next page)


Progress in Energy and Combustion Science 85 (2021) 100915
Table 7 (continued)
Reference Composition of mixtures [%vol.] Calorific Density @15°C Kinematic Water content Sulphur content Flash point Cetane
TPO DTPO DF Biodiesel Rapeseed oil Others value [kg/m3] viscosity [%wt.] [%wt.] [°C] number
(SVO) [MJ/kg] @40°C [-]
[mm2 /s]
Sharma and Murugan [201] 10 - - 90 (Jatropha) - - 39.24 883.5 (20°C) 5.73 n.a. n.a. 64.0
(2013) 20 - - 80 (Jatropha) - - 37.74 887.0 (20°C) 5.60 60.0
30 - - 70 (Jatropha) - - 36.40 892.3 (20°C) 5.41 55.0
40 - - 60 (Jatropha) - - 35.12 894.0 (20°C) 5.29 49.0
50 - - 50 (Jatropha) - - 33.89 897.5 (20°C) 5.15 44.0
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

Martínez et al. (2014) [180] 5 - 95 - - - 45.02 (HHV), 848 2.73 0.008 0.04 n.a. n.a.
42.23 (LHV)
Al Mamun et al. [93] - 50 50 - - - - 833 1.44 nil 0.621 8.0
(2016)
Tudu et al. (2016) [106] - 40 60 - - - 41.96 862 2.87 n.a. n.a. 32.0 41.17
Hürdoğan et al. [187] 90 - 10 - - - 41.91 872 4.15 75.7 n.a.
(2017) 80 - 20 - - - 42.80 843 3.13 73.5
50 - 50 - - - 44.05 839 2.91 72.5
Sharma and Murugan [17] 20 - - 80 (jatropha) - - 38.82 887 5.20 0.18 132 52
(2017) 40 - - 60 (jatropha) - - 38.60 894 4.90 0.39 119 48
60 - - 40 (jatropha) - - 38.00 901 4.79 0.56 108 45
Jeyakumar et al. [203] 20 - 80 - - - 43.17 (NCV)∗∗ 831 2.93 n.a. 65 54
(2018)
Ternary mixtures
Tudu et al. (2016) [106] - 40 50 - - 10 39.15 779 2.78 n.a. n.a. 38.8 40.60
DMC∗

24
Ambrosewicz-Walacik [185] - 1.00 49.50 49.50 (frying - - n.a. 853 3.30 0.0061 37 n.a.
et al. (2017) oil)
- 0.91 35.78 63.31 (frying - - 848 3.38 0.0063 36
oil)
- 2.00 49.00 49.00 - - 849 2.93 0.0123 31
(frying oil)
- 1.98 38.62 59.40 - - 852 3.21 0.0122 29
(frying oil)
Ambrosewicz-Walacik [39] - 5 40 - 55 - 865 11.42 0.027 55
and Walacik (2018)
Jeyakumar et al. [203] 20 - 80 - - 43.10 (NCV) ∗∗
830 2.89 - 68 55
(2018) CaCO3
20 - 80 - - TiO2 43.08 829 2.87 - 67 55
Mikulski et al. (2019) [41] 5 40 - 55 - 39.5 (LHV) 868 11.56 0.0271 54.5 n.a.
5 45 - 50 - 39.6 (LHV) 861 8.94 0.0277 55.0
5 50 - 45 - 40.1 (LHV) 860 8.43 0.0275 56.0
5 55 - 40 - 40.7 (LHV) 852 7.97 0.0276 56.5
5 65 - 30 - 41.1 (LHV) 851 6.63 0.0140 57.0


DMC - dimethyl carbonate
∗∗
NCV- net calorific valueThe distilled medium fraction of TPO obtained by Ambrosewicz-Walacik and Walacik [185] had extremely low viscosity. Hence, it was considered that only minor addition of this DTPO (between 1%vol.
and 2%vol.) would be sufficient for use as a viscosity improver. This was tested by adding it to a variety of blends comprising diesel and methyl esters produced from frying oil, all of which suffered from viscosity issues. The
thesis was verified negatively since the density and kinematic viscosity of the blends did not improve significantly after the addition of DTPO. However, even this small addition of pyrolytic oil contributed to a significant increase
in sulphur content and a reduction of flash point. The amount of sulphur was mostly determined by how much DTPO was added.
Progress in Energy and Combustion Science 85 (2021) 100915
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

desulphurisation of the baseline TPO and by use of a viscosity im- The overall fuel properties obtained by TPO/DTPO blending dif-
prover. Flash point values of all tested mixtures complied with EN fer significantly, depending on the share and quality of the individ-
590. However, it is worth noting that the claimed flash point val- ual components (bio and synthetic). However, from the above dis-
ues reported for corresponding blends in Sharma and Murugan cussion it is possible to make the general observation that pyrolytic
[17] and their earlier work [201] are significantly different (see oils and biofuels seem complementary in terms of their proper-
Table 7). Thus, the absolute values should be treated with caution ties. This relates mostly to viscosity (too high for biofuels) and
but the general statement regarding compliance of all the tested flash point (too high for TPO). This complementary relationship
mixtures with EN 590’s flash point limit still holds. also holds true, albeit to a lesser extent, to heating value (lower
Other authors considered the use of distilled TPO to produce energy density of biofuels can be compensated by TPO) and den-
binary mixtures with diesel or biodiesel. Murugan et al. [47,174] sity. Note that almost none of the authors considering a high bio
examined DTPO blends with DF, using 80%vol. and 90%vol. DTPO component share or a high TPO share in diesel (above 50%) was
and 20%vol. and 10%vol. DF. They observed little changes in anal- able to demonstrate full compliance with EN 590. Nevertheless, the
ysed properties such as density, kinematic viscosity, sulphur con- synergy of pyrolytic oils and biofuels, along with favourable results
tent and flash point compared with neat DTPO. Addition of DF ob- achieved through addification of TPO with active particles, suggests
viously improved all these properties but such high proportions of that future renewable liquid fuel research should focus on multi-
DTPO (80% or more) with DF should be rather considered as an component mixtures. Using multiple components adds complexity
academic idea because the resultant blend still falls well short of but addresses the feedstock availability issue.
EN 590. Al Mamun al. [93] investigated the parameters of a 50/50
4. Engine tests
blend of TPO and DF: their results led them to conclude even a
50/50 blend cannot satisfy EN 590. The authors, however, pointed
There is a handful of engine research results involving TPO but
out that DTPO’s very low viscosity of 0.668 mm2 /s makes it suit-
it is difficult to draw firm conclusions about how TPO impacts
able as a viscosity improver. Tudu et al. [106] added 40% DF to
combustion, performance and emissions. This ambiguity stems
their baseline DTPO and noticed the significant reduction in kine-
from the variation in the tested engine platforms, operation con-
matic viscosity and density. The calorific value and cetane number
ditions and control strategies used by different authors. Addition-
were close to the weighted (by volumetric shares) average of the
ally, differences in feedstock quality, pyrolysis conditions and post-
values of the individual components. As a cautionary note, the 40%
processing also add unwanted variables because they influence
addition of DF reduced the flash point from the neat DTPO sam-
TPO’s combustion properties and its propensity to produce toxic
ple’s 75°C to just 32°C, implying that the fuel used as DF represen-
exhaust compounds. Finally, TPOs can be admixed to different base
tative did not meet the EN 590 flash point requirement.
fuels, along with other additives, creating binary, ternary or even
Tudu et al. also prepared a ternary blend, with 40%vol. of DTPO,
more complex blends. We have attempted to use a systematic ap-
50%vol. of DF and 10%vol. of dimethyl carbonate (DMC) [106]. This
proach to this complex issue, presenting a summary of the most
had significantly lower density and slightly lower kinematic vis-
important findings which, in our opinion, form the best thinking in
cosity than the binary blend, but suffered from reduced calorific
this field. It starts with the effects of mixture formation on com-
value and cetane number index. Hariharan et al. [202] also sug-
bustion, and then deals with emissions. Additionally, results ob-
gested that diethyl ether could be used as a cetane number im-
tained from different engine platforms and for raw TPO, DTPO and
prover for TPO. The second part of Table 5 and the following dis-
more complex blends are separated in order to provide a reliable
cussion review ternary mixtures using TPO/DTPO as a component.
benchmark in each major sub-category of influential factors.
Follow-up works by Ambrosewicz-Walacik et al. [39], Mikulski
et al. [18] disregarded the sulphur constraint, assuming desulphuri- 4.1. Mixture formation and combustion
sation is possible. They increased the DTPO share to 5%vol, adding
it to FAME/DF blends and to rapeseed SVO (cold-pressed and fil- As far as we can ascertain, there are no studies specifically fo-
tered) blends with DF, mixed at different ratios. The aim was to cused on TPO spray development, atomisation and mixture forma-
find an economically feasible fuel needing only limited processing tion. Therefore, spray behaviour of TPO and its blends can be only
and with major shares of renewables. A detailed characterisation qualitatively assessed based on engine test results and explained
of the different blends can be found in Table 7. The flash point via the fuel’s physical properties.
of all the blends was close to that of EN 590 diesel. The addition In conventional diesel combustion, some portion of fuel evapo-
of 5% DTPO greatly improved viscosity (by 20%-50%, depending on rates and mixes with air during the ignition delay period. This frac-
the amount of biocomponent) although it was still well above EN tion, dependent on fuel volatility, viscosity and mixing time (gov-
590’s requirement. The parameters, however, were still considered erned by ignition delay) burns as a premixed charge. The remain-
suitable to operate modern diesel engines and so the samples were ing, usually much larger fraction, is combusted in a diffusion flame
subjected to engine tests focused on analysis of combustion and on associated with the developing spray. Crude TPO has low CN, thus
their full emission and efficiency profiles [39,41]. contributing to elongated chemical ignition delay. It also appears to
One of the newest methods used to improve fuel characteristics be more volatile as it has a lower flashpoint. Both factors should
involves nano-additives. Nanoparticles of titanium dioxide (TiO2 ) increase the premixed burn fraction, resulting in a more rapid
can reduce pollutants such as NOX and SOX [204]. The effects of pressure rise and higher peak temperatures than for DF combus-
those additives were already confirmed by Fangsuwannarak and tion. These deliberations are generally confirmed by combustion
Triratanasirchai, who tested blends of palm biodiesel with fossil and emissions studies reviewed in detail below. Note, however,
diesel and 0.1% or 0.2% addition of TiO2 . [205]. The TiO2 addition that optical studies on spray development and mixture formation
contributed to the reduction of kinematic viscosity of different fu- are highly recommended to support fundamental understanding of
els blends. Results of tests using an automotive engine showed that combustion and emission trends for TPO and its blends.
nanoparticles improved combustion completeness. Jeyakumar et al. Tests with crude TPO on basic engine platforms
[203] analysed the effect of calcium carbonate and titanium diox- Basic engine platforms include those that are naturally as-
ide nano-additives on a mixture of 20% TPO with 80% DF. They pirated, without exhaust gas recirculation (EGR) and with low-
noted slight decreases of viscosity and density, but greater reduc- pressure mechanical injection systems providing single injection.
tions of cloud and pour points. The flash point and cetane number Such engines seem obsolete when compared with modern auto-
both increased. motive designs, but their simplicity facilitates combustion study at

25
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

a fundamental level. The limited number of control parameters en-


ables reliable comparison of different research works. All the re-
search platforms included in this review section have a cylinder
bore ranging from 80 mm to 110 mm, typical of light- to mid-
duty, high-speed engines. They share a similar combustion system
design, using direct fuel injection with multi-hole injectors and in-
piston toroidal combustion chamber.
Murugan et al. [173] performed a series of experiments using
a single-cylinder, naturally-aspirated, air-cooled engine with 17.5:1
compression ratio. Fuel was delivered in a single dose, using a
low-pressure mechanical injection system. Experiments were per-
formed at a fixed speed of 1500 rpm and under variable loads,
with fixed start of injection (SOI) timing. Raw TPO was mixed
with DF in fractions of 10%, 30% and 50% vol. It was observed
that increasing the TPO fraction led to a longer auto-ignition delay,
which was attributed to TPO’s higher viscosity, poorer atomisation
and slower vaporisation. More importantly, it was noted that be-
cause of the longer ignition delay of TPO blends, these binary fu- Fig. 14. In-cylinder pressure for variable admixture of TPO to DF. Data acquired
at 1500 rpm, medium load, single low-pressure injection, SOI = 28°CA bTDC [16].
els were characterised by higher pressure rise rates (PRR), which
Courtesy of Prof. Stanisław Szwaja, Czestochowa
˛ University of Technology.
was associated with the combustion mode moving from diffusion-
controlled to premixed.
Frigo et al. [179] used a similar experimental platform, but with to substantially modify the fuel injection system to achieve appro-
a high compression ratio of 20.3:1. The single-pulse, low-pressure priate material compatibility.
injection was commenced at 12° of crank angle (CA) bTDC. The en- Tests of DTPO and DF on basic engine platforms
gine tests were performed at full load and at speeds from 20 0 0 It should be noted that addition of TPO in DF must be limited to
rpm to 3500 rpm. When the TPO content in DF did not exceed 20%, ensure that the final properties of the blended fuel remain within
the combustion process was very similar to that of pure DF. How- the specifications set by the relevant standards, especially for au-
ever, higher fractions of TPO provided substantial delay of auto- tomotive applications. Thus, the studies applied a variety of treat-
ignition, and additionally rapid increase of cyclic variability, espe- ment methods, including distillation and hydrotreatment, to up-
cially at low engine-speeds. A 40% TPO mixture at 30 0 0 rpm pro- grade the TPO’s properties to enable its blending with DF at higher
vided approximately 5°CA delay of SOC when compared to DF. As ratios. Many authors did not mention the distillation process in
with previous studies, a low CN and high viscosity were indicated their works, but the viscosity values indicated that purified, light
as reasons to limit TPO content in DF. fractions of TPO were used. Thus, we apply the term DTPO to all
Uyumaz et al. [206] used a single-cylinder, air-cooled, naturally- fuels with viscosity within or below the range specified in EN 590
aspirated engine with 18:1 compression ratio running at 2200 rpm [172].
with fixed SOI timing at 24°CA bTDC and with variable loads, i.e. Kumar Singh and Prabu [208] performed their tests using
amounts of injected fuel. Comparison of DF and its mixture with a single-cylinder, water-cooled engine with mechanical injection,
10% of TPO showed that auto-ignition delay was longer for the where SOI was fixed at 21°CA bTDC. Engine load was varied from
TPO blend. Addition of TPO delayed auto-ignition at low loads idling to 0.435 MPa of brake mean effective pressure (BMEP) at
by approximately 0.1 ms, which, at the given rotational speed, 1500 rpm. In line with other studies, raising DTPO content in DF
translated to 1.3°CA. Differences in SOC diminished as the load delayed combustion and increased the peak heat release rate (HRR)
increased. when compared with pure DF. However, the higher HRR applied
Chwist et al. [16] also used a basic engine platform, with 16.8:1 only until DTPO’s admixture rate reached 75%. With even higher
compression ratio. The fuel was injected with SOI set to 28°CA admixtures the peak HRR was reduced due to a substantial com-
bTDC. Rotational speed was 1500 rpm under fixed, medium load. bustion delay of approximately 12°CA. The peculiar effect observed
As with the above-mentioned studies, the results showed SOC de- was that auto-ignition was back advanced at 85% DTPO: the peak
lay with increasing TPO content. Using pure TPO, SOC was delayed HRR location was between the one for DF and the one for 75%
by approximately 8°CA in comparison with DF. The in-cylinder DTPO. Importantly, end of combustion occurred at nearly the same
pressure curves from this study are shown in Fig. 14. This study location for all tested fuels, from pure DF to the blend with 85%
also showed that the average SOC demonstrated higher variability, DTPO.
where ignition timing was highly affected by dropping in-cylinder Żółtowski [13] studied combustion of pure DTPO with DF-like
temperature [16]. physical properties in a generator set engine incorporating a legacy
The picture that emerges after analysis of the above studies mechanical injection system. The injection timing was set in such
shows that low admixtures - up to 20% - of raw TPO in DF can a way that while working on DF, the PRR did not exceed 6 bar/°CA
be combusted in diesel engines without any modifications. SOC is throughout a range of engine loads from 1.6 to 4.5 bar IMEP.
distinctly delayed at higher TPO concentrations, so most authors Switching to DTPO, however, increased PRR significantly at ele-
advise advancing injection timing to compensate for increased ig- vated loads. The heat release analysis revealed that physical prop-
nition delay. However, this technique also increases mixing time, erties of DTPO shift the combustion towards a more premixed
thus moving combustion more towards the premixed mode. characteristic, manifesting by higher heat release rates (refer to
Using a single-cylinder diesel engine with mechanical injection, Fig. 15 for details). The effect increased with the fuelling rate
Van de Beld et al. [207] found that it would be possible to com- because the auto-ignition delay increased, elongating the mixing
bust crude TPO provided that inlet-air temperature was increased time. Although the applied load range in this study is much below
to 100-120°C with an engine compression ratio of 17.6:1. Further- that of contemporary diesel engines, the observed effects can shed
more, if the compression ratio is increased to 22.4:1, the inlet-air some light on emission trends discussed below.
temperature could be reduced by approximately 40°C. The cited Murugan et al. [47,174] performed engine tests with fuels con-
authors also underlined that TPO’s high acidity makes it necessary taining large amounts of DTPO, up to 90%. Their experiments were

26
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 15. Heat release rates for DTPO and DF. Data acquired at 1500 rpm, variable loads expressed by IMEP. Reproduced from [13] with permission from Journal of KONES
Powertrain and Transport, European Science Society of Powertrain and Transport KONES Poland.

Fig. 16. Heat release rates for DF and mixtures of DTPO with DF. Data acquired at
1500 rpm, full load, naturally aspirated, single low-pressure injection, SOI = 23°CA
bTDC. Reproduced from [47] with permission from Fuel Processing Technology, El- Fig. 17. Heat release rates for DTPO and DF. Data acquired at 2400 rpm, full load,
sevier. turbocharged, single injection at 65 MPa, SOI = 1.1°CA bTDC for DF, SOI = 1.5°CA
bTDC for TPO, intake temperature 170°C woIC (without intercooler) and 70°C wIC
(with intercooler). Reproduced from [105] with permission from Fuel, Elsevier.
performed on the basic engine platform and under the same con-
ditions as described in the previous subsection, for crude TPO
[173]. Note that the viscosity of the DTPO used was half that of DF DF and its mixture containing 5% TPO. The combustion analysis
and approximately half that used in studies mentioned in the two showed only a negligible increase in combustion duration when
previous paragraphs ([13,208]). Increasing the DTPO content to 90% using the TPO admixture.
delayed the initial pre-combustion phase (including cold flames Vihar et al. [105] used a medium-duty, turbocharged and inter-
and negative temperature coefficient reactions) by approximately cooled engine but with a mechanical fuel-injection system. It was
2.5°CA This phase is characterised by a minor heat release rate, run at 1500 rpm and 2400 rpm at various loads. The engine was
visible in Fig. 16 between -30 and -25°CA. The high-temperature fuelled with 100% DTPO, distilled to achieve diesel-like fractions.
combustion, starting with a characteristic peak resulting from the DF served as a benchmark. To compensate for TPO’s lower CN, the
premixed phase, remained similar for 90% DTPO and for DF, as ap- authors experimented with by-passing the intercooler when fu-
parent in Fig. 16. The discrepancy between this research and other elling with DTPO, thus raising the temperature of the intake air to
results discussed in this subsection can be attributed to use of dif- help auto-ignition. The results are depicted in Fig. 17, showing how
ferent fractions of DTPO, which had different physical properties. It intake temperature compensated for DTPO’s lower CN. By-passing
is plausible that delayed combustion observed in other works re- the intercooler increased the intake air temperature from ca. 70°C
sulted primarily from physicochemical parameters of the fuel, i.e. to ca. 170°C at full load and rated speed, making DTPO’s combus-
lower volatility. Data relating to the DTPO blends used by Muru- tion rate the same as that of DF. However, at lower speeds, al-
gan et al. [174] are summarised in Table 7. though the increase of intake temperature provided the same auto-
Tests of TPO-derived fuels and DF on automotive-grade engine ignition timing, the DTPOś combustion was shifted towards a more
platforms premixed, kinetic mode.
The studies presented in the previous subsection were intended Baškovič et al. [14] performed tests on a light-duty automotive
primarily to understand TPO’s combustion, whereas the experi- diesel engine with EGR and CR injection system, enabling the use
ments using automotive engine platforms, reviewed in this section, of split fuel-injection. The engine was fuelled with pure TPO, but
are rather aimed at verification of TPO as a drop-in fuel. with a viscosity in accordance with EN 590. The authors manipu-
Martínez et al. [180] performed their research on a modern, lated the pilot injection timings to compensate for delayed auto-
light-duty, automotive engine equipped with turbocharging and ignition and the resulting severe HRR of TPO. Diesel-like combus-
EGR. It also had common-rail (CR) fuel-injection, enabling the au- tion of pure TPO was achieved by retarding pilot injection by a
thors to apply split fuel-injection. The engine was operated in the few degrees in comparison with the engine’s original control map.
steady states used in the new European drive cycle (NEDC), repro- Moreover, the higher the engine load, the larger pilot injection re-
ducing loads typical of city driving. The engine was fuelled with tardation was required.

27
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 18. Heat release rates for variable admixture of TPO to JME, with DF and pure
JME as references. Data acquired at 1500 rpm, full load, naturally aspirated, single
low-pressure injection, SOI = 23°CA bTDC. Reproduced from [105] with permission
from Fuel, Elsevier.

Tests with TPO and DTPO blended with other fuels


Sharma and Murugan [201] tested different blends of TPO and Fig. 19. Regimes of soot and NOX formation on the flame fuel-air ratio and flame
JME and DF on a basic engine platform with a single-cylinder, non- temperature. Redrawn basing on the data in [209].
EGR, mechanical injection with a three-hole injector and a com-
pression ratio of 17.5:1. The engine was run at 1500 rpm and at
various loads. SOI timing was set constant at 23°CA bTDC. Pure 4.2. Emissions
JME exhibited shorter auto-ignition delay under these conditions
than the reference DF, due to JME’s higher CN. An admixture of This section considers the validations of TPO-based fuels in
TPO, with cetane number lower than DF’s, compensated for this end-use engine performance from the perspective of emissions. It
combustion advance. Adding 20% of TPO in JME gave auto-ignition aims to shed light on the potential and limitations of using TPO as
1°CA earlier than with DF. Adding 30% of TPO in JME gave auto- an alternative to conventional diesel. As a preliminary remark, it
ignition 1°CA later than with DF. These results suggest that it could should be noted that fuel-related exhaust emission trends depend
be assumed that JME with 25% admixture of TPO would provide strongly upon (i) a fuel’s physical properties, which determine in-
auto-ignition properties the same as DF’s. However, differences in jection, atomisation, mixture formation etc., and (ii) a fuel’s chem-
viscosity mean that burn durations and HRR profiles differ, despite ical properties, which determine auto-ignition timing and combus-
the same auto-ignition timings. So, although JME ignited earlier tion chemistry itself. Therefore, both physical and chemical param-
than DF, slower vaporisation extended JME’s combustion more to- eters should be considered when discussing emissions.
wards diffusion mode. Addition of TPO, which also acted as a com- The most problematic species in diesel exhaust are particulate
bustion retarder, resulted in a substantial extension of combus- matter (PM) and NOX . A temperature/fuel-air ratio map (Fig. 19)
tion duration. These effects are clearly visible in Fig. 18, compar- helps us to physically-wise understand the paths of their cre-
ing HRR curves for different fuels and their blends. Continuation ation. It shows the soot/NOX trade-off typical of diesel-engine com-
of the above work by Sharma and Murugan was aimed at improv- bustion. Concerning NOX , their creation is firstly correlated with
ing oxidation stability of 20% TPO admixture in JME by addition of peak combustion temperature (thermal mechanism) and secondly,
phenolic antioxidants, and at verification of antioxidantséffects on with higher local excess air, supporting oxidiser entrainment. Thus,
combustion [17]. The engine’s compression ratio was raised from more premixed combustion generally tends to produce more NOX .
17.5:1 to 18.5:1, aiming to tailor the engine parameters to the new Alternatively, less premixed fuel and limited oxidiser access gener-
fuel. Some delay of combustion was noted when antioxidants were ates more soot, because combustion operates with a local oxygen
added, but not to a significant degree. deficiency. Finally, it should be noted that although the combustion
Ambrosewicz-Walacik et al. [39] distilled light TPO fraction to process always runs along the reaction line presented in Fig. 19, it
produce a low-viscosity fuel component. This DTPO was then used is the amount of fuel that undergoes combustion along the indi-
to create a ternary blend of 5% DTPO, 40% DF and 55% raw rape- vidual mechanism that determines the emissions.
seed oil. The tests were performed using an automotive diesel-
engine with CR fuel-injection. The tests covered the whole oper- 4.2.1. Nitrogen oxides
ational range and included split fuel-injection at low speeds and There are recognised paths of NOX formation during combus-
single injection at high speeds. Although the engine was operated tion. The three predominant mechanisms are: Zeldovich (thermal),
using standard calibration, the detailed HRR analyses showed some Fenimore (prompt) and N2 O (fuel). The first and the last of these
specific combustion behaviour of the ternary mixture. Namely, al- are the prevailing mechanisms in diesel engines [188,210]. Most
though at split fuel-injection the HRR traces for the mixture were NOX in raw diesel exhaust gas is NO, produced mainly in the
similar to DF’s, clear low-temperature combustion of pilot fuel was boundary layer of the diffusion flame, where the excess air ratio
observed in the case of the ternary mixture. The research was con- is nearly stoichiometric, providing favourable temperature (above
tinued by Mikulski et al. [18] using different blends of the above- 1800 K) and access to oxygen. Fuel’s effects on NOX emissions from
mentioned fuels. All the tested blends included 5% DTPO but had diesel engines were extensively investigated when researching var-
differing quantities of rapeseed oil, ranging from 20% to 55%: DF ious biofuels and waste fuels. Generally, in the case of the first
made up the remainder of each blend. The HRR curves for split generation biofuels (fatty acid esters), much engine research indi-
fuel injection did not show any significant differences between all cated that they led to increased NOX emissions when compared to
tested blends and pure DF. However, the low-temperature combus- DF [211]. In the case of paraffinic fuels, NOX emissions were re-
tion fraction was increasing as the biofuel share rose. ported to be lower [3,211]. The review by Bergthorson and Thom-

28
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 20. Trends in NOX concentrations in exhaust gas for DF and its variable mix-
tures with DTPO. Research on a basic, naturally aspirated engine with single low-
pressure injection, without EGR, full load at variable engine speeds. Reproduced
from [15] with permission from Fuel, Elsevier.

son [211] revealed that NOX emissions from pyrolytic fuels are
Fig. 21. Emissions of NOX for DTPO and DF. Data acquired at 1500 rpm and vari-
roughly half of that from DF. However, this conclusion was based able loads expressed by IMEP. Reproduced from [13] with permission from Journal
mostly on research of liquids obtained from biomass, mainly wood. of KONES Powertrain and Transport, European Science Society of Powertrain and
Tests with crude TPO on basic engine platforms Transport KONES Poland.
Murugan et al. [173] performed engine tests without EGR and
with single injection. They found that increasing the TPO content the other hand, at high loads, a severe increase in peak HRR greatly
in DF from 10% to 50% reduced NOX concentration, but only at affects the in-cylinder temperature.
low loads: NOX concentration increased at higher loads. This non- Comparison of emission performance of distilled and crude TPO
monotonic trend was explained by the fuel’s effect on a combi- from the same feedstock was performed by Murugan et al. [212].
nation of local stoichiometry and temperature. Additionally, it was They tested DF and its blends containing 30% of TPO or DTPO. They
pointed out that a higher aromatics content in TPO can increase showed that admixture of TPO did not change emission levels over
propensity to form NOX . Observed trends in NOX concentrations the whole load range whereas admixture of DTPO reduced NOX
were correlated with trends in exhaust temperature, which sug- emissions. Moreover, the difference increased in line with engine
gests predominating effects of fuel on global stoichiometry. The load, with the NOX reduction growing from ca. 10% at near-idle
trends at low loads were confirmed by Chwist et al. [16]. They conditions to ca. 20% at full load.
achieved ca. 20% reduction of NOX emissions when the fuel was Recently, Uslu [213] has attempted to optimise a blend consist-
changed from neat DF to 20% TPO/DF mixture. Interestingly, NOx ing of low viscosity TPO and DF, using legacy engine experiments.
emissions from 100% TPO were double that of neat DF. Combustion The optimised parameters were the blendś proportion of TPO and
of pure TPO was heavily delayed (refer to Fig. 14), thus giving a the injection pressure. It was found that increasing TPO admix-
higher premixed fraction and higher bulk combustion temperature. ture content above 30% led to a rapid increase in NOX emissions,
Tests with refined TPO on basic engine platforms whereas TPO’s effect on NOX was negligible when admixture was
Rising trends in NOx emissions were not consistent with in- below 30%. As could be forecast, NOX emissions fell with reduced
creasing engine loads when using mixtures containing 20%, 80% injection pressure, although other emission factors deteriorated.
and 90% of DTPO. NOX emissions reduced for higher shares of Tests with TPO-derived fuels on automotive engines
DTPO at high loads, whereas there was no significant fuel effect Martínez et al. [180] conducted testing using a multi-cylinder
observed at low loads [47,174]. The discrepancy between these re- Euro 5–certified (EGR) engine, exploring different operating points.
sults indicated that emissions were affected by the reduction in He found that 5% admixture of TPO slightly increased engine-out
viscosity (reduced by about 50% via distillation) and the removal NOX emissions for all operating conditions, except low loads. This
of heavy hydrocarbons. Similar results were achieved by Frigo et al. result contradicted others using similar platforms. Bodisco et al.
[179] on a similar engine platform and using TPO with similar vis- [214] used 10% admixture of TPO in DF to fuel a Euro 5 passenger
cosity. Both 20% and 40% TPO samples provided a reduction of NOX car with a modern common-rail fuel-injection system, but without
emissions at full load, whereas results were inconclusive at the a NOX trap. The tests were realised in real driving conditions which
medium and low loads. included equal proportions of urban and rural driving. The authors
Doğan et al. [15] used a basic engine platform with mechani- showed no significant changes to NOX emissions from fuel to fuel
cal injection system and a fuel dose fixed to the maximum torque. across the whole cycle, but noted more significant differences at
They found that increasing DTPO content in DF reduces excess air. particular load/speed combinations.
NOX concentration increased with DTPO admixture for all investi- Baškovič et al. [14] used pure DTPO to fuel an automotive
gated engine speeds (Fig. 20). Additionally, the trend was not linear common-rail, direct-injection engine, advancing the pilot injection
at low speed (1400 rpm), with rapid increase of NOX concentration to achieve nearly the same combustion course as when using DF.
at the highest (90%) admixture of DTPO. Comparison of NOX concentrations at similar HRR profiles showed
Żółtowski [13] compared DF with neat DTPO and found that that DTPO increases emissions by ca. 20%, both at low-load and
NOX emissions from both fuels were comparable at partial loads. high-load conditions, as shown in Fig. 22. After eliminating the
However, DTPO reduced NOX emissions by ca. 25% at near-idle combustion rate effect, it was concluded that DTPO produces more
regime, but elevated them when operating at high engine-load, NOX due to its fuel-bound nitrogen.
as shown in Fig. 21. Aside from the fact that the NOX levels re- Tests of TPO blends with other fuels
ported in this study were generally very high (due to the nature of Using mixtures of TPO and JME, Sharma and Murugan
the legacy engine platform), the observed trends can be explained [201] identified clear NOX emissions trends, importantly the same
by analysing the HRR curves in Fig. 15. For low engine-loads, in- for various engine loads (Fig. 23). Combustion of pure JME pro-
creased auto-ignition delay did not result in increased peak HRR, duced ca. 20% more NO than when fuelling with DF. However, in-
yet retarded combustion reduced peak in-cylinder temperature. On creasing TPO admixture from 10% to 50% served to reduce NOX

29
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 22. Comparison of NOX concentrations in the exhaust gas for DF and DTPO. Re-
search on an automotive common-rail engine with split injection, without exhaust
aftertreatment; 1500 rpm at variable engine loads (80 Nm ≈ 6.4 bar BMEP, 140 Nm
≈ 11.3 bar BMEP). TPO_SH denotes TPO with combustion optimisation by adjusting
pilot fuel injection. Reproduced from [14] with permission from Energy, Elsevier.

Fig. 24. PM emissions including non-volatile and volatile fractions for DF and 5%
TPO admixture. Research on an automotive turbocharged common-rail engine at
selected operating points from a stationary NEDC test (78 Nm ≈ 4.9 bar BMEP, 163
Nm ≈ 10.3 bar BMEP). Reproduced from [180] with permission from Fuel, Elsevier.

to the production of more PM than is the case with DF. Addi-


tionally, it is worth noting that NOX and PM emissions usually
are subjected to a trade-off when combustion shifts between pre-
mixed and diffusion-controlled. More premixed combustion pro-
duces more NOX , whereas diffusion-controlled combustion pro-
duces more PM. Finally, sulphur’s effect on PM creation cannot
be neglected when considering TPO. The sulphur content increases
the mass of sulphates bound in PM, although this effect is not sub-
Fig. 23. Trends in NO emissions for DF and variable admixtures of TPO in JME. Re- stantial, because most of the sulphur is converted to SO2 [216].
search on basic, naturally aspirated engine with single low-pressure injection, with- The majority of available research uses measurement of exhaust
out EGR. The brake power from 1.1 to 4.4 kW at 1500 rpm corresponds to BMEP
from 1.3 bar to 5.3 bar. Reproduced from [201] with permission from Fuel, Elsevier.
opacity as an indicator of PM concentration. These two parameters
are always correlated in CI engines and enable comparative anal-
yses, although opacity does not provide quantitative information
emissions, compensating for JME’s effect. These trends could be about PM emissions.
attributed to the opposite physical and auto-ignition properties of Tests with TPO/DTPO on basic engine platforms
biofuel and TPO. Biofuels ignite earlier but have a higher boiling The trends in PM and smoke emissions on engines with single-
temperature, which contributes to mixing-controlled combustion. pulse injection and in non-EGR conditions are summarised below.
TPO has a lower CN, and thus ignites later, but it contains light Murugan et al. [173] used blends of DF with TPO admixtures rang-
fractions which evaporate earlier. Overall, and under most condi- ing from 10% to 50%. They did not report any significant changes in
tions, emissions from combustion of a 50% TPO and 50% JME blend exhaust opacity across the whole load range, except at maximum
were a few percent higher than DF’s emissions. power, where 50% addition of TPO increased opacity by 7%. How-
Mikulski et al. [18] used a light fraction of DTPO to reduce vis- ever, the 40% TPO addition slightly reduced opacity compared with
cosity of raw vegetable oil. The test fuels were composed of 5% DF. In other works [47,174], the same research group found that
DTPO and from 30% to 55% of rapeseed oil, with DF making up DTPO produces more smoke than DF and they recorded that in-
the remainder. Tests were performed using a common-rail automo- creasing the proportion of DTPO admixture increases exhaust opac-
tive engine with standard calibration. There were no differences in ity. Moreover, comparison of emissions from TPO and DTPO 30%
NOX emissions between different fuels and the reference DF at low admixtures showed that TPO did not affect smoke emissions across
loads, but a clear trend was evident at higher loads. The higher the the whole range of engine loads. At the same time, admixture of
rapeseed oil content, the higher were the observed emissions. At DTPO increased smoke emission by ca. 20%, but only at higher
the 80% load, the addition of DTPO to the rapeseed oil/diesel mix- engine-loads. Fuel effects on PM and smoke emissions were negli-
ture was responsible for higher NOX emissions than a comparable gible or at least inconclusive at low loads.
mixture of FAME and diesel. Tests with TPO-derived fuels on automotive engines
Martínez et al. [180] performed detailed analysis of smoke and
4.2.2. Particulate matter and smoke opacity PM emissions from an automotive engine fuelled with a 5% TPO
PM consists of insoluble and soluble fractions. Soot and ash blend in DF. At different points of the engine operating map both
are components of the insoluble fraction; whereas the soluble one opacity and PM emissions were higher with this blend than with
consists of organic carbon, sulphate and nitrate [215]. Soot, which only DF, as shown in Fig. 24. This was attributed to TPO’s higher
is a precursor of PM, is intensively created in the first stage of aromatics content and also to its poorer vaporisation, due to higher
combustion, typically at the tip of the fuel spray, where a rich mix- final boiling point. The measurements of PM emissions were sup-
ture is burnt volumetrically and then turns into mixing-controlled ported by analysis of particle size distribution. It was found that
combustion. The production of PM is chemically increased by the the size distributions of both tested fuels were the same at low
presence of alkynes and polycyclic aromatic hydrocarbons (PAH), loads, whereas under higher loads the distributions for fuel con-
so TPO’s poorer atomisation and high PAH content should lead taining TPO shifted towards smaller particles. This was ascribed

30
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

combustion. The second dominating mechanism, more likely for


diesel engines, is low flame temperature. This can occur when
combustion is delayed towards the expansion stroke or if the fuel
is highly diluted, by recirculated exhaust, for example. CO and
HC emissions trends for TPO-derived fuels and mixtures are sum-
marised here.
Working on a simple engine platform, Murugan et al. observed
that both CO and HC concentrations generally rise with increasing
TPO admixture to DF, independently of engine load [47,173,174].
The increase in CO with addition of TPO was attributed solely
to delayed combustion and thus inhibited CO oxidation. Similarly,
Kumar Singh and Prabu [208] reported an increase of concentra-
tions of both components. The trends in both works appear con-
sistent. Shifting fuel from DF to pure TPO almost doubled CO con-
Fig. 25. Trends in exhaust gas opacity for DF and variable admixtures of TPO in centration, while the increase in HC content for the same transi-
JME. Research on basic, naturally aspirated engine with single low-pressure injec- tion would be approximately 50%. Examples of CO and HC emis-
tion, without EGR. The brake power from 1.1 to 4.4 kW at 1500 rpm corresponds
sion trends from the work by Żółtowski [13] are shown in Fig. 26.
to BMEP from 1.3 bar to 5.3 bar. Reproduced from [201] with permission from Fuel,
Elsevier.
In contrast, Chwist et al. reported a decrease of CO concentration
with increasing TPO content [16]. However, it should be noted that
the baseline CO concentrations reported in this work were already
to TPO’s sulphur and metals contents. Analysis of PM composition ultra-high, at 3%. The trend in HC emissions was not monotonic,
showed that PM generated by TPO contains a higher volatile or- but indicated an increase in the exhaust’s HC content when 10% or
ganic fraction (Fig. 24). more TPO was added to DF. Doǧan et al. [15], using a naturally as-
Tests of TPO blends with other fuels pirated engine at full load, found that the CO and HC trends were
Measurements of exhaust opacity by Sharma and Murugan dependent on engine speed. At low speed (1400 rpm), using a TPO
[201] when using blends of TPO with JME show a clear trade-off content of 90% reduced CO by 50% and HC by 20%, when com-
with NOX emissions. Compare Figs. 23 and 25 for details. Combus- pared to DF. Results were inconclusive at engine speeds beyond
tion of pure JME reduced opacity in comparison with DF. The re- this point.
duction was substantial - approximately 30% - at higher loads. In- In the case of TPO mixtures with biofuels, Sharma and Murugan
creasing TPO content in JME from 10% to 50% gradually increased [201] observed that emissions of both CO and HC when using pure
exhaust opacity to a level slightly above that when fuelling with JME are about 10-20% lower than with DF. Addition of TPO to JME
DF. Opacity with 20%-30% admixtures of TPO in JME was compara- increased both emissions. The CO and HC emissions were similar to
ble to that with DF. those from DF when the TPO admixture to JME reached about 30%.
Recent works by Ambrosewicz-Walacik et al. [39] and Mikulski Using a more advanced automotive engine platform, Martínez
et al. [18] were aimed at making it feasible to use raw bio-oils in et al. [180] evaluated emissions from a 5% admixture of TPO to
diesel engines, employing light fractions of DTPO to reduce viscos- DF. They found the admixture increased emissions of both CO and
ity. Addition of 5% of low-viscosity DTPO to a mixture containing HC by approximately 10% compared to DF, but only at near-idle
40% of DF and 55% of bio-oil produced some reduction of exhaust conditions. Differences between the test fuels were negligible at
opacity. However, a dramatic increase in smoke was observed at moderate and high loads. Furthermore, increased emissions of aro-
higher engine-loads and with split (pilot and main) injection [39]. matics were reported under all conditions and attributed to higher
This increase was found to be correlated with combustion rates, aromatic content in TPO. Similar results in terms of HC emissions
specifically advance of auto-ignition. Coupled with the poor atomi- were achieved by Baškovič et al. [14], although they compared DF
sation of highly viscous fuel, this resulted in rich combustion, pro- with pure TPO. These results revealed sensitivity to EGR of low-CN
ducing large amounts of soot. Interestingly, the fuel’s effect on fuels, like TPO.
smoke was much smaller with a single injection of fuel, because
the fuel had less of an effect on combustion rates. Finally, it has 5. A critical overview and outlook on further research
been found that injection strategy strongly affects the emissions
response to fuel type. Further research with the same 5% DTPO ad- The last decade has brought an exponential increase of research
mixture, but with variable amounts of bio-oil and DF, showed that, into the use of TPO as a combustion engine fuel. However, the un-
in general, ternary fuel reduces exhaust opacity using both single clear status of TPO production technology, wide discrepancy in TPO
and split fuel-injection. However, this effect is most noticeable at quality, the different blending strategies and the variety of engine
moderate load. Moreover, the proportion of bio-oil content did not platforms used for testing all make it difficult to pin down the cur-
show any clear trend in opacity. rent status of the research. These factors have been reviewed in
Sections 2, 3 and 4. This section’s aim is to distil the knowledge
4.2.3. CO and unburnt hydrocarbons captured in this review into a concise vision of TPO as a fuel com-
Emissions of CO and HC from diesel engines are far less prob- ponent in combustion engines. This is best done by answering the
lematic than NOX and PM [217]. The amounts are small and research questions posed in Section 1.3 of the introduction.
these species can be burnt effectively in oxidation catalytic reac- The overview and outlook below provide a focused perspective
tors, provided that the exhaust temperature is high enough [218]. on applications for TPO-derived fuels and further research in this
Nevertheless, TPO-derived fuels also affect these emissions. This field. This research is most definitely needed if TPO fuels are to be-
is mainly due to differences in the physical parameters of fu- come part of a solution capable of solving the sustainability issue
els. Highly viscous fuels have higher liquid fuel spray penetration, and GHG problem.
which can result in wall impingement by fuel. This phenomenon What is the current view on TPO well-to-tank environmental
usually produces high amounts of HC because not all the fuel is impact and economic feasibility?
burnt during combustion. CO is mainly a residue from incomplete On a well-to-tank basis, TPO as a waste-derived fuel is consid-
combustion if the mixture is too rich to oxidise CO at the end of ered as one of the negative GHG impact solutions. Taking account

31
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Fig. 26. Emissions of CO and HC for DTPO and DF. Data acquired at 1500 rpm and variable loads expressed by IMEP. Reproduced from [13] with permission from Journal of
KONES Powertrain and Transport, European Science Society of Powertrain and Transport KONES Poland.

of the added value to the circular economy, the net GHG footprint the number of works available on the topic is lacking. In fact, it
is assessed on a level of -0.9 to -1.6 kgCO2 e/kg of CO2 removed is the reliability of the studies, which in many cases are based on
from the environment [138,219]. On the gross basis, the well-to- biased assumptions, that has been found to be questionable.
tank CO2 emission of TPO production is of an order of magnitude What is the progress in pyrolysis technology with respect to
smaller than that of first-generation biofuels. In this respect, TPO maximising the TPO yield and resulting fuel quality?
surpasses both SVO and FAME/HVO derived from bio-oils. Taking The review has shown a rapid emergence of research-grade,
into account engine compatibility issues and the extent of process- pilot-scale tyre pyrolysis plants. The demonstrators are gradually
ing required to satisfy this compatibility, TPO is also GHG- superior moving towards more advanced designs, including whole tyre py-
with respect to popular waste-derived biofuels (TME, UCOME). The rolysis [75]. Process temperature is considered the most critical,
present study implies gross well-to-tank GHG emission of TPO on and the most investigated, optimisation parameter here. The best
an average level of 3.76 g CO2 e/MJ, a value comparable to bio-oil results in terms of TPO yield, calorific value and minimised energy
extraction alone. input typically are reported to come in the temperature range of
The increasing social impact of GHG emission has a pronounced 450-650°C. Further elevation of the temperature not only demands
influence on the fuel market. As shown in this study, currently greater energy input but the intensified rubber cracking results in
most waste-derived fuels and feedstock command market prices more high-volatility hydrocarbons transferred to the gaseous frac-
that are similar to or higher than first-generation SVOs and FAMEs. tion, reducing the TPO yield. Furthermore, elevated temperatures
TPO still stands apart from this trend. With a stable average market lead to undesirable increases in TPO’s viscosity and density, as well
price of $350/10 0 0 kg – a figure almost unchanged over 10 years as its sulphur and aromatics content. Turning to TPO yield, the
– it is roughly one third of the price of all the mentioned biofuels most frequently reported maximum TPO yield is 50%wt. (based on
and it is price competitive with gasoil. This creates an opportunity median from Table 2 / Table A1) but the state of the art is cur-
for TPO to be accommodated by markets that traditionally cannot rently reported to be up to 60% on the mass basis, when coupled
afford expensive fuels and thus struggle to compete with the likes with high heating values (based on median from Table 6: HHV = 43
of automotive and aviation sectors when seeking sustainable al- MJ/kg, LHV = 41 MJ/kg).
ternative fuels. Recent analysis shows that these two markets cur- Notably, even without specific reactor optimisation, the CN of
rently account for over 90% of all liquid biofuel availability [220]. TPO meets the US on-road fuel standards as well as all standards
It must be acknowledged that TPOś low market price is partly a for off-road and marine fuels. Achieving viscosity and density val-
consequence of a relatively low global demand for this fuel. If this ues that are within EN 590 limits does not pose a challenge, yet
is about to change in the current energy transition period, cou- must overcome the trade-off with the fuel’s flashpoint. TPOś high,
pled with scaling-up the production, the future price-availability diesel-like oxidative stability is another of its quality advantages
balance of TPO is still unknown. as fuel. Crude TPOś properties do not deteriorate significantly with
Adding to the list of uncertainties is the overall environmen- time, which is a common issue for most biodiesels. Hydrolysis and
tal and economic feasibility of TPO production. Though available microbial growth do not pose additional problems for TPO, assum-
LCA studies on TPO indicate that pyrolysis is currently one of the ing the feedstock or end product are properly dehydrated.
ecological and net beneficial ELT management routes, we currently However, there are fuel quality challenges for TPO, notably high
lack any reliable comparative LCA with other sustainable fuel al- sulphur and polyaromatic hydrocarbons contents and potential
ternatives. This is a serious knowledge gap. The authors´realistic acidity issues. In the light of tightening global legislation, sulphur
view of the overall financial feasibility of fuel production-oriented content is by far the biggest and most immediate challenge for di-
tyre pyrolysis is inconclusive because of the uncertain future of rect applications of TPO as an automotive fuel. Typical sulphur val-
TPO’s market price, unfavourable ELT management legislation and ues for crude TPO produced in an optimised reactor are in a range
the lack of up-to-date, quality studies on the economics of large- of 0.3 -1.0%, although figures below 0.01% were also noted for hy-
scale pyrolytic plants. It should be noted that the common opinion drogenative pyrolysis. This means desulphurisation is vital for any
that tyre pyrolysis is not commercially feasible mainly stems from automotive application where EN 590’s 0.001% sulphur limit ap-
several big implementation projects from the 1980s and 1990s. plies, even if blends are considered. TPO’s route to marine applica-
Against the backdrop of the current energy transition, this view tions looks easier, with the IMO global and ECA sulphur caps set
has to be reconsidered. at 0.5% and 0.1% respectively. The high content of PAHs poses a
As a final critical note, it is fair to point out that the mentioned challenge due to genotoxicity and carcinogenicity, as well as the
knowledge gap on TPO’s LCA and economics does not mean that propensity to form soot. This issue, as well as the sulphur content

32
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

and the flash-point/viscosity trade-off, can be partially solved by lation of TPO, yielding DTPO, generally offers regulation of viscosity
TPO-upgrading, as summarised in the next subsection. and cold properties. Accurate fractionation supports the removal of
In terms of TPO acidity, the results so far are inconclusive. PAHs, which have high boiling points. The desulphurisation effects
Reports vary from diesel-like acidity to values that are 10 times achieved by distillation vary from study to study. Application of
higher than the current automotive standards. Obviously, acidity, ionic solutions for liquid-liquid extraction was also demonstrated,
as well as other chemical properties of crude TPO, depends on showing potential to halve sulphur content on average. Bearing
pyrolysis conditions, yet this dependency has not been sufficiently in mind that the addition of an in-reactor catalyst helps to opti-
studied. mise pyrolysis, and that a TPO sulphur level below 0.01% has al-
This brings us to the general knowledge gap in the field ready been achieved, further progress towards TPO fuel applica-
of tyre pyrolysis technology. An observation from this review is tions seems feasible, if fuel blending is considered.
that most applied researchers studying how best to produce TPO Ultimately, providing a direct quantitative answer to the ques-
are focusing on a single process parameter, such as temperature, tion regarding achievable fuel quality by application of different
without considering the underlying mechanisms and parameter upgrading technologies seems impossible at this moment. This per-
cross-influences. Equally, there are many excellent fundamental- tains to the knowledge gap similar to one related to TPO produc-
level studies on mechanisms of tyre pyrolysis which lack the ap- tion, described in the previous subsection. Namely, the research di-
plied perspective towards optimising TPO quality as fuel. Notably, rected to optimising engine-related fuel properties is so far, in the
achieving optimum TPO yield and composition depends not only great majority of cases, performed without a basic chemical back-
on process temperature and on reactor type and its mechanical pa- ground and analytical support. In contrast, the fundamental-level
rameters, but also on rubber particle size (from powder to whole research on TPO fractional analysis lacks the targeted context of
tyres), inert gas flow and many other factors. On a more funda- achieving superior engine-fuel properties. For example, trade-offs
mental level, all these variables come down to the mechanisms between viscosity and flashpoint implied by some researchers for
of transient heat transfer to the rubber elements, gas diffusion DTPO can be avoided if distillation curve analysis is used to per-
through the solid material and gas retention time before cool- form targeted fractionation.
ing and condensation. Therefore, elaboration of thermodynamic What is TPOs potential as a fuel component?
and gas flow models, combined with solid decomposition and gas Although progress in desulphurisation techniques and develop-
chemical kinetics would enable understanding of the discrepan- ments in pyrolysis process control is improving its quality, TPO’s
cies between experimental studies, and thus reduce global applied most feasible application is probably not as a stand-alone fuel.
research effort. As well as the need to connect fundamental and Blends are the favoured route, not only because they facilitate the
applied-level research, comes a need for research to investigate improvement of the end-fuelś properties, but also from the per-
multi-parameter optimisation of reactor operation. spective of availability. No single alternative fuel, such as waste fu-
Finally, among the open research fronts is investigation of co- els, biofuels or synthetic electro-fuels, will be capable of becoming
pyrolysis of tyres with biomass. This is a relatively new con- a full substitute for all fossil diesel. Blending with diesel or heavy
cept, so the number of available works on this topic is limited fuel oils is the obvious strategy here.
and scattered over different feedstock configurations. The syner- A promising route for TPO-derived fuels is to take advantage
gistic effects seem evident, yet not sufficiently understood. The of their complementary properties with biofuels. Biofuels usually
research trajectory, however, seems very important from the per- suffer from high viscosity (FAME) or poor lubricity (HVO) yet offer
spective of achieving high-quality fuel, increasing the oil yield and, substantial reserves in terms of flashpoint and CN. TPO/DTPO has
at the same time, tackling the availability limitations of sustain- the opposite characteristics. Another potential route for non-road
able/renewable feedstocks. applications of DTPO, especially in developing countries, might be
What are the available TPO-upgrading technologies and what is to use its light fractions as an effective viscosity improver when
their effect on the end-fuel quality? raw vegetable oils are admixed to diesel. This was demonstrated
Assuming the pyrolysis itself is fairly well optimised, crude TPO by Mikulski et al. [18]. It was also pointed out in the literature
has the potential to fulfil the requirements of marine and of-road [17] that TPO can offer an additional benefit in biofuel blends be-
fuel standards. Automotive fuel regulations can be met, except in cause it improves storage stability when mixed with FAME.
terms of the aforementioned viscosity/flashpoint trade-off and the Fuel blending is at the bridge of fuel-engine research and the
sulphur content. Thus, when starting with a relatively high qual- major issue with the current situation comes from the fact that
ity of crude oil, it is probably not economically feasible to ap- research in this field is mainly performed by applied engine re-
ply refinery-grade cracking processes, such as hydrocracking, for searchers. They often try to find the right way to use TPO by pick-
instance. Indeed, those processes have not been widely demon- ing from a wide range of possible blending strategies, but inves-
strated with TPO and, according to available literature, refinery- tigate routes that could be immediately ruled out with a better
grade sulphur removal (HDS) is deemed to be hardly effective consideration of differences in chemical make-up. More complex,
when considering the constraints of automotive fuel market re- tailored blends should be explored. This fuel blend design chal-
quirements. It is worth noting that the 2016 EU-project SULFREE lenge should be supported by the development and use of pre-
attempted to combine state-of-the-art microwave pyrolysis with dictive models that can target specific properties or engine per-
refinery grade hydrotreating, yet failed to achieve TPO with sul- formance [222]. Development of these models should be strongly
phur levels below 0.1% [221]. encouraged in order to focus experimental testing and accelerate
Considering the above, the future TPO-upgrading technologies the development of sustainable liquid fuels: climate change does
will most probably orient towards infrastructurally low-cost solu- not allow the luxury of time.
tions, suiting the trend in distributed TPO production with low- How do different fuel compositions of TPO and TPO-derived
volume reactors with capacity to process 10 0-20 0 tonnes per day fuels perform in the engine in terms of combustion characteristics,
of ELTs. On the basis of that scenario, this review identified two efficiency and emissions?
main low-cost desulphurisation processes for TPO: fractional dis- The review of engine test results reveals several trends. Within
tillation and extraction using ionic liquids. the constraints of the quality of the TPO samples used, small ad-
Many different strategies of TPO fractionation have been tested mixtures of crude TPO (up to 10%) to diesel can be used without
while attempting to achieve fuel properties within EN 590 limits. It any profound effects on the combustion process and connected ef-
has to be said that none is completely successful. Fractional distil- ficiency, while avoiding issues with catalyst poisoning Using pro-

33
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

cessed (distilled and desulphurised) TPO can further extend this possible by adding the perspective of fuel properties and combus-
admixture limit somewhat higher than 10%. At such conditions, tion analysis. The number of works on engine emissions from TPO
NOX emissions can be slightly elevated, evident only at higher that lack this context is, in fact, far greater than those reviewed
loads, and CO and HC can increase by approximately 10%, but only in Section 4. Without a combustion perspective, the results from
at low loads. Note that these differences in state-of-the-art com- various works seem contradictory, making it impossible to draw
bustion engines are reported to disappear over the driving cycle accurate conclusions. This arouses a general concern that incom-
when using low levels of TPO additions. plete works may lead to conclusions that have a possible negative
When combustion effects are neglected, TPOs tend to emit impact on the take-up and commercialisation of alternative fuels.
more NOX due to higher amounts of fuel-bound nitrogen. Greater With this rather critical remark, we open the discussion about out-
CO and HC emissions result from higher viscosity and poorer spray look.
atomisation, limiting oxygen entrainment. TPO increases PM emis- Outlook on TPO’s perspective
sion because of its high aromatic and sulphur contents. Both these An underlying premise of this research is that waste tyres form
fractions are found to promote PM formation, which in the case a considerable feedstock for fuel production, comparable to that of
of TPO, usually is composed of small-diameter particulates. These biomass once food production has been prioritised. For instance,
smaller particles are proven to have a more deleterious impact on considering the durability of the average car tyres of 70 0 0 0 km
human health. Note, however, that reports documenting quanti- and fuel consumption at a level of 7 dm3 /100 km, a vehicleś own
tative particulate mass or number for TPO combustion are rather worn tyres can pay back 0.5% of the total fuel used.
scarce: most authors instead used smoke opacity as the basis for The evidence considered in this review indicates that TPO has
their conclusions on PM. Opacity was more likely to be influenced sufficient quality to be used directly as bunkering fuel, though is-
by local mixture stoichiometry. sues with fuel-to-fuel compatibility would have to be resolved.
The abovementioned fuel-related effects on emission can be There is a fairly small gap between the quality of fuel from the cur-
scaled with increased TPO fuel content and also can be partially rent state of the art in TPO production technology and the standard
neutralised by optimum fuel composition. Large-scale TPO admix- set for automotive diesel fuels. Continued research along the sug-
ture makes the effect of the fuelś properties on combustion be- gested lines should be capable of bridging that gap. To this end, the
come more pronounced, in turn influencing emissions to a much research must become more multidisciplinary, merging the funda-
greater extent. Delayed auto-ignition is commonly reported for mental expertise in feedstock/fuel chemistry with applied knowl-
TPO or DTPO additions above 20%, shifting combustion towards a edge on pyrolysis technology.
more premixed mode, which usually generates higher peak tem- The requirement for sustainability will shift future fuels to-
peratures, elevating NOX emissions. Long ignition delays also shift wards more complex blends accommodating all renewable re-
combustion from its optimum timing, and thus impact engine ef- sources. Feedstock co-utilisation and post-production blending
ficiency. Fuel viscosity and cetane number are primary drivers of strategies should be developed to build on the complementary
ignition delay. High viscosity is relevant only for non-processed physicochemical parameters of individual feedstocks and/or fuels.
TPO samples and harms combustion at high-load regimes where Blending TPO and FAME/UCOME have been suggested here as a
spray formation is the main limiting factor for ignition. TPO’s lower good example of this complementary approach. Another potential
cetane number is responsible for increased chemical ignition de- route is co-pyrolysis of waste tyres and biomass. According to the
lay manifesting at low loads and at high EGR ratios. Van de Beld present fuel quality and emissions review, an exemplary composi-
et al. proved that this low CN effect can be successfully mitigated tion of 7% FAME / 7% (desulphurised) TPO/DTPO / 86% diesel could
by elevating the intake-air temperature [207]. This technique al- be straightforwardly applied on the market, offering benefits over
lowed the combustion of neat TPO samples with diesel-like char- the current B7 diesel. Other combinations, for instance, blending
acteristics and efficiency. Engine control strategies have also been TPO with HVO, also should be explored. Progress here requires
demonstrated to mitigate viscosity effects for TPO, including ad- methodological reinforcement in terms of chemical expertise and
vancement of injection timing (both main and pilot injection) and predictive tools for modelling blended fuel properties and engine
increasing injection pressure. Both strategies are, however, limited performance.
by the threat of spray wall-wetting. Note that viscosity control by Before TPO-derived or any other pyrolytic fuels can be mar-
fuel pre-heating is a standard feature in flex-fuel marine diesel en- keted, engine manufacturers must accept them, which in turn,
gines capable of running on HFO. means that more vehicle and engine tests using TPO blends should
Critical evaluation of available research related to engine oper- be picked up by research organisations. The bulk of engine test
ation on TPO reveals a clear lack of reliable information on non- results with TPO involve legacy (basic) engines. Tests with mod-
regulated emissions, even though it is an important issue in the ern (EU Stage V/Euro 6) engines are lacking. Furthermore, the twin
consideration of new fuels. A notable exception is the work by threats of PM and PAH are already identified as potential issues,
Martínez et al. [180], where a multi-compound Fourier transform which in our opinion will limit the application of TPO/DTPO to
infra-red (FTIR) exhaust gas analyser was used. This study, how- blends not exceeding 10%. There are many more pending emissions
ever, used only 5% TPO admixture to DF, thus the differences in issues for TPO applications that have not been addressed at all so
emissions of most hydrocarbons, except aromatics, were below the far. Unlegislated emissions need to be considered, especially with
level of significance. Naturally, the aromatics content in exhaust respect to aldehydes and aromatics. TPO is expected to generate
gas are inherited from fuel composition. The same effect is ex- higher levels of both of these, which may lead to imposition of
pected in the case of sulphur, both in terms of gaseous exhaust additional constraints on TPO. Additionally, real driving emission
components and PMś chemical composition. There is a need for tests of vehicles fuelled with TPO should be performed. The effect
detailed emissions speciation when TPO is considered as a fuel of TPO fuels on modern exhaust aftertreatment systems has been
component. hardly studied at all. Sulphur poisoning of diesel oxidation cata-
In any observations regarding PM, it is important to point out lysts is an obvious issue here, and there may be others. Catalyst
once again that most studies considered only smoke opacity: only tests should be conducted in tandem with long-term durability en-
a few used PM concentration measurement and sizing. Refer to gine tests to check for potential deposit formation (injectors, valves
[180] for reliable information on PM distribution from TPO. etc.) and performance deterioration over time.
As a final discussion note, it is necessary to say that construct- Finally, it is worth remembering that tyre pyrolysis, although
ing this summary of emission trends related to TPO fuels was only more economically feasible in the current energy transition pe-

34
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

riod, lacks legislative incentive to encourage its large-scale up- FAME (or crude bio-oils) are most feasible since these fuels can
take as a clean and efficient way to use waste tyres. Contin- be compatible in viscosity, flash point and cold properties.
ued lobbying in this respect is required to transition from sim- ix Emissions are an important constraint for TPO fuel applica-
ple (but destructive) incineration of tyres towards more sustain- tions. Its combustion-related effects can be mitigated either by
able pyrolysis. Furthermore, the above pros and cons of pyrolytic matching fuel properties or combustion control, but TPO is fun-
oil as a fuel component should be weighed against other possi- damentally prone to emit more nitrogen oxides, CO, unburnt
ble applications for TPO, such as its use in a closed-loop, tyre- hydrocarbons and particulates than diesel fuel. However, for
manufacturing process. This comparison should include factors like typical blends containing up to 10% TPO, the effects are rather
process economy, well-to-wheel CO2 footprint and local availabil- small and also diminish at transient engine operation under
ity. It should also assess combinations of TPO with other feed- real driving conditions.
stock in fuel and tyre producton chains, demanding a complex x A major concern of using TPO is associated with particulate
decision-making equation. Multi-criteria decision-making using ar- emissions, which scale up linearly with TPO content in the fuel.
tificial intelligence-based approaches would allow complex data The size distribution of the PM, and TPO’s substantial polyaro-
from diferent sources to be merged with expert knowledge into matic hydrocarbons content, which translates to PMś chemical
reasoned and unbiased decisions [223,224]. composition, are also problematic due to their carcinogenic na-
ture.
6. Conclusions xi Despite the above issues, the major hurdle for TPO remains the
waste tyre management legislation. Regulations in this sphere
The objective of this work was to review the current status of fail to provide the necessary incentive for pyrolysis to compete
tyre pyrolytic oil in combustion engine applications. This fills an successfully with incineration. The economic and ecological fea-
important knowledge gap on the sustainability transition roadmap, sibility of TPO needs to be rigorously re-evaluated in the con-
because the vision of how to best utilise waste tyres is not clear. temporary context of energy transition.
This feedstock, with its substantial availability and good heating xii The present work puts forward the necessity of professionalisa-
value, can contribute to solving the short-term fuel transition prob- tion of research in the field of TPO. Current infrastructural and
lem for off-road machinery, heavy-duty road vehicles and marine methodological commitment to this issue is too low when com-
transport. This review’s focus was on diesel engine applications be- pared to other fuel alternatives. Future endeavours should ulti-
cause these are the prime-movers in these sectors. mately become more multidisciplinary to address the apparent
To this end, this review answers the governing research ques- disconnect between fundamental level research and TPO appli-
tions related to TPO production and quality; post-processing and cations.
quality improvement methods; and engine validation of the fi-
The above conclusions, complemented by exact numbers in the
nal TPO-derived fuels. These questions are discussed in detail in
adjoining discussion, aim to provide an unbiased input to help
Section 5. Below, we present the main conclusions drawn from the
shape the future fuel roadmap for TPO. Accordingly, these conclu-
discussion:
sions do not pre-judge the outcome of such work and so they do
i TPO as waste-derived fuel has a negative GHG impact, removing not convey a final decision. That decision-making process should
up to 1.6 kgCO2 e/kg of waste tyres from the ecosystem. At the weigh TPO against different feedstocks available for sustainable
same time, gross GHG emissions of TPO production are three fuel production, and also against possible competing applications
times lower than those of typical first-generation biofuels. of these feedstocks.
ii TPOś future availability is comparable to that of vegetable oils
and the average market price of $350/10 0 0 kg makes it the
Declaration of Competing Interest
cheapest sustainable fuel alternative, creating an incentive for
its use as an admixture to marine and off-road fuel supply
The authors declare that they have no known competing finan-
chains.
cial interests or personal relationships that could have appeared to
iii The current state of the art in tyre pyrolysis provides average
influence the work reported in this paper.
yields of crude oil of 50%, without any external energy input.
iv With nearly diesel-like lower heating value of approximately 41
MJ/kg., the properties of crude TPO technically enable direct fu- Acknowledgments
elling of compression ignition engines.
v The fuel can meet the low-sulphur fuel oil standards for The authors acknowledge expert support from Dr. Katriina
bunkering fuel and so can be used in all ships subject to the Sirviö (VEBIC Fuel Laboratory, University of Vaasa, Finland) and Dr.
IMO global sulphur cap (max 0.5% fuel sulphur content) with- Kamil Duda (Fuel Laboratory, University of Warmia and Mazury,
out needing scrubbers. The quality is insufficient for automo- Poland). Their critical views were invaluable during the prepara-
tive applications governed by the EN 590 diesel fuel standard tion of the manuscript and its revision.
or corresponding US norms. We would like to thank David Wilcox for his help in proofread-
vi New desulphurisation technology, such as the ionic solution ing the manuscript and engaging his expertise in polishing our pa-
washing process, can reduce sulphur content below 0.1%. The per.
process is not infrastructurally demanding and supports the Jacek Hunicz and Maciej Mikulski acknowledge financial sup-
economics of distributed TPO production. port from the project Lublin University of Technology-Regional Ex-
vii TPO viscosity can be tailored using distillation, but the trade- cellence Initiative, funded by the Polish Ministry of Science and
off between viscosity and flash point should be considered. Higher Education (contract no. 030/RID/2018/19).
Fractional distillation should be targeted at a narrower boiling
range than currently exploited to avoid this trade-off.
viii The limitations of viscosity, flash point, sulphur content and Appendix. Type of reactors used during pyrolysis and most
polyaromatic hydrocarbons fraction narrow down automotive relevant process conditions affecting TPO yield
applications of TPO/DTPO to low-level blends (up to 10%) with
mineral diesel and other fuels. Mixtures including DTPO and Table A1.

35
Table A1
Type of reactors used during pyrolysis and most relevant process conditions affecting TPO yield

Reactor type Reactor Process Feed material Feed Rate Inert gas / flow Temperature TPO Yield [%] Conditions for Reference Manuscript objective
volume rate range [°C] optimum TPO yield
[dm3 ]

fluidized bed 0.9 continuous shreds (1.4-2.3 mm) 0.2 - 0.5 kg/h N2 + air 700-880 25-38 700°C, feed rate 0.3 [77] Process evaluation
kg/h
n.a. n.a. ground n.a. N2 450-800 n.a. n.a. [48] Process evaluation
n.a. fast pyrolysis shreds (1-2 mm) 1 kg/h N2 /steam 500-600 51-65 500°C [78] Product analysis
n.a. n.a. ground 5 kg/h steam 360-810 30-50 450°C, 0.32 mm [79] Process evaluation
particle size, short
gas residence
n.a. n.a. shreds (0.8-1.6 mm) 3 kg/h N2 /pyro 600-700 over 36 600°C, short gas [80] Process evaluation
gas/steam residence
n.a. continuous, fast shreds (0.3-1.18 0.6 - 1.5 kg/h N2 350-600 25-45 400°C, particle size [81] Process evaluation
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

pyrolysis mm) 1 mm, feed rate


0.78 kg/h
0.23 fast, porous co-pyrolysis 0.2 kg/h N2 (0.25 400-650 n.a. n.a. [82] Product analysis
catalysts (zeolites) dm3 /min)
4.1 continuous, fast sherds (3-4 mm) n.a. N2 350 26-38 400-450°C [83] Process evaluation
pyrolysis
n.a. intermittent (4-60 ground, steel and n.a. H2 375-425 42-47 375°C, reaction [84] Product analysis
min) fabric free time 4 min.
fixed bed 1.5 intermittent (15 ground, steel free 0.05 kg/run He (0.2 350-550 30-38 550°C [85] Process evaluation
min) dm3 /min)
1.2 intermittent (60 shreds (1.0-1.4 mm) 0.2 kg/run N2 process 35-56 430°C, zeolite [65] Process evaluation
min), microporous constant 500 ZSM-5 catalyst
catalysts (zeolites) (430-600

36
catalyst)
0.4 intermittent (60 shreds (2.0 mm) 0.01 kg/run N2 (0.1 375-500 46-60 425°C [70] Process evaluation
min), perlite dm3 /min)
catalyst
3.5 intermittent (30 strips (2-3 cm 0.2 kg/run N2 (1 dm3 /min) 300-700 5-39 700°C [86] Product analysis
min) wide)
2.2 intermittent (120 shreds (1-2 mm), 0.05 kg/run N2 (2 dm3 /min) 500-800 30-38 500°C [87] Product analysis
min) steel and fabric free
0.59 intermittent (60 shreds (1.5-2 mm), 0.13 kg/run N2 (0.025 550-800 47-56 n.a. [88] Process evaluation
min) steel and fabric free dm3 /min)
0.95 intermittent (120 shreds (20 mm) 0.3 kg/run N2 (0.4 400-700 30-43 500°C [51] Product analysis
min) dm3 /min)
0.004 intermittent (15 ground 0.001 kg/run N2 (0.15 400-600 36-45 n.a. [89] Product analysis
min) dm3 /min)
2.1 intermittent (50 cross section 0.8 kg/run N2 (2 dm3 /min) 375-575 up to 49 475°C [90] Fuel production
min)
0.08 intermittent shreds (<3 mm) 0.01 kg/run n.a. 850 20-58 n.a. [91] Process evaluation
1.7 n.a. shreds (1-2 mm) n.a. N2 (5 dm3 /min) 500-800 33-41 n.a. [92] Product analysis
6.7 intermittent (160 shreds (10-50 mm), n.a. N2 330-380 38-88 n.a. [93] Fuel production
min) steel and fabric free
n.a. intermittent (20 shreds (1-1.6 mm), n.a. N2 (0.5- 1.5 400-500 41-53 450°C [94] Fuel production
min) steel and fabric free dm3 /min)
2.8 intermittent (15 shreds 0.25 kg/run n.a. 400 25-55 n.a. [95] Fuel production
min)
179 intermittent shreds 20 kg/run N2 300-500 30-42 400°C [96] Process evaluation
0.5 intermittent shreds (10-30 mm) 0.2 kg/run N2 (0.1 550-800 45-65 400°C [10] Product analysis
dm3 /min)
(continued on next page)
Progress in Energy and Combustion Science 85 (2021) 100915
Table A1 (continued)

Reactor type Reactor Process Feed material Feed Rate Inert gas / flow Temperature TPO Yield [%] Conditions for Reference Manuscript objective
volume rate range [°C] optimum TPO yield
[dm3 ]

moving bed 0.15 continuous shreds (2 mm) 6 kg/h N2 (1.2 600 54 n.a. [97] Process evaluation
dm3 /min)
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al.

n.a. continuous shreds (5 mm), 3.5 - 8 kg/h N2 (0.2 600-800 28-48 n.a. [98] Process evaluation
steel and fabric free dm3 /min)
vacuum 0.1 intermittent shreds 0.1 kg/run -4 kPa vacuum 450-600 33-50 480°C [69] Process evaluation
n.a. intermittent shreds (6-13 mm) 1 kg/run -3 kPa vacuum 240-500 10-55 415°C [99] Product analysis
848 continuous n.a. 19 kg -13 kPa vacuum 500 50 n.a. [100] Fuel production
848 continuous n.a. n.a. -80 kPa vacuum 480-500 n.a. n.a. [101] Product analysis
n.a. intermittent shreds 1 kg/run -70 kPa vacuum 500 n.a. n.a. [102] Product analysis
848 n.a. shreds / whole tyres n.a. -10 kPa vacuum 480-520 43-56 n.a. [103] Fuel production
n.a. n.a. n.a. n.a. n.a. 400-600 30-53 n.a. [104] Process evaluation
n.a. n.a. n.a. n.a. n.a. 500-700 n.a. n.a. [105] Fuel production
40640 intermittent shreds 10000 kg/run n.a. 40-45 n.a. [106] Fuel production
not specified n.a. intermittent (30 ground 0.004 kg/run N2 (0.075 350-700 18-56 550-575°C [107] Process evaluation
min) dm3 /min)

37
n.a. n.a. rubber wastes from n.a. He (0.1 n.a. n.a. n.a. [52] Product analysis
tyre production dm3 /min)
n.a. intermittent (60 n.a. n.a. N2 (0.025 500 n.a. n.a. [108] Process evaluation
min), Pt-supported dm3 /min)
catalysts (zeolites)
n.a. intermittent (180 ground 0.04 kg/run N2 (0.03 400-600 28-41 400°C [109] Product analysis
min) dm3 /min)
spouted bed 11.2 continuous ground, steel free 0.2 kg/h n.a. 425-600 44-55 n.a. [110] Product analysis
n.a. continuous shreds (2.8-3.3 0.1 kg/h n.a. 425-575 54-58 n.a. [58] Process evaluation
mm), steel and
fabric free
28.3 intermittent, ground, steel free n.a. n.a. 450 40 n.a. [71] Fuel production
catalysts (zeolites)
0.16 intermittent (90 ground, steel and 0.12 kg/run H2 450-600 10-37 550°C [111] Product analysis
min), catalysts fabric free
(zeolites)
rotary (auger) n.a. continuous shreds (1-2 mm), 0.3 kg/h N2 500-600 39-51 n.a. [112] Process evaluation
reactor steel and fabric free
n.a. n.a. shreds (5 mm), n.a. n.a. n.a. n.a. n.a. [113] Fuel production
steel free
Progress in Energy and Combustion Science 85 (2021) 100915
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

References [30] Cheng JJ, Timilsina GR. Status and barriers of advanced biofuel technologies:
A review. Renew Energy 2011;36:3541–9. doi:10.1016/j.renene.2011.04.031.
[1] BP Energy Outlook 2019 edition 2019:144. www.bp.com/content/dam/bp/ [31] Lam SS, Chase HA. A review on waste to energy processes using microwave
business-sites/en/global/corporate/pdfs/energy-economics/energy-outlook/ pyrolysis. Energies 2012;5:4209–32. doi:10.3390/en5104209.
bp- energy- outlook- 2019.pdf (accessed October 15, 2020). [32] Oil World AnnualGlobal Analysis of all major oilseeds, oils and oil meals sup-
[2] Energy demand: Three drivers. ExxonMobil; 2019 https://corporate.exxonmobil. ply, demand and price outlook. Oil World Annu 2018;2. https://www.oilworld.
com:443/Energy- and- environment/Looking- forward/Outlook- for- Energy/ biz/t/publications/annual.
Energy-demand%0Ahttps:// corporate.exxonmobil.com/ Energy- and- environment/ [33] Ikura M, Stanciulescu M, Hogan E. Emulsification of pyrolysis derived
Looking- forward/Outlook- for- Energy/Energy- demand#threeDrivers access 12 bio-oil in diesel fuel. Biomass and Bioenergy 2003;24:221–32. doi:10.1016/
Oct. 2019. (accessed October 15, 2020). S0961- 9534(02)00131- 9.
[3] Hunicz J, Matijošius J, Rimkus A, Kilikevičius A, Kordos P, Mikulski M. Effi- [34] ASTM D6751-20a Standard Specification for Biodiesel Fuel Blend Stock (B100) for
cient hydrotreated vegetable oil combustion under partially premixed con- Middle Distillate Fuels. PA: West Conshohocken; 2020. doi:10.1520/D6751-20A.
ditions with heavy exhaust gas recirculation. Fuel 2020;268:117350. doi:10. [35] ASTM D975-20c Standard Specification for Diesel Fuel. PA: West Conshohocken;
1016/j.fuel.2020.117350. 2020. doi:10.1520/D0975-20C.
[4] Zhang Z, Lohr L, Escalante C, Wetzstein M. Food versus fuel: What do [36] DIN 51605, Fuels for vegetable oil compatible combustion engines - Fuel from
prices tell us? Energy Policy 2010;38:445–51. doi:10.1016/j.enpol.2009.09. rapeseed oil - Requirements and test methods. Hamburg, Germany: 2020.
034. [37] ISO 8217:2017 Petroleum products — Fuels (class F). Geneva, Switzerland: Spec-
[5] Chen R, Lun L, Cong K, Li Q, Zhang Y. Insights into pyrolysis and co- ifications of marine fuels; 2017.
pyrolysis of tobacco stalk and scrap tire: Thermochemical behaviors, kinetics, [38] EN 590 Automotive fuels - Diesel - Requirements and test methods. Brussels,
and evolved gas analysis. Energy 2019;183:25–34. doi:10.1016/j.energy.2019. Belgium; 2009.
06.127. [39] Ambrosewicz-Walacik M, Wierzbicki S, Mikulski M, Podciborski T. Ternary
[6] Cossu R, Lai T. Automotive shredder residue (ASR) management: An overview. fuel mixture of diesel, rapeseed oil and tyre pyrolytic oil suitable for modern
Waste Manag 2015;45:143–51. doi:10.1016/j.wasman.2015.07.042. CRDI engines. Transport 2018;33:727–40. doi:10.3846/transport.2018.5163.
[7] De SK, White JR. Rubber technologist’s handbook. iSmithers Rapra Publishing. [40] Laboy-Nieves ENEnergy recovery from scrap tires: A sustainable option
1st ed; 2001. for small islands like Puerto Rico. Sustain 2014;6:3105–21. doi:10.3390/
[8] Torretta V, Rada EC, Ragazzi M, Trulli E, Istrate IA, Cioca LI. Treatment and su6053105.
disposal of tyres: Two EU approaches. A review. Waste Manag 2015;45:152– [41] Mikulski M, Ramesh S, Bekdemir C. Reactivity Controlled Compression Igni-
60. doi:10.1016/j.wasman.2015.04.018. tion for clean and efficient ship propulsion. Energy 2019;182:1173–92. doi:10.
[9] Antoniou N, Zabaniotou A. Features of an efficient and environmentally at- 1016/j.energy.2019.06.091.
tractive used tyres pyrolysis with energy and material recovery. Renew Sus- [42] International Maritime Organization The 2020 Global Sulphur Limit; 2016.
tain Energy Rev 2013;20:539–58. doi:10.1016/j.rser.2012.12.005. S1084-9521(07)0 0 085-7 [pii]\r10.1016/j.semcdb.20 07.06.0 03.
[10] Kumar Singh R, Ruj B, Jana A, Mondal S, Jana B, Kumar Sadhukhan A, et al. [43] DiesielNet, Eco Point Inc. Emission Standards 2020. https://dieselnet.com/
Pyrolysis of three different categories of automotive tyre wastes: Product standards/(accessed October 10, 2019).
yield analysis and characterization. J Anal Appl Pyrolysis 2018;135:379–89. [44] Czajczyńska D, Krzyżyńska R, Jouhara H, Spencer N. Use of pyrolytic gas from
doi:10.1016/j.jaap.2018.08.011. waste tire as a fuel: A review. Energy 2017;134:1121–31. doi:10.1016/j.energy.
[11] Yazdani E, Hashemabadi SH, Taghizadeh A. Study of waste tire pyrolysis in 2017.05.042.
a rotary kiln reactor in a wide range of pyrolysis temperature. Waste Manag [45] Verma P, Zare A, Jafari M, Bodisco TA, Rainey T, Ristovski ZD, et al. Diesel
2019;85:195–201. doi:10.1016/j.wasman.2018.12.020. engine performance and emissions with fuels derived from waste tyres. Sci
[12] Xu S, Lai D, Zeng X, Zhang L, Han Z, Cheng J, et al. Pyrolysis characteristics of Rep 2018;8:2457. doi:10.1038/s41598- 018- 19330- 0.
waste tire particles in fixed-bed reactor with internals. Carbon Resour Convers [46] Leung DYC, Wang CL. Kinetic study of scrap tyre pyrolysis and combustion. J
2018;1:228–37. doi:10.1016/j.crcon.2018.10.001. Anal Appl Pyrolysis 1998;45:153–69. doi:10.1016/S0165-2370(98)0 0 065-5.
[13] Żółtowski A. Tyre Pyrolysis Oil As an Engine Fuel. J KONES Powertrain Transp [47] Murugan S, Ramaswamy MC, Nagarajan G. Performance, emission and com-
2014;21:295–302. doi:10.5604/12314005.1134118. bustion studies of a DI diesel engine using Distilled Tyre pyrolysis oil-diesel
[14] Žvar Baškovič U, Vihar R, Seljak T, Katrašnik T. Feasibility analysis of 100% blends. Fuel Process Technol 2008;89:152–9. doi:10.1016/j.fuproc.20 07.08.0 05.
tire pyrolysis oil in a common rail Diesel engine. Energy 2017;137:980–90. [48] Williams PT, Besler S, Klinar D, Likozar BPyrolysis- thermogravimetric analy-
doi:10.1016/j.energy.2017.01.156. sis of tires[1]Lah B. Pyrolysis of natural, butadiene, styrene-butadiene rubber
[15] Doǧan O, Elik MB, Özdalyan B. The effect of tire derived fuel/diesel and tyre components: Modelling kinetics and transport phenomena at differ-
fuel blends utilization on diesel engine performance and emissions. Fuel ent heating rates and formulations. Chem Eng S. Fuel 1995;74:1277–83.
2012;95:340–6. doi:10.1016/j.fuel.2011.12.033. [49] Juma M, Koreňová Z, Markoš J, Jelemensky L, Bafrnec M. Experimental study
[16] Chwist M, Grab-Rogaliński K, Szwaja S. Pyrolysis oil combustion in the CI en- of pyrolysis and combustion of scrap tire. Polym Adv Technol 2007;18:144–8.
gine. Combust Engines 2019;179:126–31. doi:10.19206/CE- 2019- 420. doi:10.1002/pat.811.
[17] Sharma A, Murugan S. Effect of blending waste tyre derived fuel on oxida- [50] Pradhan D, Singh RK. Thermal Pyrolysis of Bicycle Waste Tyre Using Batch
tion stability of biodiesel and performance and emission studies of a diesel Reactor. Int J Chem Eng Appl 2011;2:332–6. doi:10.7763/ijcea.2011.v2.129.
engine. Appl Therm Eng 2017;118:365–74. doi:10.1016/j.applthermaleng.2017. [51] Berrueco C, Esperanza E, Mastral FJ, Ceamanos J, García-Bacaicoa P. Pyrolysis
03.008. of waste tyres in an atmospheric static-bed batch reactor: Analysis of the
[18] Mikulski M, Ambrosewicz-Walacik M, Duda K, Hunicz J. Performance and gases obtained. J Anal Appl Pyrolysis 2005;74:245–53. doi:10.1016/j.jaap.2004.
emission characterization of a common-rail compression-ignition engine fu- 10.007.
elled with ternary mixtures of rapeseed oil, pyrolytic oil and diesel. Renew [52] Seidelt S, Müller-Hagedorn M, Bockhorn H. Description of tire pyrolysis by
Energy 2020;148:739–55. doi:10.1016/j.renene.2019.10.161. thermal degradation behaviour of main components. J Anal Appl Pyrolysis
[19] Scott E, ETRma - European Tyre & Rubber manufacturers’ association. End-of- 2006;75:11–18. doi:10.1016/j.jaap.2005.03.002.
life Tyre Report. 2015. [53] Han J, Li W, Liu D, Qin L, Chen W, Xing F. Pyrolysis characteristic and mech-
[20] ETRmaEuropean Tyre & Rubber manufacturers’ association. ELT Managment anism of waste tyre: A thermogravimetry-mass spectrometry analysis. J Anal
report 2018 2019. Appl Pyrolysis 2018;129:1–5. doi:10.1016/j.jaap.2017.12.016.
[21] World Business Council for Sustainable Development Global ELT Management [54] Kumaravel ST, Murugesan A, Kumaravel A. Tyre pyrolysis oil as an alternative
–A global state of knowledge on regulation,management systems. Geneva: im- fuel for diesel engines – A review. Renew Sustain Energy Rev 2016;60:1678–
pacts of recoveryand technologies; 2019. 85. doi:10.1016/j.rser.2016.03.035.
[22] ETRma European Tyre & Rubber manufacturers’ association. End of life tyres; [55] Ding K, Zhong Z, Zhang B, Song Z, Qian X. Pyrolysis Characteristics of Waste
2011. Tire in an Analytical Pyrolyzer Coupled with Gas Chromatography/Mass
[23] Sebola MR, Mativenga PT, Pretorius J. A Benchmark Study of Waste Tyre Recy- Spectrometry. Energy & Fuels 2015;29:3181–7. doi:10.1021/acs.energyfuels.
cling in South Africa to European Union Practice. Procedia CIRP 2018;69:950– 5b00247.
5. doi:10.1016/j.procir.2017.11.137. [56] Xu F, Wang B, Yang D, Ming X, Jiang Y, Hao J, et al. TG-FTIR and Py-GC/MS
[24] ETRma - European Tyre & Rubber manufacturers’ association. ELT Managment study on pyrolysis mechanism and products distribution of waste bicycle
report 2016. 2017. tire. Energy Convers Manag 2018;175:288–97. doi:10.1016/j.enconman.2018.09.
[25] ETRma - European Tyre & Rubber manufacturers’ association. ELT Managment 013.
report 2017. 2018. [57] Lopez G, Alvarez J, Amutio M, Mkhize NM, Danon B, van der Gryp P, et al.
[26] Dry ME. High quality diesel via the Fischer-Tropsch process - A review. J Chem Waste truck-tyre processing by flash pyrolysis in a conical spouted bed reac-
Technol Biotechnol 2002;77:43–50. doi:10.1002/jctb.527. tor. Energy Convers Manag 2017;142:523–32. doi:10.1016/j.enconman.2017.03.
[27] Cormier SA, Lomnicki S, Backes W, Dellinger B. Origin and health impacts 051.
of emissions of toxic by-products and fine particles from combustion and [58] Alvarez J, Lopez G, Amutio M, Mkhize NM, Danon B, van der Gryp P, et al.
thermal treatment of hazardous wastes and materials. Environ Health Perspect Evaluation of the properties of tyre pyrolysis oils obtained in a conical
2006;114:810–17. doi:10.1289/ehp.8629. spouted bed reactor. Energy 2017;128:463–74. doi:10.1016/j.energy.2017.03.
[28] Rowat SC. Incinerator toxic emissions: A brief summary of human health 163.
effects with a note on regulatory control. Med Hypotheses 1999;52:389–96. [59] Choi G-GG, Oh S-JJ, Kim J-SS. Non-catalytic pyrolysis of scrap tires using
doi:10.1054/mehy.1994.0675. a newly developed two-stage pyrolyzer for the production of a pyrolysis
[29] ETRma - European Tyre & Rubber manufacturers’ association. Europe - 92% oil with a low sulfur content. Appl Energy 2016;170:140–7. doi:10.1016/j.
of all End of Life Tyres collected and treated in 2017 2019:92–4. apenergy.2016.02.119.

38
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

[60] Undri A, Meini S, Rosi L, Frediani M, Frediani P. Microwave pyrolysis of poly- [88] Ucar S, Karagoz S, Ozkan AR, Yanik J. Evaluation of two different scrap tires
meric materials: Waste tires treatment and characterization of the value- as hydrocarbon source by pyrolysis. Fuel 2005;84:1884–92. doi:10.1016/j.fuel.
added products. J Anal Appl Pyrolysis 2013;103:149–58. doi:10.1016/j.jaap. 20 05.04.0 02.
2012.11.011. [89] Murillo R, Aylón E, Navarro MV, Callén MS, Aranda A, Mastral AM. The ap-
[61] Song Z, Yang Y, Zhao X, Sun J, Wang W, Mao Y, et al. Microwave pyrolysis plication of thermal processes to valorise waste tyre. Fuel Process Technol
of tire powders: Evolution of yields and composition of products. J Anal Appl 2006;87:143–7. doi:10.1016/j.fuproc.2005.07.005.
Pyrolysis 2017;123:152–9. doi:10.1016/j.jaap.2016.12.012. [90] Rofiqul Islam M, Haniu H. Rafiqul Alam Beg M. Liquid fuels and chemi-
[62] Isayev AI, Yushanov SP, Chen J. Ultrasonic devulcanization of rubber vulcan- cals from pyrolysis of motorcycle tire waste: Product yields, compositions
izates. II. Simulation and experiment. J Appl Polym Sci 1996;59:815–24. doi: and related properties. Fuel 2008;87:3112–22. doi:10.1016/j.fuel.2008.04.
10.1002/(sici)1097-4628(19960131)59:5⟨815::aid-app8⟩3.3.co;2-o. 036.
[63] Isayev AI, Yushanov SP, Kim SH, Levin VY. Ultrasonic devulcanization of waste [91] Acevedo B, Barriocanal C, Alvarez R. Pyrolysis of blends of coal and tyre
rubbers: Experimentation and modeling. Rheol Acta 1996;35:616–30. doi:10. wastes in a fixed bed reactor and a rotary oven. Fuel 2013;113:817–25.
10 07/BF0 0396511. doi:10.1016/j.fuel.2012.12.077.
[64] Sun X, Isayev AI. Continuous ultrasonic devulcanization: Comparison of car- [92] Choi G-G, Oh S-J, Kim J-S. Intermediate pyrolysis of scrap tires in a fixed bed
bon black filled synthetic isoprene and natural rubbers. Rubber Chem Technol reactor and activation of the pyrolysis char using CO 2. Characteristics of py-
2008;81:19–46. doi:10.5254/1.3548195. rolysis products and activated char; 2013. p. 2–3.
[65] Williams PT, Brindle AJ. Catalytic pyrolysis of tyres: Influence of catalyst tem- [93] Al Mamun MR, Halder PK, Hasan MR, Hossain MI, Khan MZH. Fuel properties
perature. Fuel 2002;81:2425–34. doi:10.1016/S0016-2361(02)00196-5. of pyrolytic tyre oil and its blends with diesel fuel - towards waste man-
[66] Xie QCKC, Bao WR, Wei-Huang Zhao JB. Pyrolysis of Waste Tires with Copper agement. Int J Environ Waste Manag 2016;18:335. doi:10.1504/ijewm.2016.
Nitrate. Energy Sources 2004;26:397–407. doi:10.1080/00908310490281519. 10 0 02722.
[67] Qu W, Zhou Q, Wang YZ, Zhang J, Lan WW, Wu YH, et al. Pyrolysis of waste [94] Hopa DY, Yilmaz A, Bahtli TA. Recovery of waste tyres by pyrolysis in a fixed
tire on ZSM-5 zeolite with enhanced catalytic activities. Polym Degrad Stab bed reactor for liquid fuel production: effects of pyrolysis conditions on oil
2006;91:2389–95. doi:10.1016/j.polymdegradstab.2006.03.014. yield. Res Eng Struct Mater 2017. doi:10.17515/resm2016.58en0701.
[68] Shah J, Rasul Jan M, Mabood F. Catalytic pyrolysis of waste tyre [95] Odejobi OJ, Sanda O, Abegunrin IO, Oladunni AA, Sonibare JA. Production of
rubber into hydrocarbons via base catalysts. Iran J Chem Chem Eng pyrolysis oil from used tyres and the effects of pyrolysis oil-gasoline blends
2008;27:103–9. on the performance of a gasoline-powered electric generator. Sci African
[69] Zhang X, Wang T, Ma L, Chang J. Vacuum pyrolysis of waste tires with 2020;10:e00639. doi:10.1016/j.sciaf.2020.e00639.
basic additives. Waste Manag 2008;28:2301–10. doi:10.1016/j.wasman.2007. [96] Aziz MA, Rahman MA, Molla H. Design, fabrication and performance test of
10.009. a fixed bed batch type pyrolysis plant with scrap tire in Bangladesh. J Radiat
[70] Kar Y. Catalytic pyrolysis of car tire waste using expanded perlite. Waste Res Appl Sci 2018;11:311–16. doi:10.1016/j.jrras.2018.05.001.
Manag 2011;31:1772–82. doi:10.1016/j.wasman.2011.04.005. [97] Aylón E, Fernández-Colino A, Navarro MV, Murillor R, García T, Mastral AM.
[71] Ayanoğlu A, Yumrutaş R. Production of gasoline and diesel like fuels from Waste tire pyrolysis: Comparison between fixed bed reactor and moving bed
waste tire oil by using catalytic pyrolysis. Energy 2016;103:456–68. doi:10. reactor. Ind Eng Chem Res 2008;47:4029–33. doi:10.1021/ie071573o.
1016/j.energy.2016.02.155. [98] Aylón E, Fernández-Colino A, Murillo R, Navarro MV, García T, Mastral AM.
[72] Fahed Banihani F, Bani Hani ZF. The Effect of Catalyst Ratio on the Pyrol- Valorisation of waste tyre by pyrolysis in a moving bed reactor. Waste Manag
ysis Yields for Waste Tyre. Am J Chem Eng 2018;6:60. doi:10.11648/j.ajche. 2010;30:1220–4. doi:10.1016/j.wasman.2009.10.001.
20180604.14. [99] Roy C, Labrecque B, de Caumia B. Recycling of scrap tires to oil and carbon
[73] Miandad R, Barakat MA, Rehan M, Aburiazaiza AS, Gardy J, Nizami AS. Ef- black by vacuum pyrolysis. Resour Conserv Recycl 1990;4:203–13. doi:10.1016/
fect of advanced catalysts on tire waste pyrolysis oil. Process Saf Environ Prot 0921-3449(90)90 0 02-L.
2018;116:542–52. doi:10.1016/j.psep.2018.03.024. [100] Benallal B, Roy C, Pakdel H, Chabot S, Poirier MA. Characterization of
[74] Debek
˛ C, Walendziewski J. Hydrorefining of oil from pyrolysis of whole tyres pyrolytic light naphtha from vacuum pyrolysis of used tyres comparison
for passenger cars and vans. Fuel 2015;159:659–65. doi:10.1016/j.fuel.2015.07. with petroleum naphtha. Fuel 1995;74:1589–94. doi:10.1016/0016-2361(95)
024. 00165-2.
[75] Ryms M, Januszewicz K, Lewandowski WM, Klugmann-Radziemska E. Pyrol- [101] Chaala A, Roy C. Production of coke from scrap tire vacuum pyrolysis
ysis Process of Whole Waste Tires as A Biomass Energy Recycling /Piroliza oil. Fuel Process Technol 1996;46:227–39. doi:10.1016/0378-3820(95)0 0 065-
Opon Samochodowych Jako Energetyczny Recykling Biomasy. Ecol Chem Eng 8.
S 2013;20:93–107. doi:10.2478/eces- 2013- 0 0 07. [102] Roy C, Darmstadt H, Benallal B. Amen-Chen C. Characterization of naphtha
[76] Lewandowski WM, Januszewicz K, Kosakowski W. Efficiency and proportions and carbon black obtained by vacuum pyrolysis of polyisoprene rubber. Fuel
of waste tyre pyrolysis products depending on the reactor type—A review. J Process Technol 1997;50:87–103. doi:10.1016/S0378-3820(96)01044-2.
Anal Appl Pyrolysis 2019;140:25–53. doi:10.1016/j.jaap.2019.03.018. [103] Roy C, Chaala A, Darmstadt H. The vacuum pyrolysis of used tires. J Anal Appl
[77] Lee JMJS, Lee JMJS, Kim JR, Kim SD. Pyrolysis of waste tires with partial Pyrolysis 1999;51:201–21. doi:10.1016/s0165-2370(99)0 0 017-0.
oxidation in a fluidized-bed reactor. Energy 1995;20:969–76. doi:10.1016/ [104] Rombaldo CFS, Lisbôa ACL, Méndez MOA, dos Reis Coutinho A. Effect of op-
0360-5442(95)0 0 049-M. erating conditions on scrap tire pyrolysis. Mater Res 2008;11:359–63. doi:10.
[78] Kaminsky W, Mennerich C. Pyrolysis of synthetic tire rubber in a fluidised- 1590/S1516-143920 080 0 030 0 021.
bed reactor to yield 1,3-butadiene, styrene and carbon black. J Anal Appl Py- [105] Vihar R, Seljak T, Rodman Oprešnik S, Katrašnik T. Combustion characteris-
rolysis 2001;58–59:803–11. doi:10.1016/S0165-2370(0 0)0 0129-7. tics of tire pyrolysis oil in turbo charged compression ignition engine. Fuel
[79] Dai X, Yin X, Wu C, Zhang W, Chen Y. Pyrolysis of waste tires in a circulating 2015;150:226–35. doi:10.1016/j.fuel.2015.01.087.
fluidized-bed reactor. Energy 2001;26:385–99. doi:10.1016/S0360-5442(01) [106] Tudu K, Murugan S, Patel SK. Effect of tyre derived oil-diesel blend on the
0 0 0 03-2. combustion and emissions characteristics in a compression ignition engine
[80] Kaminsky W, Mennerich C, Zhang Z. Feedstock recycling of synthetic and nat- with internal jet piston geometry. Fuel 2016;184:89–99. doi:10.1016/j.fuel.
ural rubber by pyrolysis in a fluidized bed. J Anal Appl Pyrolysis 2009;85:334– 2016.06.065.
7. doi:10.1016/j.jaap.2008.11.012. [107] González JF, Encinar JM, Canito JL, Rodríguez JJ. Pyrolysis of automobile tyre
[81] Edwin Raj R, Robert Kennedy Z, Pillai BC. Optimization of process parameters waste. Influence of operating variables and kinetics study. J Anal Appl Pyrolysis
in flash pyrolysis of waste tyres to liquid and gaseous fuel in a fluidized bed 2001;58:667–83. doi:10.1016/S0165-2370(0 0)0 0201-1.
reactor. Energy Convers Manag 2013;67:145–51. doi:10.1016/j.enconman.2012. [108] Dung˜ NA, Wongkasemjit S, Jitkarnka S. Effects of pyrolysis temperature and
11.012. Pt-loaded catalysts on polar-aromatic content in tire-derived oil. Appl Catal B
[82] Zhang H, Nie J, Xiao R, Jin B, Dong C, Xiao G. Catalytic co-pyrolysis of Environ 20 09;91:30 0–7. doi:10.1016/j.apcatb.2009.05.038.
biomass and different plastics (polyethylene, polypropylene, and polystyrene) [109] Czajczyńska D, Czajka K, Krzyżyńska R, Jouhara H. Waste tyre pyrolysis – Im-
to improve hydrocarbon yield in a fluidized-bed reactor. Energy and Fuels pact of the process and its products on the environment. Therm Sci Eng Prog
2014;28:1940–7. doi:10.1021/ef4019299. 2020;20:100690. doi:10.1016/j.tsep.2020.100690.
[83] Wang WC, Jan JJ. From laboratory to pilot: Design concept and techno- [110] López G, Olazar M, Aguado R, Bilbao J. Continuous pyrolysis of waste tyres in
economic analyses of the fluidized bed fast pyrolysis of biomass. Energy a conical spouted bed reactor. Fuel 2010;89:1946–52. doi:10.1016/j.fuel.2010.
2018;155:139–51. doi:10.1016/j.energy.2018.05.012. 03.029.
[84] Mastral AM, Murillo R, Callen MS, Garcia T. Optimisation of scrap automotive [111] Ramirez-Canon A, Muñoz-Camelo Y, Singh P. Decomposition of Used Tyre
tyres recycling into valuable liquid fuels. Resour Conserv Recycl 20 0 0;29:263– Rubber by Pyrolysis: Enhancement of the Physical Properties of the Liq-
72. doi:10.1016/S0921-3449(0 0)0 0 051-3. uid Fraction Using a Hydrogen Stream. Environments 2018;5:72. doi:10.3390/
[85] Díez C, Martínez O, Calvo LF, Cara J, Morán A. Pyrolysis of tyres. Influence of environments5060072.
the final temperature of the process on emissions and the calorific value of [112] Choi GG, Oh SJ, Kim JS. Scrap tire pyrolysis using a new type two-stage py-
the products recovered. Waste Manag 2004;24:463–9. doi:10.1016/j.wasman. rolyzer: Effects of dolomite and olivine on producing a low-sulfur pyrolysis
20 03.11.0 06. oil. Energy 2016;114:457–64. doi:10.1016/j.energy.2016.08.020.
[86] De Marco Rodríguez I, Laresgoiti MF, Cabrero MA, Torres A, Chomón MJ, Ca- [113] Martínez JD, Veses A, Mastral AM, Murillo R, Navarro MV, Puy N, et al. Co-
ballero B. Pyrolysis of scrap tyres. Fuel Process Technol 2001;72:9–22. doi:10. pyrolysis of biomass with waste tyres: Upgrading of liquid bio-fuel. Fuel Pro-
1016/S0378-3820(01)00174-6. cess Technol 2014;119:263–71. doi:10.1016/j.fuproc.2013.11.015.
[87] Choi GG, Jung SH, Oh SJ, Kim JS. Total utilization of waste tire rubber through [114] Martínez JD, Puy N, Murillo R, García T, Navarro MV, Mastral AM. Waste tyre
pyrolysis to obtain oils and CO2 activation of pyrolysis char. Fuel Process Tech- pyrolysis - A review. Renew Sustain Energy Rev 2013;23:179–213. doi:10.1016/
nol 2014;123:57–64. doi:10.1016/j.fuproc.2014.02.007. j.rser.2013.02.038.

39
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

[115] Luo S, Feng Y. The production of fuel oil and combustible gas by catalytic [143] Gasoil Price Outlook. FocusEconomics; 2020 www.focus-economics.com/
pyrolysis of waste tire using waste heat of blast-furnace slag. Energy Convers commodities/ energy/ gasoil (accessed November 10, 2020).
Manag 2017;136:27–35. doi:10.1016/j.enconman.2016.12.076. [144] Chen R, Qin Z, Han J, Wang M, Taheripour F, Tyner W, et al. Life cycle energy
[116] Laresgoiti MF, De Marco I, Torres A, Caballero B, Cabrero MA, Chomón MJ. and greenhouse gas emission effects of biodiesel in the United States with
Chromatographic analysis of the gases obtained in tyre pyrolysis. J Anal Appl induced land use change impacts. Bioresour Technol 2018;251:249–58. doi:10.
Pyrolysis 20 0 0;55:43–54. doi:10.1016/S0165-2370(99)0 0 073-X. 1016/j.biortech.2017.12.031.
[117] Laresgoiti MF, Caballero BM, De Marco I, Torres A, Cabrero MA, Chomón MJ. [145] Banar M. Life cycle assessment of waste tire pyrolysis. Fresenius Environ Bull
Characterization of the liquid products obtained in tyre pyrolysis. J Anal Appl 2015;24:1215–26.
Pyrolysis 2004;71:917–34. doi:10.1016/j.jaap.2003.12.003. [146] Neste Corporation Neste Renewable Diesel Handbook; 2020 www.neste.com/
[118] Aydin H, Ilkiliç C. Optimization of fuel production from waste vehicle tires by sites/ default/ files/ attachments/ neste_renewable_diesel_handbook.pdf (accessed
pyrolysis and resembling to diesel fuel by various desulfurization methods. October 13, 2020).
Fuel 2012;102:605–12. doi:10.1016/j.fuel.2012.06.067. [147] Arabiourrutia M, Lopez G, Elordi G, Olazar M, Aguado R, Bilbao J. Product dis-
[119] Cunliffe AM, Williams PT. Properties of chars and activated carbons de- tribution obtained in the pyrolysis of tyres in a conical spouted bed reactor.
rived from the pyrolysis of used tyres. Environ Technol (United Kingdom) Chem Eng Sci 2007;62:5271–5. doi:10.1016/j.ces.2006.12.026.
1998;19:1177–90. doi:10.1080/09593331908616778. [148] López FA, Centeno TA, Alguacil FJ, Lobato B. Distillation of granulated scrap
[120] Nisar J, Ali G, Ullah N, Awan IA, Iqbal M, Shah A, et al. Pyrolysis of waste tires in a pilot plant. J Hazard Mater 2011;190:285–92. doi:10.1016/j.jhazmat.
tire rubber: Influence of temperature on pyrolysates yield. J Environ Chem Eng 2011.03.039.
2018;6:3469–73. doi:10.1016/j.jece.2018.05.021. [149] Napoli A, Soudais Y, Lecomte D, Castillo S. Scrap tyre pyrolysis: Are the efflu-
[121] Galvagno S, Casu S, Casabianca T, Calabrese A, Cornacchia G. Pyrolysis pro- ents valuable products? J Anal Appl Pyrolysis 1997;40–41:373–82. doi:10.1016/
cess for the treatment of scrap tyres: Preliminary experimental results. Waste s0165-2370(97)0 0 011-9.
Manag 2002;22:917–23. doi:10.1016/S0956- 053X(02)00083- 1. [150] Celauro B, Celauro C, Lo Presti D, Bevilacqua A. Definition of a laboratory
[122] Mastral AM, Murillo R, García T, Navarro MV, Callen MS, López JM. Study of optimization protocol for road bitumen improved with recycled tire rub-
the viability of the process for hydrogen recovery from old tyre oils. Fuel Pro- ber. Constr Build Mater 2012;37:562–72. doi:10.1016/j.conbuildmat.2012.07.
cess Technol 2002;75:185–99. doi:10.1016/S0378- 3820(02)00004- 8. 034.
[123] Haydary J, Jelemenský Ľ, Gašparovič L, Markoš J. Influence of particle size [151] Moreno-Navarro F, Sol-Sánchez M, Rubio-Gámez MC. Reuse of deconstructed
and kinetic parameters on tire pyrolysis. J Anal Appl Pyrolysis 2012;97:73–9. tires as anti-reflective cracking mat systems in asphalt pavements. Constr
doi:10.1016/j.jaap.2012.07.003. Build Mater 2014;53:182–9. doi:10.1016/j.conbuildmat.2013.11.101.
[124] Mkhize NMM, van der Gryp P, Danon B, Görgens JFF. Effect of temperature [152] Shu X, Huang B. Recycling of waste tire rubber in asphalt and portland ce-
and heating rate on limonene production from waste tyre pyrolysis. J Anal ment concrete: An overview. Constr Build Mater 2014;67:217–24. doi:10.1016/
Appl Pyrolysis 2016;120:314–20. doi:10.1016/j.jaap.2016.04.019. j.conbuildmat.2013.11.027.
[125] Alsaleh A, Sattler ML. Waste Tire Pyrolysis: Influential Parameters and [153] Mui ELK, Ko DCK, McKay G. Production of active carbons from waste tyres -
Product Properties. Curr Sustain Energy Reports 2014;1:129–35. doi:10.1007/ A review. Carbon N Y 2004;42:2789–805. doi:10.1016/j.carbon.2004.06.023.
s40518- 014- 0019- 0. [154] Lopez G, Aguado R, Olazar M, Arabiourrutia M, Bilbao J. Kinetics of scrap tyre
[126] Leung DYC, Yin XL, Zhao ZL, Xu BY, Chen Y. Pyrolysis of tire powder: Influence pyrolysis under vacuum conditions. Waste Manag 2009;29:2649–55. doi:10.
of operation variables on the composition and yields of gaseous product. Fuel 1016/j.wasman.20 09.06.0 05.
Process Technol 2002;79:141–55. doi:10.1016/S0378-3820(02)00109-1. [155] Heras F, Jimenez-Cordero D, Gilarranz MA, Alonso-Morales N, Rodriguez JJ.
[127] Kan T, Strezov V, Evans T. Fuel production from pyrolysis of natural and syn- Activation of waste tire char by cyclic liquid-phase oxidation. Fuel Process
thetic rubbers. Fuel 2017;191:403–10. doi:10.1016/j.fuel.2016.11.100. Technol 2014;127:157–62. doi:10.1016/j.fuproc.2014.06.018.
[128] Hita I, Arabiourrutia M, Olazar M, Bilbao J, Arandes JM, Castaño P. Oppor- [156] Murillo R, Navarro MV, López JM, García T, Callén MS, Aylón E, et al. Ac-
tunities and barriers for producing high quality fuels from the pyrolysis of tivation of pyrolytic tire char with CO2: Kinetic study. J Anal Appl Pyrolysis
scrap tires. Renew Sustain Energy Rev 2016;56:745–59. doi:10.1016/j.rser.2015. 2004;71:945–57. doi:10.1016/j.jaap.2003.12.005.
11.081. [157] Kyari M, Cunliffe A, Williams PT. Characterization of oils, gases, and char in
[129] Abnisa F, Wan Daud WMA. A review on co-pyrolysis of biomass: An op- relation to the pyrolysis of different brands of scrap automotive tires. Energy
tional technique to obtain a high-grade pyrolysis oil. Energy Convers Manag and Fuels 2005;19:1165–73. doi:10.1021/ef049686x.
2014;87:71–85. doi:10.1016/j.enconman.2014.07.007. [158] Fernández AM, Barriocanal C, Alvarez R. Pyrolysis of a waste from the
[130] Abnisa F, Wan Daud WMA. Optimization of fuel recovery through the grinding of scrap tyres. J Hazard Mater 2012;203–204:236–43. doi:10.1016/
stepwise co-pyrolysis of palm shell and scrap tire. Energy Convers Manag j.jhazmat.2011.12.014.
2015;99:334–45. doi:10.1016/j.enconman.2015.04.030. [159] Dȩbek C. Modification of pyrolytic oil from waste tyres as a promising
[131] Farooq MZ, Zeeshan M, Iqbal S, Ahmed N, Shah SAY. Influence of waste tire method for light fuel production. Materials (Basel) 2019;16:1–8. doi:10.3390/
addition on wheat straw pyrolysis yield and oil quality. Energy 2018;144:200– ma12060880.
6. doi:10.1016/j.energy.2017.12.026. [160] Williams PT, Bottrill RP. Sulfur-polycyclic aromatic hydrocarbons in tyre py-
[132] Kalghatgi G, Levinsky H, Colket M. Future transportation fuels. Prog Energy rolysis oil. Fuel 1995;74:736–42. doi:10.1016/0016-2361(94)0 0 0 05-C.
Combust Sci 2018;69:103–5. doi:10.1016/j.pecs.2018.06.003. [161] Cunliffe AM, Williams PT. Composition of oils derived from the batch pyroly-
[133] Ábrego J, Plaza D, Luño F, Atienza-Martínez M, Gea G. Pyrolysis of sis of tyres. J Anal Appl Pyrolysis 1998;44:131–52. doi:10.1016/S0165-2370(97)
cashew nutshells: Characterization of products and energy balance. Energy 0 0 085-5.
2018;158:72–80. doi:10.1016/j.energy.2018.06.011. [162] Li S-Q, Yao Q, Chi Y, Yan J-H, Cen K-F. Pilot-Scale Pyrolysis of Scrap Tires in
[134] Altayeb R, Ibrahim T, Jabbar NA, Aisan A. Liquid Fuel production from py- a Continuous Rotary Kiln Reactor. Ind Eng Chem Res 2004;43:5133–45. doi:10.
rolysis of waste tires: Process simulation & exergetic analysis, and life cycle 1021/ie030115m.
assessment. In: AES-ATEMA Int. Conf. Ser. - Adv. Trends Eng. Mater. their Appl.; [163] Rada EC, Ragazzi M, Dal Maschio R, Ischia M, Panaitescu VN. Energy recov-
2016. p. 65–77. ery from tyres waste through thermal option. UPB Sci Bull Ser D Mech Eng
[135] Silva VB, Rouboa A. Using a two-stage equilibrium model to simulate oxy- 2012;74:201–10.
gen air enriched gasification of pine biomass residues. Fuel Process Technol [164] Lotero E, Goodwin Jr J, Bruce A, Auwannakarn K, Liu Y, Lopez D. The Catalysis
2013;109:111–17. doi:10.1016/j.fuproc.2012.09.045. of. Biodiesel Synthesis. Spec Period Reports - Catal 2006;19:41–83. doi:10.1039/
[136] Karimi M. Exergy-based optimization of direct conversion of microalgae 9781847555229-0 0 041.
biomass to biodiesel. J Clean Prod 2017;141:50–5. doi:10.1016/j.jclepro.2016. [165] Altin R, Çetinkaya S, Yücesu HS. Potential of using vegetable oil fuels as
09.032. fuel for diesel engines. Energy Convers Manag 2001;42:529–38. doi:10.1016/
[137] Buadit T, Rattanapan C, Ussawarujikulchai A, Suchiva K, Papong S, Ma H. Life S0196-8904(0 0)0 0 080-7.
Cycle Assessment of Material Recovery from Pyrolysis Process of End-of-Life [166] Parthasarathy P, Choi HS, Park HC, Hwang JG, Yoo HS, Lee BK, et al. Influence
Tires in Thailand. Int J Environ Sci Dev 2020;11:488–98. doi:10.18178/ijesd. of process conditions on product yield of waste tyre pyrolysis- A review. Ko-
2020.11.10.1296. rean J Chem Eng 2016;33:2268–86. doi:10.1007/s11814- 016- 0126- 2.
[138] Li X, Xu H, Gao Y, Tao Y. Comparison of end-of-life tire treatment tech- [167] Nabi MN, Akhter MS, Shahadat MMZ. Improvement of engine emissions
nologies: A Chinese case study. Waste Manag 2010;30:2235–46. doi:10.1016/j. with conventional diesel fuel and diesel-biodiesel blends. Bioresour Technol
wasman.2010.06.006. 2006;97:372–8. doi:10.1016/j.biortech.2005.03.013.
[139] Shelley MD, El-Halwagi MM. Techno-Economic Feasibility and Flowsheet Syn- [168] Pietrzak R. XPS study and physico-chemical properties of nitrogen-enriched
thesis of Scrap Tire/Plastic Waste Liquefaction. J Elastomers Plast 1999;31:232– microporous activated carbon from high volatile bituminous coal. Fuel
54. doi:10.1177/009524439903100305. 2009;88:1871–7. doi:10.1016/j.fuel.2009.04.017.
[140] Hillairet F, Madiot O. Market Watch 2020. www.greenea.com/wp-content/ [169] Buyukkaya E. Effects of biodiesel on a di diesel engine performance, emis-
uploads/2020/11/Greenea- Market- Watch- November- 2020- _- EN- VOM- 2- 1. sion and combustion characteristics. Fuel 2010;89:3099–105. doi:10.1016/j.
pdf. fuel.2010.05.034.
[141] Palm and Rapeseed Oil Prices. Nestle; 2020 https:// www.neste.com/ investors/ [170] Telmo C, Lousada J, Moreira N. Proximate analysis, backwards stepwise re-
market-data/palm-and-rapeseed-oil-prices (accessed November 10, 2020). gression between gross calorific value, ultimate and chemical analysis of
[142] Tyre pyrolysis oil market price from waste tyre recycling pyroly- wood. Bioresour Technol 2010;101:3808–15. doi:10.1016/j.biortech.2010.01.021.
sis plant. DOING Holdings - Henan Doing Environ Prot Technol [171] Nautiyal P, Subramanian KA, Dastidar MG. Production and characterization
Co, Ltd; 2019 www.recyclingpyrolysisplant.com/ news/ industry_news/ of biodiesel from algae. Fuel Process Technol 2014;120:79–88. doi:10.1016/j.
tyre_pyrolysis_oil_marker _price_728.html (accessed November 10, 2020). fuproc.2013.12.003.

40
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

[172] DiesielNet, Eco Point Inc. Fuel Regulations, Automotive Diesel Fuel; EN 590 [199] Ahmad W, Ahmad I, Yaseen M. Desulfurization of liquid fuels by air assisted
1993. https://dieselnet.com/standards/eu/fuel_automotive.php (accessed Oc- peracid oxidation system in the presence of Fe-ZSM-5 catalyst. Korean J Chem
tober 1, 2019). Eng 2016;33:2530–7. doi:10.1007/s11814- 016- 0099- 1.
[173] Murugan S, Ramaswamy MC, Nagarajan G. The use of tyre pyrolysis oil in [200] Koc AB, Abdullah M. Performance of a 4-cylinder diesel engine running on
diesel engines. Waste Manag 2008;28:2743–9. doi:10.1016/j.wasman.2008.03. tire oil-biodiesel-diesel blend. Fuel Process Technol 2014;118:264–9. doi:10.
007. 1016/j.fuproc.2013.09.013.
[174] Murugan S, Ramaswamy MC, Nagarajan G. A comparative study on the per- [201] Sharma A, Murugan S. Investigation on the behaviour of a DI diesel engine
formance, emission and combustion studies of a DI diesel engine using dis- fueled with Jatropha Methyl Ester (JME) and Tyre Pyrolysis Oil (TPO) blends.
tilled tyre pyrolysis oil-diesel blends. Fuel 2008;87:2111–21. doi:10.1016/j.fuel. Fuel 2013;108:699–708. doi:10.1016/j.fuel.2012.12.042.
20 08.01.0 08. [202] Hariharan S, Murugan S, Nagarajan G. Effect of diethyl ether on Tyre pyroly-
[175] Islam MR, Parveen M, HANIU H, Sarker MRI. Innovation in Pyrolysis Technol- sis oil fueled diesel engine. Fuel 2013;104:109–15. doi:10.1016/j.fuel.2012.08.
ogy for Management of Scrap Tire: a Solution of Energyand Environment. Int 041.
J Environ Sci Dev 2010:89–96. doi:10.7763/ijesd.2010.v1.18. [203] Jeyakumar N, Narayanasamy B, John K, Kathiresh, Markus Solomon J. Prepa-
[176] Ilkiliç C, Aydin H. Fuel production from waste vehicle tires by catalytic pyrol- ration, characterization and effect of calcium carbonate and titanium diox-
ysis and its application in a diesel engine. Fuel Process Technol 2011;92:1129– ide nano additives on fuel properties of tire oil diesel blend. Energy Sources,
35. doi:10.1016/j.fuproc.2011.01.009. Part A Recover Util Environ Eff 2018;40:1798–806. doi:10.1080/15567036.2018.
[177] Sebola R, Pilusa J, Muzenda E. Characteristics of tyre derived fuel-diesel 1486919.
blends. In: 3rd Int. Conf. Med. Sci. Chem. Eng.; 2013. p. 27–31. [204] Lusvardi G, Barani C, Giubertoni F, Paganelli G. Synthesis and characterization
[178] Wongkhorsub C, Chindaprasert N. A Comparison of the Use of Pyrolysis of TiO2 nanoparticles for the reduction of water pollutants. Materials (Basel)
Oils in Diesel Engine. Energy Power Eng 2013;05:350–5. doi:10.4236/epe.2013. 2017;10:1208. doi:10.3390/ma10101208.
54b068. [205] Fangsuwannarak K, Triratanasirichai K. Improvements of Palm Biodiesel
[179] Frigo S, Seggiani M, Puccini M, Vitolo S. Liquid fuel production from waste Properties by Using Nano-TiO 2 Additive, Exhaust emission and Engine
tyre pyrolysis and its utilisation in a Diesel engine. Fuel 2014;116:399–408. Performance The Romanian Review Precision Mechanics. Opt Mechatronics
doi:10.1016/j.fuel.2013.08.044. 2013:111–18.
[180] Martínez JD, Rodríguez-Fernández J, Sánchez-Valdepeñas J, Murillo R, Gar- [206] Uyumaz A, Aydoğan B, Solmaz H, Yılmaz E, Yeşim Hopa D, Aksoy Bahtli T,
cía T. Performance and emissions of an automotive diesel engine using a et al. Production of waste tyre oil and experimental investigation on combus-
tire pyrolysis liquid blend. Fuel 2014;115:490–9. doi:10.1016/j.fuel.2013.07. tion, engine performance and exhaust emissions. J Energy Inst 2019;92:1406–
051. 18. doi:10.1016/j.joei.2018.09.001.
[181] Tudu K, Murugan S, Patel SK. Light oil fractions from a pyrolysis plant-An [207] Van De Beld B, Holle E, Florijn J. The use of pyrolysis oil and pyrolysis oil de-
option for energy use. Energy Procedia 2014;54:615–26. doi:10.1016/j.egypro. rived fuels in diesel engines for CHP applications. Appl Energy 2013;102:190–
2014.07.303. 7. doi:10.1016/j.apenergy.2012.05.047.
[182] Al-Lal AM, Bolonio D, Llamas A, Lapuerta M, Canoira L. Desulfurization [208] Kumar Singh R, Prabu M. Experimental investigation of a di diesel engine
of pyrolysis fuels obtained from waste: Lube oils, tires and plastics. Fuel using tyre pyrolysis oil-diesel blends as a biodiesel. Int J Mech Eng Technol
2015;150:208–16. doi:10.1016/j.fuel.2015.02.034. 2014;5:74–90.
[183] Islam MN, Nahian MR. Improvement of Waste Tire Pyrolysis Oil and Perfor- [209] Johnson T. Diesel engine emissions and their control: An overview. Platin Met
mance Test with Diesel in CI Engine. J Renew Energy 2016;2016:1–8. doi:10. Rev 2008;52:23–37. doi:10.1595/147106708X248750.
1155/2016/5137247. [210] Turns SR. Understanding NOx formation in nonpremixed flames: Experi-
[184] Wang WC, Bai CJ, Lin CT, Prakash S. Alternative fuel produced from thermal ments and modeling. Prog Energy Combust Sci 1995;21:361–85. doi:10.1016/
pyrolysis of waste tires and its use in a di diesel engine. Appl Therm Eng 0360-1285(94)0 0 0 06-9.
2016;93:330–8. doi:10.1016/j.applthermaleng.2015.09.056. [211] Bergthorson JM, Thomson MJ. A review of the combustion and emissions
[185] Ambrosewicz-Walacik M, Walacik M. Production of fuel blends from diesel properties of advanced transportation biofuels and their impact on existing
oil and waste products. Combust Engines 2017;171:255–8. doi:10.19206/ and future engines. Renew Sustain Energy Rev 2015;42:1393–417. doi:10.1016/
CE- 2017- 443. j.rser.2014.10.034.
[186] Bhatt PM, Prajapati A V. Experimental Investigation of Single Cylinder Diesel [212] Murugan S, Ramaswamy MRC, Nagarajan G. Influence of distillation on per-
Engine using Tyre Pyrolysis Oil (TPO) Blends 2017:212–7. formance, emission, and combustion of a di diesel engine, using tyre pyroly-
[187] Hürdoğan E, Ozalp C, Kara O, Ozcanli M. Experimental investigation on sis oil diesel blends. Therm Sci 2008;12:157–67. doi:10.2298/TSCI0801157M.
performance and emission characteristics of waste tire pyrolysis oil–diesel [213] Uslu S. Optimization of diesel engine performance and emission parame-
blends in a diesel engine. Int J Hydrogen Energy 2017;42:23373–8. doi:10. ters operating waste tire pyrolysis oil–diesel blends using response surface
1016/j.ijhydene.2016.12.126. methodology. Proc Inst Mech Eng Part I J Syst Control Eng 2019. doi:10.1177/
[188] Heywood JB. Internal Combustion Engine Fundamentals. 2nd edt. New York: 0959651819864851.
McGraw-Hill Education; 2018. [214] Bodisco TA, Rahman SMA, Hossain FM, Brown RJ. On-road NO x emissions of
[189] Santana RC, Do PT, Santikunaporn M, Alvarez WE, Taylor JD, Sughrue EL, a modern commercial light-duty diesel vehicle using a blend of tyre oil and
et al. Evaluation of different reaction strategies for the improvement of cetane diesel. Energy Reports 2019;5:349–56. doi:10.1016/j.egyr.2019.03.002.
number in diesel fuels. Fuel 2006;85:643–56. doi:10.1016/j.fuel.2005.08. [215] Khobragade R, Singh SK, Shukla PC, Gupta T, Al-Fatesh AS, Agarwal AK, et al.
028. Chemical composition of diesel particulate matter and its control. Catal Rev -
[190] Antoniou NA, Zorpas AA. Quality protocol and procedure development to de- Sci Eng 2019;61:447–515. doi:10.1080/01614940.2019.1617607.
fine end-of-waste criteria for tire pyrolysis oil in the framework of circular [216] Corro G. Sulfur impact on diesel emission control -a review. React Kinet Catal
economy strategy. Waste Manag 2019;95:161–70. doi:10.1016/j.wasman.2019. Lett 2002;75:89–106. doi:10.1023/A:1014853602908.
05.035. [217] Neeft JPA, Makkee M, Moulijn JA. Diesel particulate emission control. Fuel
[191] Petchsoongsakul N, Ngaosuwan K, Kiatkittipong W, Wongsawaeng D, Ass- Process Technol 1996;47:1–69. doi:10.1016/0378-3820(96)01002-8.
abumrungrat S. Different water removal methods for facilitating biodiesel [218] Russell A, Epling WS. Diesel oxidation catalysts. Catal Rev - Sci Eng
production from low-cost waste cooking oil containing high water content in 2011;53:337–423. doi:10.1080/01614940.2011.596429.
hybridized reactive distillation. Renew Energy 2020;162:1906–18. doi:10.1016/ [219] Prepared by ICF for the U.S Environmental Protection Agency Office of Re-
j.renene.2020.09.115. source Conservation and Recovery.. Tires: Documentation for Greenhouse Gas
[192] Cowley LT, Stradling RJ, Doyon J. The Influence of Composition and Proper- Emission and Energy Factors Used in the Waste Reduction Model (WARM);
ties of Diesel Fuel on Particulate Emissions from Heavy-Duty Engines 1993. 2019 https:// www.epa.gov/ sites/ production/ files/ 2019-06/ documents/ warm_v15_
10.4271/932732. tires.pdf (accessed November 10, 2020).
[193] Zhao B. Why will dominant alternative transportation fuels be liquid fu- [220] Corporation EM. ExxonMobil Outlook on energy 2020 2020. https:
els, not electricity or hydrogen? Energy Policy 2017;108:712–14. doi:10.1016/j. //corporate.exxonmobil.com/Energy- and- environment/Looking- forward/
enpol.2017.06.047. Outlook- for- Energy/(accessed November 17, 2020).
[194] Kalghatgi G. Fuel/Engine Interactions. Warrendale, PA: SAE International; 2013. [221] Fawcett S, Poole K, Kerley L, Vila B, Vari M, Papadimitriou V, et al. Tyre Re-
doi:10.4271/R-409. cycling Pyrolysis for producing oil with less than 0.2% sulphur content, low cost
[195] Shakirullah M, Ahmad I, Ahmad W, Ishaq M. Desulphurization study of sulphur impregnated carbon for reducing mercury air emissions, with simultane-
petroleum products through extraction with aqueous ionic liquids. J Chil ous elemental. Altrincham, UK; 2016.
Chem Soc 2010;55:179–83. doi:10.4067/S0717-97072010 0 0 020 0 0 07. [222] Kroyan Y, Wojcieszyk M, Larmi M, Kaario O, Zenger K. Modeling the Im-
[196] Hita I, Gutiérrez A, Olazar M, Bilbao J, Arandes JM, Castaño P. Upgrading pact of Alternative Fuel Properties on Light Vehicle Engine Performance and
model compounds and Scrap Tires Pyrolysis Oil (STPO) on hydrotreating Greenhouse Gases Emissions. SAE Tech Pap Ser 2019;1. doi:10.4271/2019-01-
NiMo catalysts with tailored supports. Fuel 2015;145:158–69. doi:10.1016/j. 2308.
fuel.2014.12.055. [223] Çolak M, Kaya İ. Prioritization of renewable energy alternatives by using an
[197] Hita I, Palos R, Arandes JM, Hill JM, Castaño P. Petcoke-derived functionalized integrated fuzzy MCDM model: A real case application for Turkey. Renew Sus-
activated carbon as support in a bifunctional catalyst for tire oil hydropro- tain Energy Rev 2017;80:840–53. doi:10.1016/j.rser.2017.05.194.
cessing. Fuel Process Technol 2016;144:239–47. doi:10.1016/j.fuproc.2015.12. [224] Erdoğan S, Balki MK, Aydın S, Sayin C. The best fuel selection with hybrid
030. multiple-criteria decision making approaches in a CI engine fueled with their
[198] Murena F. Kinetics of sulphur compounds in waste tyres pyrolysis. J Anal Appl blends and pure biodiesels produced from different sources. Renew Energy
Pyrolysis 20 0 0;56:195–205. doi:10.1016/S0165-2370(0 0)0 0 091-7. 2019;134:653–68. doi:10.1016/j.renene.2018.11.060.

41
M. Mikulski, M. Ambrosewicz-Walacik, J. Hunicz et al. Progress in Energy and Combustion Science 85 (2021) 100915

Maciej Mikulski (PhD) is an associate professor in com- Jacek Hunicz (PhD, DSc.) is an associate professor and
bustion engine technology at the University of Vaasa, Fin- head of Powertrains Laboratory at the Lublin University
land and a Principle Scientist in the VEBIC internal com- of Technology, Poland. His track record includes experi-
bustion engine laboratory (Finland). He has 14 years of mental engine research and renewable low-carbon fuels.
professional experience in combustion engine research in- In the area of combustion research, his studies are cen-
volving alternative fuels in industry and academia. He tered on the control strategies for low-temperature com-
has worked in several relevant engine/fuel development bustion in HCCI engines, including the NVO fuel reform-
projects with the world’s leading automotive, marine and ing. With over 23 years of professional expertise in com-
off-road OEMs and co-established the first full-load ca- bustion engines and powertrain development, prof Hunicz
pable, ultra-efficient RCCI natural gas engine. For his re- is a grant holder of several relevant nationwide projects
search on low temperature combustion, he was recently funded by the Ministry of Science and Higher Education
honored with the Finish Ostrobothnia Chamber of Com- and National Science Centre. He is a member of several
merce award. He is an author of over 80 research papers international research groups. He is also an innovation
in the field and a member of several international expert groups and combustion consultant for domestic off-road vehicle and bus manufacturer Ursus and techni-
research associations. He is currently (2020-2023), amongst others, a leader in 2 cal advisor to the Polish military industry in the field of powertrain testing. Since
major (over 25M€ in total founding) Finish Industry-Academia consortia the CLEAN 2018, he has been a board member of the Polish Scientific Society for Combustion
PROPULSION TECHNOLOGY and SILENT, both committed to advanced combustion Engines.
and powertrain development.
Szymon Nitkiewicz (PhD) obtained Master’s degree in
Marta Abrosewicz Walacik (PhD) is a professional Mathematics and graduated a PhD in Biocybernetics
chemist and a former head of the Biofuel Laboratory at and Biomedical Engineering. He is an owner of several
the University of Warmia and Mazury in Poland. Since the Patents, awarded at international exhibitions. His expe-
beginning of the carrier, she has focused on tyre pyroly- rience in mathematics allows him to analyze in detail a
sis and fuel upgrading. Over 35 published research papers wide range of issues. As part of Dr. Ambrosewicz’s team
involve research on TPO, FAME and crude oil refining for he specializes in assessing the impact of technology on
different applications, including combustion engine fuels. human health and supports the current work with his
She is responsible for critically assessing the analytical critical outlook in this respect. His scientific interests os-
works related to this review and gathering the informa- cillate in topics related to supporting medical diagnostics,
tion on the state of the art in pyrolysis process. as well as in assessing the impact of human activity on
the environment. He is a member of the project team In-
tereg Baltic Sea Region financed by the EU.

42

You might also like